You are on page 1of 16

Article

Synergistic TiO2/VO2 Window Coating with


Thermochromism, Enhanced Luminous
Transmittance, and Photocatalytic Activity
Yuxia Ji, Andreas Mattsson,
Gunnar A. Niklasson, Claes G.
Granqvist, Lars Österlund
lars.osterlund@angstrom.uu.se

HIGHLIGHTS
Coatings for energy-efficient and
air-cleaning glazing

Synergistically enhanced
thermochromic, luminous, and
photocatalytic properties

Boosted photocatalytic
degradation due to absorption of
near-infrared solar light

Bilayer coatings of TiO2/VO2 lead to glazing that can purify indoor air and lower
solar energy inflow. They are important for energy-efficient buildings with reduced
needs for ventilation and air conditioning. The underlying thermochromic VO2
layer modulates solar energy throughput by absorbing infrared light film that heats
up the photocatalytic TiO2 film and yields boosted photocatalytic degradation of
pollutants. The top photocatalytic layer also serves as an antireflection coating and
provides enhanced solar light modulation and luminous transmittance.

Ji et al., Joule 3, 2457–2471


October 16, 2019 ª 2019 Elsevier Inc.
https://doi.org/10.1016/j.joule.2019.06.024
Article
Synergistic TiO2/VO2 Window Coating
with Thermochromism, Enhanced Luminous
Transmittance, and Photocatalytic Activity
Yuxia Ji,1,2 Andreas Mattsson,1,2 Gunnar A. Niklasson,1 Claes G. Granqvist,1 and Lars Österlund1,3,*

SUMMARY Context & Scale


Modern buildings tend to be ‘‘energy guzzlers’’ and have indoor environments The world’s appetite for energy is
with unhealthy air. Glazing with TiO2/VO2 bilayer coatings (1) exhibits enhanced rising, and much of this energy is
photocatalytic air purification compared with single-layer TiO2 due to heating used for the cooling, heating, and
from the underlying infrared-absorbing VO2 film, (2) is thermochromic thanks lighting of buildings. Energy-
to the VO2 and admits less solar energy inflow when there is a cooling demand, efficient buildings are therefore
and (3) has significantly improved luminous transmittance as a result of antire- urgently needed. There is also
flection due to TiO2. These coatings were deposited by reactive DC magnetron increasing awareness that poor air
sputtering onto heated glass; they were nanocrystalline, and the anatase phase quality affects human health and
prevailed in TiO2. The VO2 coatings showed well-developed thermochromism. leads to millions of premature
The photocatalytic degradation rate of stearic acid was almost doubled for deaths yearly. A ‘‘holy grail’’ for
the TiO2/VO2 bilayer film compared with that for single-layer TiO2. Our results building development is a means
demonstrate an important, and hitherto unexplored, synergy between photoca- to control energy flows through
talysis, thermochromism, and high luminous transmittance which exploits spec- glazing while keeping indoor
tral-selective material properties for solar illumination. environments healthy and
comfortable. Here, we
demonstrate that a bilayer coating
INTRODUCTION
on the glazing can purify indoor air
People in industrialized countries spend as much as 80%–90% of their time indoors, and lower solar energy ingress
in buildings, or in vehicles.1 This environment is fraught with problems from the per- while preserving the glazing’s
spectives of energy and health, and new techniques are needed to alleviate these main function, to enable good
problems.2 Here we demonstrate that novel glazing, i.e., windows and glass fa- indoor-outdoor contact and
cades, can decompose polluting organic molecules at the same time as solar energy daylighting. One part of the
inflow is modulated to achieve energy efficiency; furthermore, these benign features bilayer is thermochromic and
can operate synergistically and achieve enhanced catalytic and optical properties gives temperature-dependent
exploiting broadband spectral-selective materials properties—beyond what has throughput of solar energy, and
hitherto been reported for multilayer structures.3–5 the other part is photocatalytic
and able to destroy pollutants in
Today’s buildings are responsible for 30%–40% of the global use of primary energy,6 and indoor air. These functions
their role is growing in many countries. For example, the buildings’ share of the energy operate synergistically and do not
consumption in the USA was 34% in 1980 but as high as 41% in 2010.7 The ensuing effects hamper the glazing’s visible
of energy production on climate change and increasing sea level are well known.8 Clearly performance, thus enabling
it is urgent that the buildings’ energy footprint be reduced, which calls for diminished arti- indoor-air-cleaning windows.
ficial heating and cooling, as well as more efficient air treatment, all of which have to be
implemented without imperiling human well-being and productivity.9

One of the important technologies for lowering the energy footprint of buildings is
‘‘smart’’ glazing, characterized by its ability to regulate the throughput of luminous
radiation and solar energy via an external stimulus.2,10 An option of present interest
employs self-adaptive thermochromic coatings, which are able to produce energy

Joule 3, 2457–2471, October 16, 2019 ª 2019 Elsevier Inc. 2457


efficiency by having a lower (higher) solar transmittance (Tsol) at a high (low) temper-
ature, with the property change occurring close to the human comfort tempera-
ture.11,12 Such coatings are based on vanadium dioxide (VO2)13–15 and utilize its
well-known metal-insulator phase transition, associated with a reversible crystal
structure transformation between monoclinic and tetragonal, at a ‘‘critical’’ temper-
ature tc z 68 C.16 Below this transition temperature, at t < tc, VO2 is semiconduct-
ing and thin films are infrared transmitting, whereas such films are metallic-like and
infrared reflecting at t > tc. The transition temperature can be lowered, especially by
addition of some tungsten.17 The luminous transmittance (Tlum) can be significant
irrespective of temperature. Pure VO2 films are not good enough for glazing,17
but the films’ properties can be enhanced by antireflection treatment,18,19 incorpo-
ration of magnesium or strontium,20,21 fluorination,22 nanostructuring,23,24 and
protective top coatings.25,26

It is well known that poor indoor air quality has adverse effects on humans.27 These
effects are aggravated by contaminated outdoor air and by accumulation of pollut-
ants in the indoor environment, typically due to emissions from materials and
human activity inside the building.28 Characteristic examples of volatile organic in-
door pollutants are ketones (in particular aldehydes), BTXS (i.e., benzene, toluene,
xylenes, and styrene), and terpenes.29–31 These contaminants, which commonly are
in the ppb concentration range, are also associated with the so-called sick building
syndrome,32 which affects the users of up to 30% of newly constructed or refur-
bished buildings in Europe and North America.33 Active indoor air remediation
can be done by filtering with active carbon or, as is the common practice, by
increasing the ventilation of the building. However, the latter measure requires
that the outdoor air is clean enough or that it is filtered. Ventilation is also associ-
ated with enhanced demands for air conditioning, thereby increasing the buildings’
energy consumption, which highlights the need for innovation in passive air clean-
ing.34 It has been demonstrated that photocatalytic technologies that allow efficient
air purification permit down-sizing of heating, ventilation, and air conditioning
(HVAC) systems by 25% with maintained air quality, and these technologies are
able to reduce the energy demand by up to 30% without increased investment
costs.35

Ideally, air purification in a photocatalytic system should rely on nothing but avail-
able electromagnetic radiation, such as solar irradiation during the day and indoor
lighting. Photocatalysis is a technique capable of meeting these requirements and
is able to remove almost any type of organic air pollutant36 in a manner that is
both energy and cost effective.35 Titanium dioxide (TiO2) is the most commonly
used photocatalyst36; it exists as different polymorphs, the naturally occurring
ones being anatase, rutile, and brookite.37 Rutile is the thermodynamically most sta-
ble structure, whereas anatase stands out as the most photocatalytically active
phase,38–40 though still not sufficiently active for sustained solar-light-activated pho-
tocatalysis for efficient indoor air cleaning. Hence new materials or methods are
intensively being looked for to realize solar-light-activated photocatalysis for air
purification.
1Department of Engineering Sciences, the
Ångström Laboratory, Uppsala University,
Possibilities to use thermochromic VO2 thin films for energy-efficient glazing were P. O. Box 534, Uppsala SE-75121, Sweden
discussed already several decades ago,41,42 and photocatalytic TiO2 thin films are 2These authors contributed equally
in commercial use in self-cleaning (or ‘‘easy-clean’’) glazing.36,43 These materials 3Lead Contact
have also been combined, and TiO2 and other materials have been employed as *Correspondence:
antireflection coatings on top of VO2 films to improve luminous transmittance and lars.osterlund@angstrom.uu.se
thermochromic solar energy modulation.5,18,44,45 More recently, there have been https://doi.org/10.1016/j.joule.2019.06.024

2458 Joule 3, 2457–2471, October 16, 2019


reports on multifunctional coatings for glazing, in particular bilayers or nanocompo-
sites of TiO2 and VO23–5,46–48; these coatings have shown temperature-modulated
optical transmittance, antireflection, and self-cleaning properties when investigated
by monitoring decomposition of stearic acid layers, decoloration of dye solutions, or
super-wetting properties under illumination by ultraviolet (UV) light.

Here we take another, significant, step toward self-adaptive, multifunctional glazing


and demonstrate pronounced synergies between the two components in a TiO2/
VO2 bilayer coating with regard to increased indoor comfort and energy savings;
importantly we employ the full solar spectrum. Specifically, we show that the tem-
perature rise due to the absorption of light, in particular near-infrared light, in VO2
produces a very marked enhancement of the photocatalytic reactivity of TiO2 at
the same time as antireflection enhances the glazing’s luminous performance, thus
realizing and implementing our previously reported concept for boosting the reac-
tivity of a photocatalyst by purposely adjusting its temperature and water coverage
at the surface.49

RESULTS AND DISCUSSION


Morphological and Structural Characterization
Our films were deposited by reactive magnetron sputtering, which allows tuning of
their microstructures through variation of deposition parameters as discussed in
detail in the Supplemental Information. In order to obtain the desired crystal struc-
tures of the VO2 and TiO2 films, it was necessary to heat the substrates during depo-
sition and to exercise precise control of the oxygen partial pressure. Surface
morphologies for films of VO2 (53 nm), TiO2 (297 nm), and VO2 (53 nm)/TiO2
(297 nm) were obtained by use of scanning electron microscopy (SEM) and atomic
force microscopy (AFM) and are shown in Figure S2. The two sets of data were
consistent and demonstrated that the VO2 film had irregular grains with diameters
of 80 nm together with rod-like features up to 600 nm in length; the root-mean-
square surface roughness (RRMS) was found to be 34 nm. The VO2 nanorods
were principally similar to those shown in earlier work.50 The TiO2 film was smoother
and displayed grains with sizes of 60–70 nm and RRMS z 7 nm. Finally, the TiO2/VO2
bilayer film had surface features resembling those of the TiO2 film, albeit the RRMS
was larger because of the underlying VO2 and amounted to 14 nm. It is noteworthy
that the TiO2 film provided surface smoothing of the top surface of the bilayer film;
this effect is consistent with data in prior work in which surface roughness has been
seen to decrease after sputter deposition of thin films of TiO2.51

X-ray diffraction (XRD) and Raman data in Figure S3 show convincingly that the VO2
and TiO2 films have very good phase purity and crystallinity. Crystallite sizes were
extracted by means of Scherrer analysis applied to the VO2 (011) and TiO2 (101)
peaks; these sizes were 20 nm for VO2 and 15 nm for TiO2. The latter value is
consistent with the size deduced from the Eg peak at 145 cm–1 in the Raman spec-
trum in Figure S3B.52

Considering now the TiO2/VO2 bilayer film, the XRD data (Figure S3A) show diffrac-
tion peaks due to monoclinic VO2 and anatase TiO2 and additional small peaks at
2w = 36.1 and 2w = 41.3 whose positions agree with those of rutile TiO2 (PDF
Card 21–1276). The occurrence of some rutile in the TiO2/VO2 sample, but not in
the TiO2 film, can be explained by the presence of the underlying VO2 film for the
bilayer samples. The TiO2 layer was deposited at elevated temperature, implying
that VO2 was in its metal-like phase with rutile structure, and it is plausible that

Joule 3, 2457–2471, October 16, 2019 2459


this VO2 phase promotes the formation of rutile TiO2 at the interface between the
two layers. This surmise is supported by the fact that the lattice constants of the
two materials are similar, with the rutile TiO2 lattice constant being slightly larger.53
It should also be noted that rutile TiO2 has previously been used as a growth
template for VO2 films.54 The rutile phase was, however, not detected by Raman
spectroscopy (Figure S3B), which is known to be more sensitive than XRD for docu-
menting the phase composition in TiO2,55 and hence the amount of rutile in the film
is small and presumably confined only to the interface between TiO2 and VO2 (where
the Raman signal is significantly damped).

Optical Properties and Thermochromism


Spectral transmittance and reflectance, T(l) and R(l), were measured in the 300 < l <
2,500 nm wavelength range corresponding to solar irradiation and spectral absorp-
tance, A(l), was obtained from

A(l) = 1 – T(l) – R(l). (Equation 1)

Data for T(l), R(l), and A(l) are shown in Figures 1A–1C, respectively, for the VO2 and
TiO2/VO2 films at 25 and 90 C. Additional data on A(l) for the TiO2 film are reported
in Figure 1C.

The VO2 sample displays thermochromic features that are characteristic for thin
(50 nm) sputter- deposited films.23,25 Specifically, T(l) is larger for t < tc than for
t > tc at l > 900 nm; the difference is increased toward longer wavelengths, and
at l = 2,500 nm the transmittance is 80% for t < tc and 25% for t > tc. In the lu-
minous range at 400 < l < 700 nm, the transmittance is rather temperature indepen-
dent and decreases toward shorter wavelengths. The reflectance data for the VO2
film are less conspicuous but demonstrate that R(l) is smaller for t < tc than for
t > tc at l > 1,500 nm, as expected for a metallic-like material.

Temperature-dependent modulation of luminous and solar transmittance was ob-


tained from
R
4lum;sol ðlÞTðl; tÞdl
Tlum;sol ðtÞ = R ; (Equation 2)
4lum;sol ðlÞdl

where 4lum is the spectral sensitivity of the human eye56 and 4sol is the AM 1.5 solar
irradiance spectrum.57

The difference in luminous and solar transmittance, DTlum,sol, for temperatures


below and above tc is then readily obtained from

DTlum,sol = Tlum,sol(t < tc) – Tlum,sol(t > tc). (Equation 3)

Table 1 gives specific values for the VO2 film and demonstrates that DTlum is slightly
negative whereas DTsol = 4.3%.

The optical transmittance data reported in Figure 1 for the TiO2/VO2 bilayer sample
indicates that the TiO2 film enhances T(l) and introduces clear-cut optical interfer-
ence. The optical properties of TiO2 films depend on deposition parameters and
thickness, and the refractive index was 2.3 for films prepared under conditions
similar to those in the present work,58 whereas the refractive index for VO2 films is
2.8 at t < tc.18 The top layer therefore serves essentially as a three-quarter-
wavelength antireflection (AR) coating for mid-solar-spectrum and mid-luminous

2460 Joule 3, 2457–2471, October 16, 2019


Figure 1. Optical Properties of Films
Spectral Transmittance (A) reflectance (B) and absorptance (C) at room temperature (below the thermochromic ‘‘critical’’ temperature t c ) and at 90  C (at
t > t c ) for films of VO2 (53 nm) and TiO 2 (297 nm)/VO2 (53 nm). (C) also shows spectral absorptance for a film of TiO 2 (297 nm). (D) shows transmittance at a
wavelength of 2,000 nm recorded for increasing and decreasing temperature, indicated by arrows, for films of VO 2 (53 nm) and TiO 2 (297 nm)/VO 2
(53 nm).

radiation. This layer may optically be somewhat less efficient than a conventional
quarter-wavelength coating, but the TiO2 thickness must not be too small if it is to
be used as a photocatalytic film since the penetration depth for the photocatalyti-
cally active UV radiation is of the order 1 mm for nanostructured TiO2 at
l z 370 nm.59 Even though the AR treatment is not optimal, it is clear that the
TiO2 coating nevertheless is effective for boosting both Tlum, which is enhanced to
47% at t < tc and 48% at t > tc, and Tsol which is correspondingly enhanced to
as much as 53% and 44%. Importantly, DTsol = 8.8% for the TiO2/VO2 bilayer,
i.e., twice as large as for the single-layer VO2 film. Our data are in good qualitative
agreement with earlier work on VO2 films AR coated by SiO244,60,61 or TiO245,60; on
three-layer structures in which VO2 was positioned between metal oxides such as
TiO2, SiO2, or In2O3:Sn4,18,62–64; and on VO2 nanoparticles embedded in a SiO2
matrix.65 The visible optical properties of the TiO2/VO2 film change only marginally
at tc, which implies that the thermochromic transition is essentially invisible.

Joule 3, 2457–2471, October 16, 2019 2461


Table 1. Luminous and Solar Transmittance Below and Above the ‘‘Critical’’ Temperature tc, and
Corresponding Transmittance Modulation, for Films of VO2 (53 nm) and TiO2 (297 nm)/VO2
(53 nm)
Film Tlum(25 C) Tlum(90 C) DTlum (%) Tsol(25 C) Tsol(90 C) DTsol (%)
Material (%) (%) (%) (%)
VO2 36.3 39.1 -2.8 41.4 37.1 4.3
TiO2/VO2 47.3 48.0 -0.7 53.2 44.4 8.8

Figure 1C also reports spectral absorptance for the TiO2/VO2 bilayer and compares
these results with those for the TiO2 film. The absorptance of the TiO2 film is negli-
gible in almost the entire studied spectral range, but absorption sets in at the short-
est wavelengths and corresponds to an optical band gap of 3.3 eV as derived via
Tauc-type analysis.66 Clearly the absorptance is much larger in the thermochromic
bilayer than in the TiO2 film, especially in the near-infrared region at 700 < l <
2,500 nm, owing to the influence of VO2. This difference is essential for understand-
ing the photocatalytic efficiency, as discussed below. Furthermore, most of the UV
radiation is blocked, which is beneficial for human health.

The temperature range for the thermochromic transition was studied by measuring
optical transmittance at l = 2,000 nm, where the temperature-induced change is
particularly large, for films of VO2 and TiO2/VO2. Figure 1D shows that the transmit-
tance change occurs in an interval around 68 C, i.e., at the ‘‘critical’’ temperature
for bulk VO216; the transition exhibits hysteresis and has a mid-transition width of
12 C. It is well known that strain and crystallite size affect tc,53,63,67 and hence
our data indicate that the VO2 film has a relaxed structure with good crystallinity;
this result is consistent with the occurrence of a distinct VO2 (011) peak in XRD (Fig-
ure S3A). The change in the appearance of the hysteresis curve in Figure 1D with the
addition of a TiO2 layer is attributed to a difference in strain exerted on the VO2 film
as a result of lattice mismatch between the two materials.53

Photocatalysis
Photocatalytic properties of the thin-film samples were evaluated by use of stearic
acid (SA), which is a common probe molecule for assessing the photocatalytic and
self-cleaning properties of TiO2 films.68–70 Importantly, the photocatalytic decom-
position kinetics is not crucially dependent on SA coverage.68,69,71 In the present
work, SA was deposited by spin coating, which previously has been shown to create
thin uniform layers on TiO2 surfaces.72,73 The long carbon chain in the SA molecule
makes it a good choice for monitoring photocatalytic activity by use of Fourier trans-
form infrared (FTIR) spectroscopy, and the C–H stretching bonds in the molecule
give rise to three distinct vibration bands in the 2,700–3,000-cm–1 wavenumber re-
gion, as illustrated in Figure 2 for SA coated onto TiO2 and TiO2/VO2. Specifically,
the asymmetric, nas(CH), and symmetric, ns(CH), stretching modes of the CH2 group
lie at 2,922 and 2,854 cm–1, respectively, and the in-plane stretch of the CH3 group
lies at 2958 cm–1.69 Illumination of the samples by using a filtered (AM 1.5 solar ra-
diation) and unfiltered Xe lamp (see the Supplemental Information for details)
decreased the intensity of the n(CH) bands as a function of illumination time.

Figure 2 reports FTIR data and demonstrates that the degradation rate is much faster
for the TiO2/VO2 bilayer coating than for the TiO2 film and that this difference pre-
vails both for filtered and unfiltered Xe-lamp illumination. The data were taken with
samples mounted in a holder maintained at a temperature ts of 25 C, and additional
data (not shown) were recorded at 80 C. Furthermore, data were taken with a UV-
LED source irradiating samples at 25 C (not shown). We foreshadow that the

2462 Joule 3, 2457–2471, October 16, 2019


Figure 2. Photocatalytic Properties of Films
Absorbance of stearic acid (SA) layers irradiated with a Xe lamp through solar-energy-mimicking filters (A and B) and without filters (C and D) for films of
TiO2 (297 nm) (A and C) and films of TiO 2 (297 nm)/VO2 (53 nm) (B and D). Irradiation times are shown. (E) SA coverage in nanomoles cm –2 as determined
from the integrated area, A p , of the n(CH) stretching bands versus time. (F) Logarithm of the ratio between A p and the initial integrated area, A 0 , for films
of TiO 2 and TiO2 /VO2 versus time. Data in (E) and (F) were obtained with simulated solar light irradiation (triangles) and with unfiltered Xe-lamp
irradiation (circles). Straight lines in (F) were used to evaluate initial photodegradation rates.

Joule 3, 2457–2471, October 16, 2019 2463


irradiation from the Xe lamp increases the actual temperature of the films, as dis-
cussed in the Supplemental Information; this effect will be important for analyzing
the data.

The degradation rate for SA was quantified by studying the integrated peak area, Ap, of
the n(CH) vibrations in the range of 2,700–3,000 cm–1 (which is proportional to the SA
concentration) as a function of illumination time t. It has previously been shown that 1
integrated absorbance unit, 1 A.U. (h 1 cm–1), corresponds to A1.A.U. = 9.7 3 1015 SA
molecules cm–2 per A.U. for SA layers prepared in a manner similar to ours.68 In the
present experiments we obtain Ap(t = 0) h A0 = 1.29, 0.95, 0.97, and 0.97 cm–1, respec-
tively, for the various films in Figures 2A–2D. Thus the SA layer thicknesses for our films
correspond to about 18 monolayers, except for the film giving the data in Figure 2A
whose thickness is 24 monolayers; these results were obtained by using the bulk density
and molar mass of SA (941 kg m–3 and 284.48 g mole–1, respectively) and a monolayer
thickness of 2.6 nm.70

Figures 2E and 2F display data on the photodegradation rate of SA. This rate has
previously been reported to follow zero-order kinetics for thick layers and first-order
kinetics in the case of thin films.69–71 Although rather thick layers were inferred from
literature-calibrated FTIR data, as mentioned above, the initial decomposition ki-
netics was found to be consistent with first-order kinetics, i.e.,

Ap ðtÞ = A0 ekt ; (Equation 4)

where k is photodegradation rate constant.

However, inspecting the plot of ln(Ap/A0) versus t (Figure 2F), reveals that after about
10 to 15 min of illumination, there was a change in the reaction kinetics to a slower
degradation rate that also obeys first-order reaction kinetics, although with a smaller
rate constant. This change in reaction kinetics occurred after shorter illumination
times for samples exhibiting higher degradation rates, which is consistent with faster
depletion of radicals and/or readily degradable C–H moieties. We attribute the
change in reaction kinetics to a combination of depletion of hydroxyl radicals at
the interface of the SA film and accumulation of strongly bonded residues that inhibit
reactions in the SA.74 Therefore only the first, linear regime in Figure 2F was used to
determine rate constants.

The initial photodegradation rate of SA molecules, rSA in units of moles per unit area
and time, is defined as the negative derivative of the first-order reaction kinetics
(Equation 4) evaluated at t = 0. Thus

kA1:A:U: A0
rSA = ; (Equation 5)
NA
where NA is Avogradro’s constant.

Specific data on k and rSA are given in Table 2 for samples at ts being 25 C and 80 C
and demonstrate that the decomposition of SA molecules was more efficient on the
TiO2/VO2 bilayer than on the single-layer TiO2 film under both types of illumination.
Also given in Table 2 is the quantum yield (QY) determined as the ratio between the
initial photodegradation rate and the number of absorbed photons in the TiO2 film
according to

rSA NA
QY = R 2500 ; (Equation 6)
200
Fph ðlÞAðlÞdl

2464 Joule 3, 2457–2471, October 16, 2019


Table 2. Photodegradation Rate Constant k (min–1), Initial Photodegradation Rate rSA (nmol min–1 cm–2) and Quantum Yield QY (Converted
Molecules per Absorbed Photon) for Stearic Acid on Films of TiO2 (297 nm) and TiO2 (297 nm)/VO2 (53 nm) Kept at the Nominal Temperatures ts in
Synthetic Air with Zero Relative Humidity
Film Xe Lamp (AM 1.5) Xe Lamp (No Filter) UV-LED (370 nm)
Material   
ts = 25 C ts = 80 C ts = 25 C ts = 25 C
3 3 3
k rSA QY310 k rSA QY310 k rSA QY310 k rSA QY3103
TiO2 0.056 G 1.17 G 2.1 G 0.2 0.029 G 0.82 G 1.5 G 0.2 0.103 G 1.60 G 0.50 G 0.05 0.013 G 0.22 G 18 G 2
0.011 0.24 0.006 0.17 0.021 0.32 0.002 0.05
TiO2/VO2 0.101 G 1.57 G 2.8 G 0.3 0.054 G 1.37 G 2.5 G 0.3 0.189 G 2.96 G 0.93 G 0.09 0.015 G 0.18 G 15 G 2
0.020 0.32 0.011 0.28 0.038 0.59 0.002 0.04

Irradiation was performed by use of the indicated light sources, whose irradiance spectra can be found in Figure S1.
The error bars show the experimental variations of photon flux, stearic acid surface coverage, and absorptance in Equations 4–6.

where Fph(l) is the photon flux from the light source, the integration limits are in
nanometers, and A(l) is obtained from Equation 1.

Data in Table 2 lead to the following observations: (1) the reaction proceeds
faster—manifested by larger degradation rate, initial degradation rate, and quantum
yield—on the TiO2/VO2 bilayer sample than on the single-layer TiO2 sample when
broadband Xe-light sources were used; (2) SA degradation progresses at higher rate
with unfiltered broadband illumination than with AM 1.5 filtered solar-light-mimicking
Xe-lamp illumination; (3) SA degradation proceeds at lower rate for elevated substrate
temperature than at room temperature; (4) for UV-LED illumination, the reaction rate is
slightly lower on the TiO2/VO2 bilayer sample than on the single-layer TiO2 sample;
and (5) the QY depends on which light source was used and was largest for UV-LED
illumination and lowest for unfiltered Xe-lamp illumination.

The difference in QY for the various light sources is related to the spectral distribu-
tion and photon flux of the illumination. Simulated solar light and the unfiltered Xe
lamp emit light whose intensity increases with increasing wavelength in the UV and
near-UV regions. Thus a large fraction of the photon flux stems from photons with
energies close to and below the band gap energy of TiO2, whereas the UV-LED
mainly emits photons with energy well above this band gap. This difference in
phonon flux from the illumination sources impacts the QY calculated by Equation 6
since high-energy photons give larger photocatalytic yield than photons with low en-
ergy.75,76 Furthermore, it has been shown from measurements utilizing different UV
light intensities that the QY decreases with increasing number of adsorbed pho-
tons.77 Thus the highest QY was obtained with UV-LED illumination, which provides
low photon flux with high-energy photons, whereas the lowest QY was found with
unfiltered Xe illumination, which gives a large photon flux and a broad distribution
of photon energy so that there is a large fraction of photons with energies close to
and below the band gap energy.

The somewhat higher values of rSA and QY for the single-layer TiO2 film compared to
the corresponding values for the TiO2/VO2 bilayer film for measurements done with
weak UV-LED illumination—which is opposite to the results for broadband Xe-light
sources—may indicate that differences of structure and morphology noted in SEM,
AFM, and XRD influence the reactivity. Figure S3A shows a slightly larger ratio of the
<004> to <101> diffraction peaks in the single-layer TiO2 film than in the TiO2/VO2
bilayer film, which may account for this difference since (001) facets are reported to
be more reactive than (101) facets,78 and in particular sputter-deposited films with
<001> orientation have been shown to be more reactive than randomly oriented
films.79 More important for our results is, however, that this higher photocatalytic

Joule 3, 2457–2471, October 16, 2019 2465


activity of the TiO2 film compared to the data for the TiO2/VO2 bilayer film under UV-
LED illumination implies that the minor occurrence of rutile-phase TiO2 in the TiO2/
VO2 bilayer films, which could be noted in the XRD analysis (Figure S3A), does not
contribute to the rate enhancement of the bilayer films under Xe-lamp illumination.
This result provides support for the interpretation above that the slight rutile-
anatase phase inter-mixing is confined to the buried TiO2/VO2 interface and does
not affect the reactivity, since it has been reported that rutile-anatase nanoparticle
mixtures can display higher photocatalytic activity than phase-pure anatase.80–82

The lower degradation rate at ts = 80 C than at ts = 25 C with the same illumination
and humidity is per se not surprising, since it is well known that both temperature and
humidity influence gas-phase photocatalysis.74,83,84 Low relative humidity or high
photocatalyst temperature will dry out the TiO2 surface, thereby opening up reac-
tion routes that lead to stable intermediate surface species causing blocking of
active sites, whereas too high relative humidity will create a water layer on the sur-
face that limits reactant transport. At intermediate conditions, a balance can be ob-
tained between free adsorption sites, O2 coverage, and water coverage on the
surface.49,84 Thus a TiO2 surface at 80 C is expected to be considerably dehydrated
compared to the case of such a surface at ambient temperature, which accounts
for the lower reaction rate. The importance of hydroxyls on the surface have been
shown by UHV studies, wherein the lack of hydroxyls on the photocatalyst surface
is a limiting factor for the photocatalytic degradation.36,85 This effect of the hy-
droxyls was corroborated in separate measurements done at ts = 80 C using a
slightly humid airflow with a relative humidity of 0.8% (corresponding to a water
coverage of about 0.02 monolayers at steady-state conditions49) at 80 C, which,
as expected,49,84 yielded higher photocatalytic activity than for measurements in
dry air (Table S2).

We must therefore draw the following conclusions: The higher magnitudes of rSA and
QY for the TiO2/VO2 bilayer samples than for the single-layer TiO2 samples under
broadband Xe-lamp illumination stem from different optical absorptance of the films
(Figure 1C). For the bilayers, the optical absorptance of visible and near-infrared (IR)
light in the VO2 film leads to heating of the TiO2 photocatalyst through thermal con-
duction and thus accelerates SA degradation compared to the situation for the sin-
gle-layer TiO2 samples which have optical absorptance only in the UV. This effect of
photocatalyst heating was, as expected, most pronounced when unfiltered light
from the Xe lamp was employed. Here we note that the degradation rate is expected
to depend linearly on photon intensity in our range of photon fluxes,86 and therefore
the larger flow of photons with energies higher than the TiO2 bandgap for the unfil-
tered Xe lamp contributes to increased SA degradation. For the TiO2 film, rSA
increased from 1.17 to 1.60 nmole min–1 cm–2 as a result of this increased flux of
UV photons. However, for the bilayer sample, the increase in rSA was still larger, spe-
cifically from 1.57 to 2.96 nmole min–1 cm–1. Thus, we can conclude that the primary
reason for this increase in rSA is the heating of the photocatalyst as a consequence of
the higher absorption of photons in the visible and near-IR wavelength ranges.

Finally, we assess the reaction data in Table 2 with due consideration of actual sam-
ple temperature, ts*, which is larger than ts owing to the illumination from the broad-
band radiation sources, as elaborated in the Supplemental Information. Figure 3
shows QY as a function of ts* and strongly suggests that the heating due to optical
absorption in the VO2 film enhances the temperature as well as the photocatalytic
reaction rate and QY of the TiO2/VO2 bilayer film. Specifically, Table 2 shows an
enhancement of rSA and QY of 34% at ts = 25 C and about 67% at ts = 80 C

2466 Joule 3, 2457–2471, October 16, 2019


Figure 3. Quantum Yield as a Function of Actual Substrate Temperature, ts*, for the Illumination
Sources Shown in the Legend and Indicated by the Different Symbols
The data pertain to films of TiO 2 (297 nm) (blue symbols) and TiO 2 (297 nm)/VO 2 (53 nm) (red
symbols). Note the logarithmic scale on the ordinate axis. The Error Bars Show the Experimental
Variations of Photon Flux, Stearic Acid Surface Coverage and Absorptance in Equations 4–6.

compared with corresponding numbers for a single-layer TiO2 film in the case of AM
1.5-filtered radiation. When unfiltered Xe-lamp illumination is employed, the corre-
sponding enhancement is about 86% at ts = 25 C. In the Supplemental Information,
we calculated that the temperature rise was about 17 C with AM 1.5 filtered Xe illu-
mination and about 35 C for unfiltered Xe-lamp illumination for the bilayer films in
the in situ measurement cell. Data on the temperature increase have some uncer-
tainties associated with thermal losses due to convection and thermal conductance
in the in situ cell but suffice for semi-quantitative arguments.

For a thermally activated degradation reaction, the reaction rate, r, is expected to


increase as a function of temperature according to rfeEa =kB t , where Ea is the activa-
tion energy for photo-oxidation and kB is Boltzmann’s constant. Measurements on
thick SA layers on compact TiO2 films, intended for self-cleaning window applica-
tions, have obtained an activation energy of 0.2 eV for the temperature range
used here;70 this activation energy corresponds to a doubling of the reaction rate
for a temperature increase of 30 C, which is very close to the data we obtain for
the bilayer film as shown in Table S1. In principle, an increased photocatalyst
temperature can also enhance the probability for indirect interband transitions,
and the Arrhenius-type analysis used here may disguise contributions from such
photo-thermal effects. We can disregard a temperature-dependent decrease of
the optical band gap at the temperatures we use here.87 As noted above, a temper-
ature-dependent change of water coverage due to desorption can also alter the
photo-oxidation reaction rate84 because the presence of water can modify the over-
all reaction between SA and radicals derived from water and O2 on the surface if the
reaction is governed by the diffusivity of the reactants. The latter mechanism would,
however, give a temperature dependence of the reaction that is proportional to the
relative ratio of the temperature of the heated and non-heated catalyst film. Here we
find good agreement between experimental data, calculated temperature rise of the
film due to illumination, and expected thermal activation of the degradation rate.
We conclude that the SA photodegradation reaction on the TiO2 film is not
controlled by reactant diffusion in a water layer but an effect of heating due to light
absorption in the thermochromic VO2 film.

Joule 3, 2457–2471, October 16, 2019 2467


It should be emphasized that the photocatalytic and thermochromic components in
the TiO2/VO2 bilayer coatings operate in unison and that, furthermore, the luminous
performance is enhanced by the photocatalytic top layer which may also serve as
protection of the underlying thermochromic layer against further oxidation.5 The
various functionalities thus operate synergistically, which points toward new avenues
to accomplish superior glazing for energy-efficient buildings with healthy and
comfortable indoor environments.

Conclusions
By combining a thermochromic VO2 film with a photocatalytic TiO2 layer, a multi-
functional coating is obtained whose solar energy modulation, luminous transmit-
tance, and indoor air cleaning ability are enhanced compared to the properties
obtainable with the corresponding single-layer films. The thermochromic layer is
used to modulate the solar energy inflow into a building while maintaining high lu-
minous transmittance. The addition of a TiO2 photocatalytic coating on top of the
VO2 film not only adds indoor air cleaning capability but the TiO2 layer also serves
as an antireflection and protective coating that enhances both luminous transmit-
tance and solar energy modulation. Additionally, the thermochromic film below
the photocatalytic coating boosts the photocatalytic properties, which is a result
of heating of the photocatalyst due to optical absorption, predominantly in the
near-IR range, in the thermochromic film.

A glazing based on the principles delineated in this paper should use at least two
glass panes and allow indoor air to be circulated between these panes preferably
by natural convection.88 The outer pane or panes should be of iron-free glass with
high UV transmittance; the TiO2/VO2 bilayer film should be on one of the surfaces
facing the air gap between the panes, as depicted in the graphical abstract; and a
low-emittance coating is desired on the other surface facing the air gap in order
to provide improved thermal insulation.

In conclusion, the utilization of photocatalytic and thermochromic bilayer coatings,


as demonstrated here, paves the way toward more energy-efficient fenestration
and holds the potential to clean indoor air and at the same time reduce the
need for air conditioning and ventilation. These benign properties are a result of
the synergistic effects between the two coatings, which result in improved photo-
catalytic efficiency, luminous transmittance, and solar energy modulation of the
glazing.

EXPERIMENTAL PROCEDURES
All details of the experimental procedure can be found in the Supplemental
Information.

SUPPLEMENTAL INFORMATION
Supplemental Information can be found online at https://doi.org/10.1016/j.joule.
2019.06.024.

ACKNOWLEDGMENTS
This work was funded by the Swedish Research Council grant agreement no. 2016-
05904 and the European Research Council under the European Community’s
Seventh Framework Program (FP7/2007–2013, ERC grant agreement 267234
[‘‘GRINDOOR’’]).

2468 Joule 3, 2457–2471, October 16, 2019


AUTHOR CONTRIBUTIONS
Y.J. and A.M. jointly conducted the experiments. A.M. wrote the original draft of the
manuscript. L.Ö. conceived the original idea and supervised the project. G.A.N. and
C.G.G. provided advice and expertise. All authors contributed to the final version of
the manuscript.

DECLARATION OF INTERESTS
The authors declare no conflicts of interest.

Received: March 22, 2019


Revised: May 13, 2019
Accepted: June 23, 2019
Published: July 17, 2019

REFERENCES
1. Leech, J.A., Nelson, W.C., Burnett, R.T., Aaron, 10. C.M. Lampert, and C.G. Granqvist, eds. (1990). 21. Dietrich, M.K., Kuhl, F., Polity, A., and Klar, P.J.
S., and Raizenne, M.E. (2002). It’s about time, a Large-area chromogenics. materials and (2017). Optimizing thermochromic VO2 by co-
comparison of Canadian and American time– devices for transmittance control (The doping with W and Sr for smart window
activity patterns. J. Expo. Anal. Environ. International Society of Optical Engineering). applications. Appl. Phys. Lett. 110.
Epidemiol. 12, 427–432.
11. Saeli, M., Piccirillo, C., Parkin, I.P., Binions, R., 22. Khan, K.A., and Granqvist, C.G. (1989).
2. Smith, G.B., and Granqvist, C.G. (2011). Green and Ridley, I. (2010). Energy modelling studies Thermochromic sputter-deposited vanadium
nanotechnology, solutions for sustainability of thermochromic glazing. Energy Build. 42, oxyfluoride coatings with low luminous
and energy in the built environment (CRC 1666–1673. absorptance. Appl. Phys. Lett. 55, 4–6.
Press).
12. Hoffmann, S., Lee, E.S., and Clavero, C. (2014). 23. Li, S.-Y., Niklasson, G.A., and Granqvist, C.G.
3. Wilkinson, M., Kafizas, A., Bawaked, S.M., Examination of the technical potential of near- (2014). Thermochromic undoped and Mg-
Obaid, A.Y., Al-Thabaiti, S.A., Basahel, S.N., infrared switching thermochromic windows for doped VO2 thin films and nanoparticles,
Carmalt, C.J., and Parkin, I.P. (2013). commercial building applications. Sol. Energy optical properties and performance limits for
Combinatorial atmospheric pressure chemical Mater. Sol. Cells 123, 65–80. energy efficient windows. J. Appl. Phys. 115.
vapor deposition of graded TiO2–VO2 mixed-
phase composites and their dual functional 13. Warwick, M.E.A., and Binions, R. (2014). 24. Li, M., Magdassi, S., Gao, Y., and Long, Y.
property as self-cleaning and photochromic Advances in thermochromic vanadium dioxide (2017). Hydrothermal synthesis of VO2
window coatings. ACS Comb. Sci. 15, 309–319. films. J. Mater. Chem. A 2, 3275–3292. polymorphs, advantages, challenges and
prospects for the application of energy
4. Powell, M.J., Quesada-Cabrera, R., Taylor, A., 14. Granqvist, C.G. (2016). Recent progress in efficient smart windows. Small 13.
Teixeira, D., Papakonstantinou, I., Palgrave, thermochromics and electrochromics, A brief
R.G., Sankar, G., and Parkin, I.P. (2016). survey. Thin Solid Films 614, 90–96. 25. Ji, Y.-X., Li, S.-Y., Niklasson, G.A., and
Intelligent multifunctional VO2/SiO2/TiO2 Granqvist, C.G. (2014). Durability of
15. Chang, T., Cao, X., Dedon, L.R., Long, S., thermochromic VO2 thin films under heating
coatings for self-cleaning, energy-saving Huang, A., Shao, Z., Li, N., Luo, H., and Jin, P.
window panels. Chem. Mater. 28, 1369–1376. and humidity, effect of Al oxide top coatings.
(2018). Optical design and stability study for Thin Solid Films 562, 568–573.
ultrahigh-performance and long-lived
5. Top, I., Binions, R., Warwick, M.E.A., Dunnill,
vanadium dioxide-based thermochromic 26. Ji, H., Liu, D., Zhang, C., and Cheng, H. (2018).
C.W., Holdynski, M., and Abrahams, I. (2018).
coatings. Nano Energy 44, 256–264. VO2/ZnS core-shell nanoparticle for the
VO 2 /TiO 2 bilayer films for energy efficient
windows with multifunctional properties. adaptive infrared camouflage application with
16. Morin, F.J. (1959). Oxides which show a metal- modified color and enhanced oxidation
J. Mater. Chem. C 6, 4485–4493. to-insulator transition at the neel temperature. resistance. Sol. Energy Mater. Sol. Cells 176,
Phys. Rev. Lett. 3, 34–36. 1–8.
6. United Nations (2009). Buildings and climate
change, summary for decision makers (U.N 17. Li, S.-Y., Niklasson, G.A., and Granqvist, C.G. 27. Sundell, J. (2004). On the history of indoor air
Development Program), Sustainable (2012). Thermochromic fenestration with VO2- quality and health. Indoor Air 14, 51–58.
Buildings & Climate Initiative. based materials, three challenges and how
they can be met. Thin Solid Films 520, 3823– 28. Gligorovski, S., and Abbatt, J.P.D. (2018). An
7. U.S. Department of Energy (2011). Buildings 3828. indoor chemical cocktail. Science 359, 632–633.
energy data book, 2011 (United States
Department of Energy). 18. Mlyuka, N.R., Niklasson, G.A., and Granqvist, 29. Brown, S.K., Sim, M.R., Abramson, M.J., and
C.G. (2009). Thermochromic VO2-based Gray, C.N. (1994). Concentrations of volatile
8. Stocker, T.F., Qin, D., Plattner, G.-K., Tignor, multilayer films with enhanced luminous organic compounds in indoor air – a review.
M., Allen, S.K., Boschung, J., Nauels, A., Xia, Y., transmittance and solar modulation. Phys. Indoor Air 4, 123–134.
Bex, V., and Midgley, P.M. (2013). Climate Status Solidi (a) 206, 2155–2160.
change 2013-The physical science Basis, 30. Wolkoff, P., Wilkins, C.K., Clausen, P.A., and
working group I contribution to the fifth 19. Xu, G., Jin, P., Tazawa, M., and Yoshimura, K. Nielsen, G.D. (2006). Organic compounds in
assessment report of the intergovernmental (2004). Optimization of antireflection coating office environments - sensory irritation, odor,
panel on climate change (Cambridge for VO2-based energy efficient window. Sol. measurements and the role of reactive
University Press). Energy Mater. Sol. Cells 83, 29–37. chemistry. Indoor Air 16, 7–19.
9. Altomonte, S., Saadouni, S., Kent, M.G., and 20. Mlyuka, N.R., Niklasson, G.A., and Granqvist, 31. Jones, A.P. (1999). Indoor air quality and health.
Schiavon, S. (2017). Satisfaction with indoor C.G. (2009). Mg doping of thermochromic VO2 Atmos. Environ 33, 4535–4564.
environmental quality in BREEAM and non- films enhances the optical transmittance and
BREEAM certified office buildings. Archit. Sci. decreases the metal-insulator transition 32. World Health Organization (1982). Indoor air
Rev. 60, 343–355. temperature. Appl. Phys. Lett. 95. pollutants, exposure and health effects (Euro

Joule 3, 2457–2471, October 16, 2019 2469


reports and studies no. 78) (WHO Regional 49. Österlund, L., Mattsson, A., Brischetto, M., 64. Sun, G.Y., Cao, X., Zhou, H., Bao, S., and Jin, P.
Office of Europe). Byberg, J.J., Stefanov, B.I., Ji, Y.-X., and (2017). A novel multifunctional thermochromic
Niklasson, G.A. (2018). Spectral selective solar structure with skin comfort design for smart
33. World Health Organization (1986). Indoor air light enhanced photocatalysis, TiO2/TiAlN window application. Sol. Energy Mater. Sol.
quality research (Euro Reports and Studies No Bilayer Films. Top. Catal. 61, 1607–1614. Cells 159, 553–559.
103) (World Health Organization).
50. Li, S.-Y., Namura, K., Suzuki, M., Niklasson, 65. Zhao, L., Miao, L., Liu, C., Li, C., Asaka, T., Kang,
34. Tham, K.W. (2016). Indoor air quality and its G.A., and Granqvist, C.G. (2013). Y., Iwamoto, Y., Tanemura, S., Gu, H., and Su,
effects on humans—a review of challenges and Thermochromic VO2 nanorods made by H. (2014). Solution-processed VO2-SiO2
developments in the last 30 years. Energy sputter deposition, Growth conditions and composite films with simultaneously enhanced
Build. 130, 637–650. optical modeling. J. Appl. Phys. 114. luminous transmittance, solar modulation
ability and anti-oxidation property. Sci. Rep. 4,
35. Hodgson, A.T., Sullivan, D.P., and Fisk, W.J. 51. Duparré, A., and Walther, H.G. (1988). Surface 7000.
(2005). Evaluation of ultra-violet photocatalytic smoothing and roughening by dielectric thin
oxidation (UVPCO) for indoor air applications, film deposition. Appl. Opt. 27, 1393–1395. 66. Tauc, J., Grigorovici, R., and Vancu, A. (1966).
conversion of volatile organic compounds at Optical properties and electronic structure of
low part-per-billion concentrations (National 52. Lejon, C., and Österlund, L. (2011). Influence of amorphous germanium. Phys. Stat. Sol. b 15,
Laboratory). phonon confinement, surface stress, and 627–637.
zirconium doping on the Raman vibrational
36. Fujishima, A., Zhang, X., and Tryk, D. (2008). properties of anatase TiO2 nanoparticles. 67. Suh, J.Y., Lopez, R., Feldman, L.C., and
TiO2 Photocatalysis and related surface J. Raman Spectrosc. 42, 2026–2035. Haglund, R.F., Jr. (2004). Semiconductor to
phenomena. Surf. Sci. Rep. 63, 515–582. metal phase transition in the nucleation and
53. Muraoka, Y., and Hiroi, Z. (2002). Metal- growth of VO2 nanoparticles and thin films.
37. Hanaor, D.A.H., and Sorrell, C.C. (2011). Review insulator transition of VO2 thin films grown on J. Appl. Phys. 96, 1209–1213.
of the anatase to rutile phase transformation. TiO2 (001) and (110) substrates. Appl. Phys.
J. Mater. Sci. 46, 855–874. Lett. 80, 583–585. 68. Mills, A., and Wang, J. (2006). Simultaneous
monitoring of the destruction of stearic acid
38. Mattsson, A., and Österlund, L. (2010). 54. Kawatani, K., Kanki, T., and Tanaka, H. (2014). and generation of carbon dioxide by self-
Adsorption and photoinduced decomposition Formation mechanism of a microscale domain cleaning semiconductor photocatalytic films.
of acetone and acetic acid on anatase, and effect on transport properties in strained J. Photochem. Photobiol. A 182, 181–186.
brookite, and Rutile TiO 2 Nanoparticles. VO2 thin films on TiO2(001). Phys. Rev. B 90.
J. Phys. Chem. C 114, 14121–14132. 69. Sawunyama, P., Jiang, L., Fujishima, A., and
55. Li, J.-G., and Ishigaki, T. (2004). Brookite /
Hashimoto, K. (1997). Photodecomposition of a
39. Di Paola, A., Garcı́a-López, E., Marcı́, G., and rutile phase transformation of TiO2 studied
Langmuir-Blodgett film of stearic acid on TiO2
Palmisano, L. (2012). A survey of photocatalytic with monodispersed particles. Acta Mater. 52,
film observed by in situ atomic force
5143–5150.
materials for environmental remediation. microscopy and FT-IR. J. Phys. Chem. B 101,
J. Hazard. Mater. 211–212, 3–29. 56. Wyszecki, G., and Stiles, W.G. (1982). Colour 11000–11003.
science, concepts and methods, quantitative
40. Schneider, J., Matsuoka, M., Takeuchi, M., 70. Mills, A., Lepre, A., Elliott, N., Bhopal, S.,
data and formulae, Second edition (Wiley).
Zhang, J., Horiuchi, Y., Anpo, M., and Parkin, I.P., and O’Neill, S.A. (2003).
Bahnemann, D.W. (2014). Understanding TiO2 57. American Society for Testing Materials (2008). Characterization of photocatalyst Pilkington
photocatalysis, Mechanisms and materials. Standard tables of reference solar spectral ActivTM, a reference film photocatalyst?
Chem. Rev. 114, 9919–9986. irradiance, direct normal and hemispherical on J. Photochem. Photobiol. A 160, 213–224.
37 tilted surface, 14 (ASTM), G173–G103.
41. Jorgenson, G.V., and Lee, J.C. (1986). Doped 71. Ghazzal, M.N., Barthen, N., and Chaoui, N.
vanadium oxide for optical switching films. Sol. 58. Rodrı́guez, J., Gómez, M., Ederth, J., Niklasson, (2011). Photodegradation kinetics of stearic
Energy Mater. 14, 205–214. G.A., and Granqvist, C.G. (2000). Thickness acid on UV-irradiated titania thin film
dependence of the optical properties of separately followed by optical microscopy and
42. Babulanam, S.M., Eriksson, T.S., Niklasson, sputter deposited Ti oxide films. Thin Solid Fourier transform infrared spectroscopy. Appl.
G.A., and Granqvist, C.G. (1987). Films 365, 119–125. Catal. B 103, 85–90.
Thermochromic VO2 films for energy-efficient
windows. Sol. Energy Mater. 16, 347–363. 59. Siefke, T., Kroker, S., Pfeiffer, K., Puffky, O., 72. Paz, Y., Luo, Z., Rabenberg, L., and Heller, A.
Dietrich, K., Franta, D., Ohlı́dal, I., Szeghalmi, (1995). Photooxidative self-cleaning
43. Parkin, I.P., and Palgrave, R.G. (2005). Self- A., Kley, E.-B., et al. (2016). Materials pushing transparent titanium dioxide films on glass.
cleaning coatings. J. Mater. Chem. 15, 1689– the application limits of wire grid polarizers J. Mater. Res. 10, 2842–2848.
1695. further into the deep ultraviolet spectral range.
Adv. Opt. Mater. 4, 1780–1786. 73. Remillard, J.T., McBride, J.R., Nietering, K.E.,
44. Lee, M.-H., and Cho, J.-S. (2000). Better Drews, A.R., and Zhang, X. (2000). Real time
thermochromic glazing of windows with anti- 60. Chen, Z., Gao, Y., Kang, L., Du, J., Zhang, Z., in situ spectroscopic ellipsometry studies of the
reflection coating. Thin Solid Films 365, 5–6. Luo, H., Miao, H., and Tan, G. (2011). VO2- photocatalytic oxidation of stearic acid on
based double-layered films for smart windows: titania films. J. Phys. Chem. B 104, 4440–4447.
45. Jin, P., Xu, G., Tazawa, M., and Yoshimura, K. optical design, all-solution preparation and
(2002). A VO2-based multifunctional window improved properties. Sol. Energy Mater. Sol. 74. Österlund, L. (2010). Fourier-transform infrared
with highly improved luminous transmittance. Cells 95, 2677–2684. and Raman spectroscopy of pure and doped
Jpn. J. Appl. Phys. 41, L278–L280. TiO2 photocatalysts. In On Solar Hydrogen &
61. Zhu, M., Qi, H., Wang, B., Wang, H., Zhang, D., Nanotechnology, V.L., ed. (John Wiley & Sons),
46. Chen, Z., Cao, C., Chen, S., Luo, H., and Gao, Y. and Lv, W. (2018). Enhanced visible pp. 189–238.
(2014). Crystallised mesoporous TiO2 (A). transmittance and reduced transition
J. Mater. Chem. A 2, 11874–11884. temperature for VO2 thin films modulated by 75. Coutts, J.L., Levine, L.H., Richards, J.T., and
index-tunable SiO2 anti-reflection coatings. Mazyck, D.W. (2011). The effect of photon
47. Zheng, J., Bao, S., and Jin, P. (2015). TiO2(R)/ RSC Adv. 8, 28953–28959. source on heterogeneous photocatalytic
VO2(M)/TiO2(A) multilayer film as smart oxidation of ethanol by a silica-titania
window, combination of energy-saving, 62. Jin, P., Xu, G., Tazawa, M., and Yoshimura, K. composite. J. Photochem. Photobiol. A 225,
antifogging and self-cleaning functions. Nano (2003). Design, formation and characterization 58–64.
Energy 11, 136–145. of a novel multifunctional window with VO2 and
TiO2 coatings. Appl. Phys. A 77, 455–459. 76. Xu, C., Yang, W., Ren, Z., Dai, D., Guo, Q.,
48. Li, Y., Ji, S., Gao, Y., Luo, H., and Kanehira, M. Minton, T.K., and Yang, X. (2013). Strong
(2013). Core-shell VO2@TiO2 nanorods that 63. Miller, M.J., and Wang, J. (2016). Multilayer photon energy dependence of the
combine thermochromic and photocatalytic ITO/VO2/TiO2 thin films for control of solar and photocatalytic dissociation rate of methanol on
properties for application as energy-saving thermal spectra. Sol. Energy Mater. Sol. Cells TiO2(110). J. Am. Chem. Soc. 135, 19039–
smart coatings. Sci. Rep. 3, 1370. 154, 88–93. 19045.

2470 Joule 3, 2457–2471, October 16, 2019


77. Ohko, Y., Hashimoto, K., and Fujishima, A. 81. Li, G., and Gray, K.A. (2007). The solid-solid anatase TiO2. J. Phys. Chem. B 109, 10886–
(1997). Kinetics of photocatalytic reactions interface, Explaining the high and unique 10895.
under extremely low-intensity UV illumination photocatalytic reactivity of TiO2-based
on titanium dioxide thin films. J. Phys. Chem. A nanocomposite materials. Chem. Phys. 339, 85. Linsebigler, A.L., Lu, G., and Yates, J.T. (1995).
101, 8057–8062. 173–187. Photocatalysis on TiO2 surfaces, Principles,
mechanisms, and selected results. Chem. Rev.
78. Wu, Q., Liu, M., Wu, Z., Li, Y., and Piao, L. (2012). 82. Scanlon, D.O., Dunnill, C.W., Buckeridge, J., 95, 735–758.
Is photooxidation activity of {001} facets truly Shevlin, S.A., Logsdail, A.J., Woodley, S.M.,
lower than that of facets for anatase TiO2 Catlow, C.R.A., Powell, M.J., Palgrave, R.G., 86. Ollis, D.F. (2005). .Kinetic disguises in
crystals? J. Phys. Chem. C 116, 26800–26804. Parkin, I.P., et al. (2013). Band alignment of heterogeneous photocatalysis. Top. Catal. 35,
rutile and anatase TiO2. Nat. Mater. 12, 217–223.
79. Stefanov, B.I., Niklasson, G.A., Granqvist, C.G.,
798–801.
and Österlund, L. (2015). Quantitative relation
between photocatalytic activity and degree of 87. Zhang, F., Zhang, R.-J., Zhang, D.-X., Wang,
h001i orientation for anatase TiO 2 thin films. 83. Kozlov, D.V., Vorontsov, A.V., Smirniotis, P.G., Z.-Y., Xu, J.-P., Zheng, Y.-X., Chen, L., -Huang,
J. Mater. Chem. A 3, 17369–17375. and Savinov, E.N. (2003). Gas-phase Y., Sun, R.-Z., Chen, Y., et al. (2013).
photocatalytic oxidation of diethyl sulfide over Temperature-dependent optical properties of
80. Vandermeulen, T., Mattson, A., and Österlund, TiO2, Kinetic investigations and catalyst titanium oxide thin films studied by
L. (2007). A comparative study of the deactivation. Appl. Catal. B 42, 77–87. spectroscopic ellipsometry. Appl. Phys. Exp. 6.
photocatalytic oxidation of propane on
anatase, rutile, and mixed-phase anatase-rutile 84. Hägglund, C., Kasemo, B., and Österlund, L. 88. Granqvist, C.G., and Lindqvist, S.-E. (2010).
TiO2 nanoparticles, Role of surface (2005). In situ reactivity and FTIR study of the Pollutant decomposition device. US Patent
intermediates. J. Catal. 251, 131–144. wet and dry photooxidation of propane on 7731915.

Joule 3, 2457–2471, October 16, 2019 2471

You might also like