Sie sind auf Seite 1von 10

Surface & Coatings Technology 277 (2015) 308317

Contents lists available at ScienceDirect

Surface & Coatings Technology


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

Electrochromic thin lms of sodium intercalated vanadium(V) oxide


xerogels: Chemical bath deposition and characterization
Metodija Najdoski a,b,, Violeta Koleva c, Sasho Stojkovikj a,b, Toni Todorovski a,1
a
b
c

Institute of Chemistry, Faculty of Natural Sciences and Mathematics, Ss. Cyril and Methodius University, POB 162, Arhimedova 3, 1000 Skopje, Republic of Macedonia
Research Center for Environment and Materials, Macedonian Academy of Sciences and Arts, Krste Misirkov 2, 1000 Skopje, Republic of Macedonia
Institute of General and Inorganic Chemistry, Bulgarian Academy of Sciences, G. Bonchev Str. Bldg. 11, 1113 Soa, Bulgaria

a r t i c l e

i n f o

Article history:
Received 26 February 2015
Revised 20 July 2015
Accepted in revised form 21 July 2015
Available online 29 July 2015
Keywords:
Thin lms
Vanadium(V) oxide xerogels
Chemical synthesis
Electrochromism
Optical properties
Electrochemical properties

a b s t r a c t
An optimized chemical bath method is applied to obtain well-structured thin lms with composition
Na0.33V2O5nH2O (n = 1 and 1.3). The method is based on a controlled precipitation reaction that takes place
in the system of sodium metavanadate and diethyl sulfate at 85 C. The lm structure, morphology and the
changes occurring during prolonged aging are examined by XRD, IR spectroscopy, TG-DTA, SEM and AFM. The
electrochemical and electrochromic properties are studied by cyclic voltammetry and UVvis spectroscopy.
The as-deposited thin lms are characterized with high optical transmittance varying between 40 and 70% at
the 500 nm visible region in dependence on lm thickness. The Na0.33V2O5nH2O thin lms exhibit stable electrochemical cycling combined with relatively high electrochromic activity. The reproducibility of the transmittance variance of 55% after 500 cycles in the electrochromic cell is a promising result for the potential
application of Na0.33V2O5nH2O thin lms in electrochromic devices.
2015 Published by Elsevier B.V.

1. Introduction
Vanadium(V) oxide and derived compounds have been extensively
studied due to valuable chemical and physical properties which determine a wide range of applications in catalysis, high-energy lithium batteries and a variety of electric and optical devices. The synthetic procedures
at ambient conditions usually produce hydrated vanadium(V) oxides,
V2O5nH2O, known as xerogels which adopt layered structures with
V2O5 layers and interstitial water molecules [13]. Vanadium(V) oxide
xerogels like crystalline V2O5 are typical intercalation compounds with
multiple valence state of vanadium which enables redox-dependent
properties [46]. Due to the high intercalation capacity (for instance, a
lithium intercalation capacity about 1.4 times larger than that of crystalline V2O5) [7] they have a great potential for applications like reversible
cathodes for lithium batteries [8,9], micro-batteries [6], supercapacitors
[10], electrodes [11] and humidity sensors [12]. The reversible cation
intercalation/deintercalation within the xerogel framework is concomitant with reversible reduction/oxidation of V(V) to V(IV) or to a lower
Corresponding author at: Institute of Chemistry, Faculty of Natural Sciences and
Mathematics, Ss. Cyril and Methodius University, POB 162, Arhimedova 3, 1000 Skopje,
Republic of Macedonia.
E-mail addresses: metonajd@pmf.ukim.mk (M. Najdoski), vkoleva@svr.igic.bas.bg
(V. Koleva), sashostojkovikj@gmail.com (S. Stojkovikj), toni.todorovski@irbbarcelona.org
(T. Todorovski).
1
Present address: Institute for Research in Biomedicine, Parc Cientc de Barcelona,
08028 Barcelona, Spain.

http://dx.doi.org/10.1016/j.surfcoat.2015.07.041
0257-8972/ 2015 Published by Elsevier B.V.

valence vanadium state giving rise to easy color changes: yellow


(V(V)), blue (V(IV)), green (V(III) or a mixture of V(V) and V(IV)) and
violet (V(II)) [2,6]. The multi-colored electrochromism demonstrated
by V2O5nH2O xerogels makes them very attractive since it provides
the opportunity to extend the range of functions of the electrochromic
materials. Vanadium(V) oxide xerogels under the form of thin lms on
electroconductive glass substrates have been used in electrochromic
devices [13,14], electrochromic mirrors [14], smart windows designed
for architectural purposes to control light transmittance [1517] and
controlled reectance mirrors for vehicles [18].
The thin lm properties, including V2O5nH2O xerogels, are well
known to depend essentially on its microscopic characteristics [19,20]
such as structure, crystallinity and morphology, which can be governed
by the deposition method and the deposition parameters (kind and
concentration of the precursors, rate of deposition, temperature, pressure, etc.). Therefore, the choice of the suitable synthetic procedure is
a powerful tool for control and optimization of the material properties.
In this regard, we have recently developed a simple chemical bath
deposition method to obtain well-dened ammonium intercalated
vanadium(V) oxide xerogels with the compositions (NH4)xV2O5
1.3H2O (x = 0.15 and 0.30) [21,22]. The method is based on the direct
acidication of NH4VO3 solutions by acetic acid at different temperatures between 50 and 85 C. Through rational selection of the deposition
parameters like vanadium concentration, temperature and deposition
time, (NH4)xV2O51.3H2O thin lms exhibiting high values of the transmittance variance (T) of 55% at 400 and 900 nm were designed.

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

However, the use of the same simple synthetic procedure in the case
of initial NaVO3, i.e. acidication of NaVO3 solution with acetic acid at
75 C (pH = 3), leads to the formation of unstructured amorphous
thin lms as we have previously established [23]. In that case a further
thermal treatment at 400 C was needed in order to obtain crystalline
lms which represent a two-phase mixture of sodium vanadium oxides
such as NaV6O15 and Na1.1V3O7.9. Thus prepared thin lms exhibited insufciently high T values of about 20% in the voltage range of 2 V.
The present work is focused on the examination of well-structured
hydrated sodium vanadium oxide thin lms with electrochromic properties. Such thin lms are prepared by an optimized chemical bath
method that ensures one-step deposition of the thin lms at low
temperature. For the purpose we have applied a different synthetic approach: instead of a direct acidication, we have used here an indirect
acidication through the hydrolysis of diethyl sulfate present in the
chemical bath. This controlled precipitation reaction gives rise to the deposition of thin lms of vanadium(V) oxide xerogel with composition of
Na0.33V2O5H2O having well-organized layered structure in the nanoscale region. These lm characteristics are advantageous to achieving
electrochemical stability and high T value of 55% which is reproducible
for 500 cycles. The changes during the lm aging are studied in respect
to the lm structure, V(V) reduction and electrochromic effect.
It is worth emphasizing that the as-prepared well-structured
Na0.33V2O5H2O composition (solid or lm) is highly benecial for
further obtaining a variety of chemical compositions. Thus, once obtained it can be used to produce either lower hydrated xerogels
Na0.33V2O5nH2O (0.3 b n b 1) or a single phase of NaV6O15 by thermal
treatment at an appropriate temperature. All these compositions having
a layered or tunnel structure can serve as host matrices for intercalation
processes which open opportunities for different applications. From this
point of view the difference with the previously studied (NH4)xV2O5
1.3H2O compositions is obvious: their thermal treatment produces the
well studied V2O5.

2. Material and methods


Thin lm deposition is performed onto commercially available glass
substrates. They are coated with a conductive, transparent thin layer of
SnO2:F (FTO) with 80% optical transparency in the visible spectrum and
electrical resistance of 1020 /cm2. Before deposition, the substrates
were cut into pieces with the dimensions 40 mm 25 mm 2 mm
and cleaned in the following order: with detergent, alkaline solution,
1:1 diluted hydrochloric acid, hexane, acetone and rinsed with deionized water and dried at room temperature. Commercial diethyl sulfate
(Sigma-Aldrich), propylene carbonate (Sigma-Aldrich), lithium perchlorate (Sigma-Aldrich), sodium metavanadate min 98 wt.% (Carlo
Erba), ethanol 96% (Alkaloid), sodium hydroxide (Merck), hydrochloric
acid (Merck), hexane (Merck) and acetone (Merck) are used without
further purication.
2.1. Preparation of the thin lms
The lms are deposited from a chemical bath with optimized composition and process conditions. The stock solution for preparation of
the chemical bath is obtained by heating at 65 C of 0.50 g sodium
metavanadate and 250 ml deionized water. The chemical bath solution
is prepared in a 120 ml beaker by mixing 100 ml sodium metavanadate
solution (0.016 M) and 0.5 ml of diethyl sulfate. The addition of diethyl
sulfate results in a change of the solution color from yellow to dark
orange. The cleaned substrates are vertically supported to the wall of
the beaker with the non-conductive side facing the wall. The deposition
system is then heated up to 85 C (deposition temperature) with continuous stirring and the achieved temperature and stirring are maintained during the deposition time. pH of the chemical bath during the
deposition is about 5.5. The beginning of the deposition reaction is

309

observed as the appearance of turbidity of the liquid phase which is


due to the formation of a solid substance (precipitate). Just at this moment the substrate is removed from the chemical bath and quickly,
but carefully, is wiped with cotton soaked with ethanol (96%). This procedure is needed in order to remove weakly sticking grains while the
well adherent grains remain on the surface. The substrate is then
returned into the chemical bath at the same position and the deposition
time starts to be measured. In such a manner we have prepared thin
lms for 5, 10 and 15 min deposition times. To prepare a thicker lm,
the lm already obtained for the 15 min deposition time is reinserted
into a fresh chemical bath for a further 15 min (2 15 min) which is
done 5 min after the beginning of the deposition reaction in the second
chemical bath. Thus the obtained lm will be further designated as that
prepared for the 30 min deposition time. By the deposition procedure
used the lms are deposited on the both sides of the substrates. To remove the thin lm from the non-conductive side of the substrate, this
site is carefully wiped with cotton wetted with 2 M aqueous solution
of sodium hydroxide. Finally, the as-deposited thin lms are wiped
with cotton soaked with ethanol, rinsed with ethanol and left vertically
to dry at room temperature. The thin lms prepared for the 5 to 15 min
deposition time have yellow color, while the thicker lm obtained for
the 30 min deposition time has yellow-brown color.
The precipitate from the chemical bath is separated by vacuum
ltration, washed with ethanol and dried in air at room temperature
for 34 h. The fresh precipitate has brown color.

3. Characterization of the thin lms


The composition and structure of both thin lms and precipitate
from the chemical bath were examined using a Rigaku Ultima IV X-ray
diffractometer with CuK radiation. The thermal studies (TG and
DTA) were carried by a LABSYS Evo apparatus (SETARAM) in a
temperature interval of up to 500 C in an airow at a heating rate of
10 C/min. Infrared spectra were recorded with a Perkin-Elmer System
2000 infrared interferometer using KBr disks.
The morphology of the thin lms was observed by scanning electron
microscopy (JEOL JSM-5510). The topography of the lm surfaces was
examined by AFM in taping mode at room temperature using
NanoScopeV system (Veeco Instruments Inc.).
In-situ optical spectra of the thin lms were recorded by a Varian
Cary 50 Scan spectrophotometer ranging from 350 to 900 nm at voltages in the range of 2.5 V. The electrochromic cell was a home-made
cell of 3 mm thick window glass. The cell is actually a Vis cuvette on a
squared glass base with holes that allow the cuvette to be attached to
the spectrophotometer. The electrochromic cell was used as a twoelectrode system: one electrode was a blank FTO substrate and the
other electrode was FTO substrate with a thin lm. The distance
between the electrodes was about 1 cm and 1 M LiClO4 in propylene
carbonate (PC) was used as an electrolyte (30 ml). The surface of each
electrode was about 8 cm2. The prolonged cycling up to 500 cycles
was performed in the same two-electrode electrochromic cell with alternative square pulse voltage of +2.5 V and switching time of 60 s.
The electrochemical behavior of Na0.33V2O5H2O thin lms were examined by cyclic voltammetry in 1 M LiClO4 (PC) in a conventional
three-electrode cell using a micro AUTOLAB II equipment (Eco-Chemie)
in the potential range initially between 2.5 and +2.5 V, and then reduced to 1 and +1 V. The prepared thin lm is the working electrode,
the reference electrode is Ag/AgCl (3 M KCl) and the auxiliary electrode
is a platinum wire. The CV curves are recorded at 10 and 50 mV/s
scanning rates.
The lm thickness was measured by a Alpha Step D-100
prolometer (measuring parameters: stylus force 5 mg, length 8 mm,
range 10 m and speed 0.07 mm/s). All as-deposited Na0.33V2O5nH2O
xerogels thin lms with different thickness have passed an adhesion
tape test.

310

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

4. Results and discussion


4.1. Chemical consideration for lm formation
The chemistry of the deposition process is based on a controlled precipitation reaction resulting from the acidication of the NaVO3 solution
that occurs in the presence of diethyl sulfate. The main idea for the precipitation process was taken from a previous study of one of the authors
[24]. Above 65 C the hydrolysis of diethyl sulfate takes place according
to the reaction:
CH3 CH2 O2 SO2 aq 4H2 Ol 2CH3 CH2 OHaq 2H3 O aq SO4 2 aq:

The concentration of H3O+ gradually increases (decreasing pH to


about 5.5) and the conditions for precipitation of vanadium(V) oxide
xerogels are fullled [6].
4.2. Composition and structure of the thin lms
Fig. 1 shows the X-ray powder diffraction (XRD) patterns of the asdeposited lm (Fig. 1a) and the precipitate from the chemical bath, so
called the brown sample (Fig. 1b).
Besides the peaks due to the FTO substrate (PDF 461088) the two
patterns are very similar, except the difference in the intensities of
some peaks, for instance at 25.49, 29.85, 47.31 and 50.54 (2). In
general, such a difference is reasonable and it is mainly due to the occurrence of texture in thin lm growth [25]. The similarity between the
XRD patterns evidences the same phase composition of the lm and
precipitate. This is of importance since we are able to undertake some
studies that require more amount of the sample (for example TG-DTA
analyses) using the precipitate and the obtained data are then referred
to the lm.
Both patterns display comparatively small number of diffraction
peaks, about ten peaks, all broad. The rst peak centered at 7.89
(d = 11.20 ) has much higher intensity than the other peaks located
above 23 (2 scale). The comparison with the patterns from PDF database showed that our diffractograms do not correspond exactly to any
known X-ray powder patterns of vanadium compounds. However,

Fig. 1. XRD patterns of (a) as-deposited lm, (b) brown precipitate and (c) greenish
precipitate.

there is a great similarity with the patterns of V2O5nH2O xerogel


(PDF 401296) with the difference that our patterns exhibit additional
peaks. From the literature survey on vanadium oxides [6,2629] it is
well known that the X-ray patterns of V2O5nH2O xerogel are characterized by a small number of 35 broad peaks from the series of 00l reections (missing 002 peak). Moreover, the intensity of the rst order
diffraction 001 located generally around d = 11.5 is much larger
than that of the other terms. The structure of xerogel was resolved by
Petkov et al. [29]. The xerogel is found to be an assembly of double
V2O5 sheet forming slabs that are stacked along the c-axis of a monoclinic unit cell. The slabs are separated by water molecules. The basal
distance between the layers depends on the amount of water and
increases by step of about 2.8 for each water layer: 11.55 for
n 1.51.6 and 8.75 for n 0.5 [30]. Due to the layered structure,
vanadium(V) oxide gels are able to intercalate a wide variety of
inorganic and organic guest species without change in the onedimensional stacking of the layers [6,3135]. It was reported that the intercalation of cations like Na+ and TMA+ (tetramethyl ammonium)
into the xerogel leads to the appearance of extra diffraction peaks in
the diffraction patterns of M0.3V2O51.5H2O [31]. Durupthy et al. [31]
have suggested that the observation of hkl set of reections (instead of
00l only) is related to the loss of ordered stacking of the double V2O5
layers. It is important that our diffractograms resemble to a great extent
the patterns given in the above paper. Based on all above we consider
that our synthesis product in the form of precipitate or lm is Na+
intercalated vanadium(V) oxide xerogel, NaxV2O5nH2O, so that the
oxidation number of vanadium in our samples becomes less than 5.
The incorporation of cations between the layers of the gels is a result
of ion-exchange reactions with the acid protons of the gels, so the
amount of the intercalated ions is around 0.30.4 per mole of V2O5
[31,3336]. For the sodium ions the equilibrium amount is found to be
0.33 [32]. Concerning our compositions (precipitate and lm) we have
one more argument in favor of the above ratio: during the annealing
at 400 C they completely transform into a single phase of monoclinic
NaV6O15 (Na0.33V2O5) (PDF 86120) without any additional diffraction
peaks (XRD patters are not shown) which conrms a Na:V ratio of 1:6 in
the as-prepared compositions.
The formation of vanadium(V) oxide xerogel is further supported by
IR spectroscopy (Fig. 2). It is clearly seen that the spectral characteristics
of the synthesis product are different from those of the initial reagent
NaVO3 (Fig. 2a). The IR spectra of the as-deposited scraped lm
(Fig. 2b) and precipitate (Fig. 2d) are practically identical (within the
limit of the experimental resolution), thus conrming the same phase
composition of the two samples.
The IR spectra (Fig. 2b, d) are dominated by strong absorptions in the
1020400 cm1 region associated with the vibrations of the vanadium
oxygen framework. The band at 1012 cm1 is attributed to the
stretching vibration of terminal V_O groups (shorter V\\O bond), the
band at 763 cm1 is due to asymmetric stretching vibrations of the
bridged V\\O\\V units (longer V\\O bonds), and the band at
514 cm1 is assigned to the V\\O\\V symmetric stretch mixed with
bending vanadium oxygen vibrations [3638]. The absorption at
917 cm1 is likely to be associated with a V_O stretch strongly
perturbed by the water molecules [38]. In addition, a weak band near
1230 cm1 appears but its origin is unclear (Fig. 2). The presence of
water molecules in the V2O5nH2O xerogel is clearly manifested by
the bands at 3590 cm1 (shoulder) and 3400 cm1 (OH stretching
vibrations) and at 1614 cm1 (HOH bending vibration) (Fig. 2b, d).
The band near 3600 cm1 has been assigned to water molecules nearly
free of hydrogen bonding, which are presumably directly bonded to
vanadium through their oxygen atom [3739]. The band below
3580 cm1, in our samples at 3400 cm1 has been attributed to the
water molecules hydrogen bonded with the oxygen either of V2O5 [37,
40] or of other H2O molecules [3739]. There is a third kind of water
molecules, that also give rise to a OH band around 3600 cm1 since
they are not involved in hydrogen bonds [37,38]. These water molecules

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

Fig. 2. IR spectra of (a) NaVO3, (b) as-deposited scraped lm obtained for 30 min, (c) lm
aged 20 months, (d) fresh brown precipitate and (e) brown-greenish precipitate aged
more than 3 years.

(n around 0.10.5) are supposed to be trapped into the cavities of the


lattice, where they are strongly retained and thus leave the structure
above 250 C just before the crystallization of the sample [30,37,38].
It is noticeable that the band positions in the prepared Na0.33V2O5
nH2O are very close to those for pristine V2O5nH2O xerogels: 1015,
760 and 515 cm1 [3739]. This nding evidences that the V2O5 framework in Na0.33V2O5nH2O is not essentially affected by the intercalated
sodium ions. All the same, the small downshift with 3 cm1 of the
band at 1012 cm1 in the as-prepared xerogel (Fig. 2) implies more
V(IV) sites than that in the pristine V2O5nH2O xerogels which can be
related to the presence of Na+ ions in the interlayer space.
The water content in the as-prepared fresh Na0.33V2O5nH2O precipitate (brown sample) was determined by the TG-DTA technique (Fig. 3).
As seen the release of the water molecules is stepwise in accord with
earlier reports [2,30,37]. The rst step is developed in the temperature
range of 40150 C (endothermic peak at 116 C) with a mass loss of
4.40% equivalent to 0.5 mol H2O. These water molecules are the most
weakly bonded ones. Between 150 and 270 C the TG curve shows a
gradual decrease with a small mass loss of 1.43% (0.16 mol H2O) without a distinct endothermic effect. A sharp endothermic effect appears
at 293 C which is accompanied by a sharp TG step with mass loss of
2.90% (0.33 mol H2O). This is the last amount of strongly retained
water molecules. Their release is immediately followed by an exothermic effect at 373 C due to the crystallization of monoclinic NaV6O15.
According to the TG curve the total mass loss is 8.73% which corresponds to one mole H2O per Na0.33V2O5.
From the XRD, IR spectroscopy and TG-DTA data we can conclude
that the synthesis product in the form of lm and precipitate is a
vanadium oxide xerogel with the composition Na0.33V2O5H2O.
It is worth mentioning that over time (for instance, in about 7 days)
both the precipitate and lms turn greenish. The color change occurs regardless the storage conditions (at ambient conditions or closed vessel)
and it is an indication for a reduction process at the V(V) sites. This

311

Fig. 3. TG-DTA curves for Na0.33V2O5H2O xerogel precipitates: (a) fresh and (b) aged
20 days.

process, however, does not appear to modify the XRD pattern of greenish precipitate (Fig. 1c) which is very similar to that of the fresh one and
only differences in the relative intensity of some peaks such as at 25.44
and 32.74 (2) are observed (Fig. 1b). Since the IR spectroscopy can
give valuable information on the change in the valence state of vanadium [41,42] we have further examined the IR spectra of the precipitate
and lm, both aged at different times, 20 days, 20 months and 3 years,
and some of them are included in Fig. 2. The inspection of the IR spectra
showed that the spectra of the precipitates aged 3 years and lm aged
20 months match closely those of the corresponding samples aged
20 days, as well as the spectra of the lms that are practically the
same with the spectra of the precipitates. So that, as an illustration in
Fig. 2 the IR spectra of a lm aged 20 months (Fig. 2c) and a precipitate
aged 3 years (Fig. 2e) are compared with the spectra of the corresponding fresh samples (Fig. 2b,d).
By IR spectroscopy it was established that the electrochemical reduction of V(V) from V2O5 to V(IV) causes a considerable modication
of the spectra as both the terminal V_O and bridging V\\O\\V
stretching vibrations are mainly affected [41,42]. Therefore, in the presence of more V(IV) species we should expect a red shift of the band at
1012 cm1 as well as a decrease in the intensity and frequency of the
band at 761 cm1 [41,42]. The close examination of the IR spectra in
Fig. 2 reveals that even prolonged aging does not result in any signicant difference in the band positions and intensities in comparison
with the fresh samples. Only a very small frequency variation between
2 and 4 cm1 is observed in some spectra, but these variations are not
consistent and, particularly the former one is within the limit of the experimental resolution (Fig. 2). These spectral features provide evidence
for the low degree of reduction of V(V) to V(IV) that takes place during
the aging, so the amount of the newly obtained V(IV) is not high enough
to affect the internal V\\O framework of the initial xerogel. We could
also suppose that the reduction process occurs mainly on the surface
of the particles, and thus the vibrational characteristics of the bulk
material appear to be unchanged.

312

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

The sensitivity of the vanadium(V) oxide xerogels toward reduction


under storage which is accompanied by progressive color change in
green after some weeks or months was commented on by Livage [6],
but the explanation of this phenomenon is still not clear. Unfortunately,
we are not able to specify what species are responsible for the electron
transfer process leading to the V(V) reduction. However, considering
the easy hydrolysis of the weakly bound water as found in [43] we
could suppose a possible role of the water molecules and protons in
the reduction process.
On the other hand, our IR absorption studies give valuable information that the degree of the reduction of V(V) in Na0.33V2O5H2O xerogel
does not increase during the prolonged aging at ambient conditions.
Whereas the vibrations of the V\\O units are not affected, some
changes are observed in the OH stretching mode region (Fig. 2). In the
greenish samples the intensity of the band around 3600 cm1 is increased and the band near 3440 cm1 is upshifted (Fig. 2c, e) compared
to the respective bands in the yellow/brown samples (Fig. 2b, d).
Obviously, some changes related to the amount and state of the water
molecules in the gel occur during storage and to elucidate this phenomenon we examined the thermal behavior of the greenish sample aged
20 days (Fig. 3). As seen the TG curves for the fresh and aged samples
differ considerably from each other. Firstly, the total mass loss for the
greenish sample is increased to 10.67% (vs. 8.73% in brown sample)
which leads to a higher water content of 1.3 mol. This water amount
is distributed in the following proportions: 1.1 mol (mass loss of
9.13%) are released between 40 and 205 C (rst strong endothermic effect at 175 C), about 0.1 mol (mass loss of 1.32%) between 205 and
270 C and the last amount of 0.1 mol between 270 and 320 C. The
comparison in the proportions between the two samples shows that
the greenish sample contains more amount of water molecules that
are removed below 200 C than the brown sample, i.e. more amount
of weakly bonded water molecules. Therefore, the extra water molecules accommodated during the storage are weakly bonded. This nding can explain the observed increase in the intensity of the (OH)
band at 3595 cm1 in the IR spectrum of the aged samples.

From the above data it follows that the vanadium reduction during
aging of our Na0.33V2O5H2O is accompanied by an increase in the
water content to Na0.33V2O51.3H2O. Such phenomenon has been previously reported for reduced xerogels [44]. Babonneau et al. [44] have
found that the reduced gels having V(IV)/V(V) of 16% contain 2.5 H2O
per V2O5 instead of 1.61.8 H2O for gels with usual V(IV)/V(V) about 1
to 4%.
4.3. Morphology of Na0.33V2O5H2O thin lms
As described in the literature [6,31] vanadium(V) oxide xerogels
normally exist in the form of long ribbons. SEM images of two thin
lms prepared for the 15 and 30 min deposition times are depicted in
Fig. 4.
Among the studied lms these are the lms exhibiting the best optical properties (see Section 4.5) and their thickness determined by the
prolometer is about ~150 and ~300 nm (for the 15 and 30 min deposition times, respectively). The cross-sectional SEM image of the lm for
the 30 min deposition time (Fig. 4d) gives a thickness of about 250 nm.
The chemical bath deposition method produces thin lms which
surface is completely covered with the deposited material (Fig. 4a, b
and c). Both lms are dense and without any porosity. The SEM image
of the thinner lm recorded at a higher magnication (Fig. 4b) clearly
shows the granular structure of the lm. It is composed of nanograins,
spherical and elongated, with sizes between 50 and 200 nm. Besides
well separated nanograins, randomly oriented ribbon-like units (about
200 nm wide and 11.5 m long) are also visible (Fig. 4a).
The observations by SEM are supported by AFM data for the same
Na0.33V2O5H2O thin lms (Fig. 5). The 2D surface topography of the
thinner lm (Fig. 5a) shows nanograins with sizes in the range of
50150 nm. In addition, there are elongated ribbon-like units with the
following dimensions: width from 150 to 300 nm and length from 0.7
to 2 m. The longer deposition time ensures the growth of the grains
and, expectedly, the thicker lm (Fig. 5b) exhibits larger grains with
sizes from 250 to 700 nm as well as larger ribbons reaching to 1 m

Fig. 4. SEM images of Na0.33V2O5H2O thin lms with different thickness: (a) and (b) ~150 nm, (c) ~300 nm and (d) cross-section view of the thin lm with ~300 nm thickness.

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

313

Fig. 5. AFM images of Na0.33V2O5H2O thin lms: (a) and (b) are 2D images with proles of lms with 150 and 300 nm thickness, respectively; (c) and (d) are 3D images of lms with 150
and 300 nm thickness, respectively.

wide and 3 m long. Moreover, AFM data reveal that the lms exhibit
comparatively high surface roughness with amplitude of the grain
height at around 600 nm for the thinner lm and 250500 nm for the
thicker one, i.e. the thickness is not uniform throughout the whole
lm area. The 3D images, and especially those for the thinner lm
(Fig. 5c), clearly illustrate the way for the formation of the ribbons.
We can see that the ribbons arise as a result of the coalescence of
nanograins in a preferred direction. We suppose that only the grains
having the most suitable orientation and that are situated closely to
each other are able to coalesce and thus to form long ribbon-like units.
Moreover, we can also speculate, that a possible growth mechanism of
Na0.33V2O5H2O thin lms is the island mechanism (Fig. 5d) in which
the coalescence of the islands generates ribbon-like units and the latter,
at a certain point, will form the compact lm layer.

4.4. Electrochemical properties


Firstly, we performed CV measurements in large potential range of
2.5 V with a blank substrate. As expected, we observed only oxygen
reduction that begins around 1.5 V and high polarization of the working electrode in the range of 1 V with ~0 mA/cm2 current density. This

allows us to perform electrochemical analysis in the relevant voltage


range of 1 V.
Fig. 6 shows cyclic voltammograms with ve scans (10 mV/s) of two
as-deposited lms with different thickness. Both cyclic voltammograms
exhibit three pairs of oxidation/reduction peaks. It is essential that the
shape and position of the respective redox peaks are very close for the
two lms (they differ within 0.04 V) which reect the same electrochemical processes. The anodic peaks (A1, A2, A3) appear at 0.15,
0.16 and 0.74 V and the cathodic peaks (C1, C2, C3) are at 0.50,
0.16, and 0.38 V (Fig. 6a). The A2/C2 pair is the most intense one.
The color changes observed in the three-electrode cell are the same as
those in the two-electrode electrochromic cell.
The survey from the literature shows that the cyclic voltammograms
of vanadium(V) oxide xerogels display one [45], two [46] and three
redox pairs [2]. The presence of one redox pair has been attributed to
the amorphous nature of the lms [45]. According to Benmoussa et al.
[46] the appearance of two redox pairs reects the crystalline nature
of the lms. The observation of three redox pairs in our lms is consistent with data of Costa et al. obtained at the same scanning speed for vanadium oxide gel prepared on exible polyethylene terephthalate/
indium tin oxide electrodes [2]. It should be mentioned that the general
view of the CV curves presented by Costa et al. [2] also having an

314

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

Fig. 6. Five cyclic voltammograms of Na0.33V2O5H2O thin lms with thickness: (a) 150 nm
and (b) 300 nm at a scanning rate of 10 mV/s.

intensive A2/C2 redox pair as well as the voltage ranges for the anodic
(from 0.23 to 0.64 V) and cathodic (from 0.8 to 0.3 V) peaks are
very similar to our data.
The calculation of the exchanged charge showed that the values
for the extracted charge for the thin lm with 150 nm thickness is
65.5 mC/cm2 and for the lm with 300 nm thickness is 75.5 mC/cm2.
The inserted charge values are much lower, 19.9 mC/cm2 and
22.3 mC/cm2 respectively. The higher values for the exchanged charge
for thicker lms are expected due to their higher capacity.
The electrochemical reversibility for 200 cycles was monitored at
scanning rate of 50 mV/s for the lm with 300 nm thickness (Fig. 7).
In conformity with the higher scanning rate all CV curves in Fig. 7
display one broad pair of oxidation/reduction peak instead of the
three pairs well-separated at the lower scanning rate (Fig. 6b). The
oxidation/reduction pair remains stable during the cycling with a slight
shift of the peak maximum to lower potentials: from 0.48 V/0.44 V for
the 5th scan to 0.43 V/ 0.41 V for the 200th scan. Moreover, the peak
maxima move to higher potentials than those of the most intensive
A2/C2 pair at the lower rate. It is also observed that the peak area
(exchanged charges) decreases during the cycling and this process is

Fig. 7. Cyclic voltammograms to 200 scans of thin lm with 300 nm thickness at a scanning
rate of 50 mV/s.

more pronounced up to the 100th cycle. Then, the electrochemical


reversibility appears to be stabilized.
As mentioned previously, over time, the as-deposited lms changed
its original yellow/yellow-brown color into greenish color. Fig. 8 compares the CV curves (up to 20 scans) of a fresh lm (yellow-brown)
and an aged 20 days lm (brown-greenish) with the same thickness.
Opposite to our expectations, the comparison shows that electrochemical behavior of the fresh and aged lm during the initial scans
does not differ considerably. In both cases there are three cathodic
and three anodic peaks, however, the rst anodic peak A1 is very well
distinguished in the fresh lm, while in the aged one it appears as a
shoulder to the second A2 peak. The corresponding peak potentials
show differences within 0.06 V.
With increasing scan number a clear difference between the fresh
and aged lm is observed (Fig. 8). During the longer cycling some
changes occur and the fresh lm is more essentially concerned. It is
clearly seen (Fig. 8a) that after 15 scans some peaks disappear in the
CV curves of the fresh lm. The A3/C3 pair surely disappears, the
A1/C1 pair strongly diminishes and the A2 peak is shifted to more positive values of the potential. Thus, the cyclic voltammograms of the fresh
lm transform from a system with three redox peaks to a system with
only one redox pair, A2/C2, which is characteristic of amorphous lms
[47]. At the same time the aged lm exhibits more stable cycling behavior and only the redox A1/C1 pair decreases in intensity. Regarding the
exchanged charges, going from the 1st to the 20th scans for the fresh
lm the extracted charge slightly decreases from 41 to 37 mC/cm2,
while the decrease in the inserted charge is more pronounced (from
15 to 8 mC/cm2). The lm aging causes a reduction mostly of the extracted charges (from 34 to 30 mC/cm2 for 5th and 20th scans, respectively), while the inserted charges are similar to those for the fresh
lm (from 14 to 8 mC/cm2 for 5th and 20th scans), respectively.
The peak disappearing in the CV curves with the increasing scan
number is observed by other scientists in the cases of vanadium oxide
xerogels [2,48] but the origin of this phenomenon is still not clear.
Several possible reasons are commented in the literature. The deactivation of some redox sites of the lm could be related either to
amorphization of the lm [45] or to crystalline phase transition

Fig. 8. Cyclic voltammograms of thin lms with 300 nm thickness: (a) fresh lm, 1st to 4th
and 16th and 20th scans, (b) aged lm, 5th, 7th, 10th, 15th and 20th scans (10 mV/s).

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

315

triggered by the oxidation/reduction cycles [2] or to changes in morphology and microstructure of the lm leading to the formation of a
more homogeneous lm [48].
The observed redox peaks in the potential range 1 V for
Na0.33V2O5H2O thin lms are related to step-wise reduction of V(V) to
V(IV) concomitant with formation of different crystalline states LixV2O5
and, accordingly reversible V(IV) oxidation and Li-deintercalation
[46,49,50].
4.5. Optical properties
Optical transmittance spectra in the 350900 nm spectral range of
the as-deposited thin lms are given in Fig. 9. It is important that the
four lms exhibit very similar dependence of the optical transmittance
on the wavelength, but the overall transmittance decreases with the
lm thickness. This is expected since more states in the thicker lm
are available for the photons to be absorbed. The thin lms have relatively high transmittance (low absorbance) at wavelengths longer
than 550 nm as the thinner lms (obtained for 5, 10 and 15 min deposition time) exhibit transmittance between 70 and 90%, while the thick
lm (30 min deposition time) exhibits a lower transmittance between
55 and 70%. The marked reduction in the transmittance at wavelengths
shorter than 550 nm is associated with the fundamental absorption. It is
essential that the absorption edge for the four lms appears in a narrow
interval of wavelengths (365385 nm). This means that the lms with
different thicknesses should have very close values of the optical band
gap which implies the same or very close stoichiometry regarding the
different vanadium sites. In this regard we have also compared the IR
spectra of the scraped lms obtained for the 5, 15 and 30 min deposition
and we have not found any variation in the positions and intensity of the
vibrational bands in comparison with the IR spectrum presented in
Fig. 2b. Considering these data we believe that the change of the lm
color with the time of deposition from yellow to yellow-brown shade
is mainly related to the lm thickness and grain growth evidenced by
AFM analysis (Fig. 5a, b) rather than the same changes in the vanadium
oxidation state during the deposition.
The optical transmittance of Na0.33V2O5H2O thin lms in the
350900 nm range is recorded in a two-electrode electrochromic cell
using LiClO4/PC electrolyte. Our previous studies on (NH4)xV2O5
1.3H2O thin lms in the voltage range between 1 and 2.5 V have
shown that the increased voltage results in increasing transmittance
variance (T) and shorter bleaching and coloration response times [30].
It is important to notice that the current densities achieved in the
used two-electrode electrochromic cell are around 0.27 mA/cm2 at
+2.5 V (at the highest voltage), but these values are still much lower
than those achieved in the three-electrode cell (CV measurements):
around 1.5 mA/cm2 at a signicantly lower voltage of 0.3 V. This fact explains the longer response time (Fig. 10) observed in the two-electrode
electrochromic cell.

Fig. 9. Optical transmittance spectra of as-deposited thin lms with time of deposition.

Fig. 10. Response time of Na0.33V2O5H2O thin lm with 300 nm thickness at 900 nm at
2.5 V.

The response time () is calculated as the time required for the thin
lm to change its color from yellow/brown to obtain 90% [51] of the blue
color (c,90% coloration response time) or vice versa, to obtain 90% of
the yellow/brown color (b,90% bleaching response time). It is seen
in Fig. 10 that the bleaching response time b,90% is 10 min, while the coloration response time c,90% is 17.5 min i.e. the reduction process followed with Li+-intercalation is almost two times slower than the oxidation
process followed with Li+-deintercalation.
The optical transmittance spectra recorded at 2.5 V with a
switching time of 20 min of four Na0.33V2O5H2O thin lms prepared
for different deposition times are shown in Fig. 11. It is seen that the dependence of the transmittance of the reduced forms on wavelength
passes through a maximum at 530 nm (Fig. 11a) and 560 nm
(Fig. 11b, c, d). The oxidized forms exhibit a steep increase in the transmittance up to 600 nm with further gradual increases up to 900 nm. It is
important that the fundamental absorption edge for both reduced and
oxidized lms exhibits a red shift (to lower energy) with the increase
in the lm thickness. The red shift of the absorption edge reects a decrease in the optical band gap and this effect has been attributed to
the increase in the grain size and effective decrease in the imperfections
at the grain-boundary regions [52].
Transmittance variance (T) dened as: T = T(bleached) T(colored)
is used to quantify the electrochromic effect of the prepared thin lms.
For all thin lms the maximum T value is obtained at 900 nm
(Fig. 11). The three thinner lms exhibit close T values of 3237%,
while for the thicker lm having 300 nm thickness T reaches 55%.
For the latter lm the optical response during a prolonged cycling up
to 500 cycles was also examined and the corresponding optical transmittance spectra are included in Fig. 11d. As seen the oxidized state is
completely recovered during the cycling and the transmittance curve
after 500 cycles well matches the initial curve. Some changes occur in
the reduced states between the 5th and 100th cycles resulting in an increase in the transmittance between 400 and 500 nm and shift of the absorption edge to lower wavelengths (high energy) as well as a slight
decrease in the transmittance between 500 and 700 nm. The observed
blue shift of the absorption edge reects a blue shift of the optical
band gap and this shift could be related to the change of the lm microstructure and increase in the imperfections during the lithium intercalation process [53]. Further changes between 100th and 500th cycles are
not observed, so that the reduced state is stabilized which is evident
from the overlap of the corresponding transmittance curves.
Therefore, the data for the optical stability of the thin lm after the
100th cycle in the electrochromic cell are in line with the data for its
electrochemical stability. Between 700 and 900 nm the transmittance
of the reduced forms in all cycles shows the same trend as the initial
scan. It is very important that the T value after 500 cycles is 53% at
900 nm i.e. it retains practically unchanged which is a demonstration
of the good optical stability of as-deposited lms. The comparison
with the literature data for electrochromic vanadium(V) oxide xerogels

316

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

Fig. 11. In-situ optical transmittance spectra of Na0.33V2O5H2O thin lms prepared for different deposition times: (a) 5 min, (b) 10 min, (c) 15 min and (d) 30 min; after 5 cycles in
(a, b, c) and up to 500 cycles in (d).

[47,51,54,55] shows that the obtained results are promising and can
serve as a base for further design of electrochromic devices.
As was mentioned before, the as-deposited yellow/yellow-brown
lms turn greenish tint over time. Fig. 12 shows the change in the optical transmittance over time (up to 20 months) of the best lm with
300 nm thickness. The optical transmittance spectra evidence that
aging does not affect the transmittance in the 350550 nm spectral region. As expected from the greenish shade of the aged lm, however, a
gradual decrease of the lm transmittance at wavelengths longer than
550 nm (i.e. an increased absorption of the red light) is observed as
the transmittance reduction in 20 days aging is about 10% at 900 nm.
It is also seen that the transmittance reduction is most signicant up
to 10th day while the lm structure still stabilizes (about 8%) and after
that it is only about 2%. Further decrease of the transmittance is not recorded even for a prolonged aging of 20 months (Fig. 12). It is very important, however, that when such a lm aged 20 days is placed in the
electrochromic cell (Fig. 12b) it is completely recovered in its oxidized
state and thus T retains its high value (56%).
5. Conclusions
Well-structured nano-sized thin lms with compositions Na0.33V2O5
nH2O (n = 1 and 1.3) are successfully deposited by an optimized chemical bath method where the acidication of NaVO3 solution occurs
through the hydrolysis of diethyl sulfate. The thin lms demonstrate appreciable optical transmittance varying between 40 and 70% at 500 nm in
dependence on the lm thickness. The lm structure and morphology,
electrochemical behavior and electrochromic activity of fresh and aged
thin lms are studied. It is established that the long-term aging of the
lms does not affect their electrochromic properties. The thin lms exhibit maximum transmittance variance (T) at 900 nm which varies between 32 and 55% depending on the lm thickness. The electrochemical

Fig. 12. Optical transmittance spectra of as-deposited thin lm with 300 nm thickness:
(a) over time and (b) brown-greenish lm aged 20 days placed in the electrochromic cell.

M. Najdoski et al. / Surface & Coatings Technology 277 (2015) 308317

reversibility and electrochromic reproducibility with T retention up to


500 cycles have been demonstrated. These characteristics can be attributed to the well-dened structure and nanosized morphology of the thin
lms provided by the suitable preparation method. The results obtained
are good prerequisite for further application of Na0.33V2O5nH2O thin
lms in different electrochromic devices.
Acknowledgment
The authors thank the Bulgarian Academy of Sciences and the
Macedonian Academy of Sciences and Arts for the nancial support
and the Alexander von Humboldt Foundation for providing the electrochemical equipment.
References
[1] W.D. Murphy, A.P. Christian, J.F. Disalvo, V.J. Waszczak, Lithium incorporation by
vanadium pentoxides, Inorg. Chem. 18 (1979) 28002803.
[2] C. Costa, C. Pinheiro, I. Henriques, C.A.T. Laia, Inkjet printing of solgel synthesized
hydrated tungsten oxide nanoparticles for exible electrochromic devices, Appl.
Mater. Interfaces 4 (2012) 52665275.
[3] A.L. Pergament, E.L. Kazakova, G.B. Stefanovich, Optical and electrical properties of
vanadium pentoxide xerogel lms: modication in electric eld and the role of
ion transport, J. Phys. D. Appl. Phys. 35 (2002) 21872197.
[4] F. Cheng, J. Chen, Transition metal vanadium oxides and vanadate materials for lithium batteries, J. Mater. Chem. 21 (2011) 98419848.
[5] D. Vernardou, State-of-the-art of chemically grown vanadium pentoxide nanostructures with enhanced electrochemical properties, Adv. Mater. Lett. 4 (2013)
798810.
[6] J. Livage, Vanadium pentoxide gels, Chem. Mater. 3 (1991) 578593.
[7] K. Takahashi, S.J. Limmer, Y. Wang, G.Z. Cao, Synthesis and electrochemical properties of single-crystal V2O5 nanorod arrays by template-based electrodeposition, J.
Phys. Chem. B 108 (2004) 97959800.
[8] J. Scarminio, A. Talledo, A.A. Andersson, S. Passerini, F. Deckers, Stress and
electrochromism induced by Li insertion in crystalline and amorphous V2O5 thin
lm electrodes, Electrochim. Acta 38 (1993) 16371642.
[9] V. Vivier, S. Belair, C.C. Vivier, J.Y. Nedelec, L.T. Yu, A rapid evaluation of vanadium
oxide and manganese oxide as battery materials with a micro-electrochemistry
technique, J. Power Sources 103 (2001) 6166.
[10] E.A. Olivetti, J.H. Kim, D.R. Sadoway, A. Asatekin, A.M. Mayes, Solgel synthesis of vanadium oxide within a block copolymer matrix, Chem. Mater. 18 (2006)
28282833.
[11] V.S.R. Channu, R. Holze, E.H. Walker Jr., S.A. Wicker Sr., R.R. Kalluru, Q.L. Williams, W.
Walters, Synthesis and characterization of lithium vanadates for electrochemical
applications, Int. J. Electrochem. Sci. 5 (2010) 13551366.
[12] H. Yin, K. Yu, H. Peng, Z. Zhang, R. Huang, J. Travas-Sejdic, Z. Zhua, Porous V2O5
micro/nano-tubes: synthesis via a CVD route, single-tube-based humidity sensor
and improved Li-ion storage properties, J. Mater. Chem. 22 (2012) 50135019.
[13] J. Legendre, J. Livage, Vanadium pentoxide gels: I. Structural study by electron
diffraction, J. Colloid Interface Sci. 94 (1983) 7583.
[14] R. Ceccato, G. Carturan, Solgel synthesis of vanadate-based thin lms as counter
electrodes in electrochromic devices, J. Sol-Gel Sci. Technol. 26 (2003) 10711074.
[15] C.G. Granqvist, Electrochromics for smart windows: oxide-based thin lms and
devices, Thin Solid Films 564 (2014) 138.
[16] S. Papaefthimiou, E. Syrrakou, P. Yianoulis, An alternative approach for the energy
and environmental rating of advanced glazing: an electrochromic window case
study, Energy Build. 41 (2009) 1726.
[17] R. Baetens, B.P. Jelle, A. Gustavsen, Properties, requirements and possibilities of
smart windows for dynamic daylight and solar energy control in buildings: a state
of the art review, Sol. Energy Mater. Sol. Cells 94 (2010) 87105.
[18] N.I. Jaksic, C. Salahifar, A feasibility study of electrochromic windows in vehicles, Sol.
Energy Mater. Sol. Cells 79 (2003) 409423.
[19] H.K. Koduru, H.M. Obili, G. Cecilia, Spectroscopic and electrochromic properties of
activated reactive evaporated nano-crystalline V2O5 thin lms grown on exible
substrates, Int. Nano Lett. 3 (2013) 2431.
[20] S. Beke, A review of the growth of V2O5 lms from 1885 to 2010, Thin Solid Films
519 (2011) 17611771.
[21] M. Najdoski, V. Koleva, A. Samet, Effect of deposition conditions on the
electrochromic properties of nanostructured thin lms of ammonium intercalated
vanadium pentoxide xerogel, J. Phys. Chem. C 118 (2014) 96369646.
[22] M. Najdoski, V. Koleva, A. Samet, Inuence of vanadium concentration and temperature on the preparation of electrochromic thin lms of ammonium intercalated
vanadium(V) oxide xerogel nanoribbons, Dalton Trans. 43 (2014) 1253612545.
[23] M. Najdoski, V. Koleva, S. Demiri, Chemical bath deposition and characterization of
electrochromic thin lms of sodium vanadium bronzes, Mater. Res. Bull. 47 (3)
(2011) 737743.

317

[24] M. Najdoski, T. Todorovski, A simple method for chemical bath deposition of


electrochromic tungsten oxide lms, Mater. Chem. Phys. 104 (2007) 483487.
[25] M. Birkholz, Thin Film Analysis by X-ray Scattering, John Wiley & Sons, New Jersey,
2006 230.
[26] R. Ceccato, S. Dir, T. Barone, G. De Santo, E. Cazzanelli, Growth of nanotubes in
solgel-derived V2O5 powders and lms prepared under acidic conditions, J.
Mater. Res. 24 (2009) 475481.
[27] T. Yao, Y. Oka, N. Yamamoto, Layered structures of vanadium pentoxide gels, Mater.
Res. Bull. 27 (1992) 669675.
[28] J. Livage, F. Beteille, C. Roux, M. Chatry, P. Davidson, Solgel synthesis of oxide
materials, Acta Mater. 46 (1998) 743750.
[29] V. Petkov, P.N. Trikalitis, E.S. Bozin, S.J.L. Billinge, T. Vogt, M.G. Kanatzidis, Structure
of V2O5nH2O xerogel solved by the atomic pair distribution function technique, J.
Am. Chem. Soc. 124 (2002) 1015710162.
[30] P. Aldebert, N. Bafer, N. Gharbi, J. Livage, Layered structure of vanadium pentoxide
gels, Mater. Res. Bull. 16 (1981) 669676.
[31] O. Durupthy, N. Steunou, T. Coradin, J. Maquet, C. Bonhomme, J. Livage, Inuence of
pH and ionic strength on vanadium(V) oxides formation. From V2O5nH2O gels to
crystalline NaV3O81.5H2O, J. Mater. Chem. 15 (2005) 10901098.
[32] L. Znaidi, N. Bafer, D. Lemordant, Kinetics of the H+/M+ ion exchange in V2O5
xerogel, Solid State Ionics 2830 (1988) 17501755.
[33] Y.J. Liu, J.A. Cowen, T.A. Kaplan, D.C. De Groot, J. Schindler, C.R. Kannewurf, M.G.
Kanatzidis, Investigation of the alkali-metal vanadium oxide xerogel bronzes:
AxV2O5nH2O (A = K and Cs), Chem. Mater. 7 (1995) 16161624.
[34] Y.J. Liu, J.L. Schindler, D.C. De Groot, C.R. Kannewurt, W. Hirpo, M.G. Kanadzidis,
Synthesis, structure, and reactions of poly(ethylene oxide)/V2O5 intercalative nanocomposites, Chem. Mater. 8 (1996) 525534.
[35] V.L. Volkov, G.S. Zakharova, Nanotubes in the V2O5nH2O xerogel-hydroquinone
H2O system, Russ. J. Inorg. Chem. 54 (2009) 3639.
[36] M. Malta, G. Louarn, N. Errien, R.M. Torresi, Redox behavior of nanohybrid material
with dened morphology: vanadium oxide nanotubes intercalated with polyaniline,
J. Power Sources 156 (2006) 533540.
[37] L. Abello, E. Husson, Y. Repelin, G. Lucazeau, Structural study of gels of V2O5: vibrational spectra of xerogels, J. Solid State Chem. 56 (1985) 379389.
[38] Y. Repelin, E. Husson, L. Abello, G. Lucazeau, Structural study of gels of V2O5: normal
coordinate analysis, Spectrochim. Acta A 41 (1985) 9931003.
[39] C. Sanchez, J. Livage, G. Lucazeau, Infrared and Raman study of amorphous V2O5,
J. Raman Spectrosc. 12 (1982) 6872.
[40] M.T. Vandenborre, R. Prost, E. Huard, L. Livage, Etude par spectroscopie infra-rouge
de l'eau adsorbee sur un xerogel d'oxyde de vanadium, Mater. Res. Bull. 18 (1983)
11331142.
[41] A. urca, B. Orel, G. Drai, B. Pihlar, Ex situ and in situ infrared spectroelectrochemical investigations of V2O5 crystalline lms, J. Electrochem. Soc. 146
(1999) 232242.
[42] M. Hajzeri, A. urca Vuk, L.S. Pere, M. olovi, B. Herbig, U. Posset, M. Krmanc, B.
Orel, Solgel vanadium oxide thin lms for a exible electronically conductive polymeric substrate, Sol. Energy Mater. Sol. Cells 99 (2012) 6272.
[43] D.S. Yakovleva, V.P. Malinenko, A.L. Pergament, G.B. Stefanovich, Electrical and optical properties of thin lms of hydrated vanadium pentoxide featuring
electrochromic effect, Tech. Phys. Lett. 33 (2007) 10221024.
[44] F. Babonneau, P. Barboux, F.A. Josein, J. Livage, Rduction des gels de V2O5, J. Chim.
Phys. 82 (1985) 761766.
[45] J. Livage, Optical and electrical properties of vanadium oxides synthesized from
alkoxides, Coord. Chem. Rev. 190192 (1999) 391403.
[46] M. Benmoussa, A. Outzourhit, R. Jourdani, A. Bennouna, E.L. Ameziane, Structural,
optical and electrochromic properties of solgel V2O5 thin lms, Act. Passive
Electron. Compon. 26 (2003) 245256.
[47] N. Ozer, S. Sabuncu, J. Cronin, Electrochromic properties of solgel deposited
Ti-doped vanadium oxide lm, Thin Solid Films 338 (1999) 201206.
[48] F.J. Anaissi, G.J.F. Demets, H.E. Toma, Electrochemical conditioning of
vanadium(V) pentoxide xerogel lms, Electrochem. Commun. 1 (1999) 332335.
[49] N.T.B. Bay, P.M. Tien, S. Badilescu, Y. Djaoued, G. Bader, F.E. Girouard, V.V. Truong,
L.Q. Nguyen, Optical and electrochemical properties of vanadium pentoxide gel
thin lms, J. Appl. Phys. 80 (1996) 70417045.
[50] S.F. Cogan, N.M. Nguyen, S.J. Perrotti, R.D. Rauh, Optical properties of electrochromic
vanadium pentoxide, J. Appl. Phys. 66 (1989) 13331337.
[51] W. Zhongchun, C. Jiefeng, H. Xingfang, Electrochromic properties of aqueous solgel
derived vanadium oxide lms with different thickness, Thin Solid Films 35 (2000)
238241.
[52] D. Sun, C.W. Kwon, G. Baure, E. Richman, J. MacLean, B. Dunn, S.H. Tolbert, The relationship between nanoscale structure and electrochemical properties of vanadium
oxide nanorolls, Adv. Funct. Mater. 14 (2004) 11971204.
[53] C.V. Ramana, R.J. Smith, O.M. Hussai, Grain size effects on the optical characteristics
of pulsed-laser deposited vanadium oxide thin lms, Phys. Status Solidi 1 (2003)
R4R6.
[54] N. Ozer, Electrochemical properties of solgel deposited vanadium pentoxide lms,
Thin Solid Films 305 (1997) 8087.
[55] M.R.J. Scherer, P.M.S. Cunha, O.A. Scherman, U. Steiner, Enhanced electrochromism
in gyroid-structured vanadium pentoxide, Adv. Mater. 24 (2012) 12171221.

Das könnte Ihnen auch gefallen