You are on page 1of 10

Progress in Organic Coatings 113 (2017) 15–24

Contents lists available at ScienceDirect

Progress in Organic Coatings


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

Preparation of photocatalytic TiO2-based self-cleaning coatings for painted MARK


surface without interlayer
Fei Xua, Tao Wangb, HongYu Chenb, James Bohlingc, Alvin M. Mauricec, Limin Wua,

Shuxue Zhoua,
a
Department of Materials Science, State Key Laboratory of Molecular Engineering of Polymers, Advanced Coatings Research Center of Ministry of Education of China,
Fudan University, Shanghai 200433, China
b
Dow Chemical Company, No. 936, Zhangheng Road, Pudong New Area, Shanghai 201203, China
c
Dow Chemical Company, 400 Arcola Rd., Collegeville, PA, 19426, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Photocatalytic TiO2 provides a promising route to produce self-cleaning coating surfaces, but it often causes
TiO2 nanoparticles decomposition of organic substrates and other coating components. An interlayer is often employed to insulate
Self-cleaning coatings photocatalytic activity from sensitive substrate layers, complicating the application of these self-cleaning
Hybrid coatings. Herein, photocatalytic TiO2-based coatings were prepared based on an aqueous TiO2 dispersion in a
Photo-catalytic degradation
hybrid binder synthesized from tetraethyl orthosilicate and methyl trimethoxy silane via a sol-gel process.
Coatings with various levels of nano TiO2 were prepared and found to be transparent to visible light; i.e., the
clear coating could be directly cast on architectural latex coatings without changing its original appearance. The
photocatalytic coating system was composed of a TiO2-based clear top coat and an opaque latex film underneath.
The coatings were evaluated through outdoor exposure studies and accelerated weathering tests.
Results of these tests showed that the photocatalytic coatings with a TiO2 content range of 33–45% exhibited
excellent self-cleaning performance, while displaying none of the expected degradation. In fact, this nano TiO2-
based clear coating actually protected the latex film from UV-induced damage. The increased stability of the
organic film may have resulted from reduced UV transmission through the photo-catalyst containing clear coat
as well as reduced water permeability. It was demonstrated that the combination of appropriate TiO2 content
and a suitable binder are crucial for the fabrication of robust photocatalytic self-cleaning coatings for painted
surfaces without the need for an interlayer.

1. Introduction efficiency of TiO2 nanoparticles will be lowered due to the encapsula-


tion of TiO2 with polymer and/or the declined UV absorption of TiO2
To date, titanium dioxide (TiO2)-based photocatalytic coatings are nanoparticles due to competition with other pigments in the coatings.
the most successful self-cleaning coatings utilized in practical applica- More critically, photocatalytic TiO2 nanoparticle can degrade most
tions. Many substrates, such as glass [1], lime [2], cement [3], marble polymeric binders [9], which seriously shorten the service life of the
[4], copper [5], polyethylene terephthalate [6], polycarbonate [7], and organic coatings. To solve the above problem, a strategy of an in-
polymethylmethacrylate [8], have been modified with TiO2-based dependent photocatalytic TiO2-based clearcoat cast onto traditional
coatings to achieve self-cleaning performance. Substrates such as metal organic coatings has been developed.
for architecture and traffic infrastructure, concrete for buildings and With the above strategy, two issues had to be addressed. One was
bridges, and polymers for outdoor objects, have all been decorated and/ the fabrication of the self-cleaning and photocatalytically active TiO2-
or protected with organic coatings, i.e. acrylic latex coatings, poly- based clearcoats, while the other was the need to avoid the photo-
urethane coatings, fluorocarbon polymer coatings, and polysiloxane catalytic decomposition of the organic coatings underneath by these
coatings. As one would expect, self-cleaning performance is a function clearcoats. The TiO2 photocatalyst employed in the clearcoat could be
of the topcoat rather than the substrate. incorporated by either a sol-gel process [10–12] or a blending method
Incorporation of TiO2 nanoparticles into organic topcoats may be a with photocatalytic TiO2 nanoparticles (for example, the commercial
way to acquire self-cleaning performance. However, the photocatalytic P25) [13–15]. Nevertheless, the sol-gel derived TiO2 coatings have the


Corresponding author.
E-mail address: zhoushuxue@fudan.edu.cn (S. Zhou).

http://dx.doi.org/10.1016/j.porgcoat.2017.08.005
Received 29 October 2016; Received in revised form 23 May 2017; Accepted 12 August 2017
0300-9440/ © 2017 Elsevier B.V. All rights reserved.
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

shortcoming of low photocatalytic activity if a high-temperature Table 1


treatment is not conducted. As for the blending method, the key issue The formulations of photocatalytic TiO2-based coatings.
involved the dispersion of TiO2 nanoparticles and the selection of an
Sample TM82/g G-TiO2/g TiO2 content/wt%
appropriate binder. We have previously reported the use of organic-
inorganic binders to produce durable photocatalytic TiO2-based self- T0 1.88 – 0
cleaning coatings [16,17]. However, until now the application of both T18 2.93 1.04 18
T33 2.20 2.09 33
sol-gel derived and blending-based TiO2 clearcoats directly onto or-
T45 1.46 3.13 45
ganic coatings and the control of the photocatalytic decomposition of T56 0.73 4.18 56
the organic coatings underneath, have seldom been reported.
Prevention of the degradation of the organic substrate (or coating)
due to the photocatalytic activity of the TiO2-based coating was his-
torically accomplished through the use of an inert interlayer [18]. For
examples, Cai [19] and coworker utilized corona treatment to enable a
uniform polysiloxane coating to form onto on the surface of a fluor-
ocarbon paint, followed by the deposition of a TiO2 film onto the
polysiloxane surface. Fateh and his coworker deposited a SiO2 barrier
layer and then TiO2 film, TiO2/ZnO film or TiO2/SiO2 film onto poly-
carbonate (PC) substrate [7,20,21]. In several Japanese patents
[22–24], a mediating protective layer formed from silicate, silicone or
alkylalkoxysilane was also reported in the fabrication of TiO2-based
self-cleaning coatings. In addition to the use of an inert interlayer, the
photocatalytic effect of a TiO2 layer has also be inhibited by the in-
corporation of a free-radical scavenging interlayer, i.e. polydopamine
(PDA) [25]. Although an interlayer can efficiently insulate the photo-
catalytic effect of the TiO2-based coating, it complicates the process of
preparing robust self-cleaning surfaces.
Herein, photocatalytic TiO2-based coatings were fabricated using
TiO2 nanoparticles and hybrid binders that were formed from tetraethyl Fig. 1. The FTIR spectra of TEOS, MTMS and dried TM82 coating.
orthosilicate (TEOS) and methyl trimethoxy silane (MTMS). It has been
found that these TiO2-based self-cleaning coatings not only did not
(2.50 mol) of H2O and 0.24 g of HCl were mixed and added drop wise
cause photocatalytic degradation of the architectural latex coating un-
to the flask over 1 h. The solution was then stirred at 70 °C for another
derneath, instead, they protected the coating from apparent UV-in-
2 h to obtain the hybrid binder denoted as TM82. The theoretical solids
duced damage, if the TiO2 content in the dried coatings was below 45
content of hybrid binder was 13.78% assuming that all of the orga-
wt%. Moreover, the TiO2-based coating had minimal impact on the
noalkoxysilanes in the formulation were completely hydrolyzed and
appearance of the original latex coating. Since no interlayer is required,
condensed.
these coatings simplify the establishment of a self-cleaning surface on a
painted object. The non-degradation of the organic coating could be
attributed to reduced UV transmission through the photo-catalyst 2.3. Preparation of photocatalytic TiO2-based coatings
containing clear coat as well as reduced water (and hydroxyl radical)
permeability to the interface between the latex coating and the clear The aqueous TiO2 dispersion (named as G-TiO2) was first prepared
coat. This principle facilitates the ability to design practical TiO2-based by deagglomeration of P25 in the presence of pre-hydrolyzed GPS
self-cleaning coatings for painted surfaces. dispersant that was synthesized according to previous work [26]. The
G-TiO2 dispersion had a solids content of 12.74% (GPS/TiO2=1.00/
2. Experimental 3.04 in mass ratio) and a dispersed size of 102 nm.
The TM82 hybrid binder and G-TiO2 were ultrasonically mixed in
2.1. Materials the amounts shown in Table 1 to form the photocatalytic TiO2-based
coatings. The coatings were cast onto glass slides by a drawdown rod
TEOS (chemical grade), aqueous HCl (37%), ethanol (analytical (wet film thickness: 20 μm) or applied onto dried latex coatings by a
grade) and xylene (analytical grade) were purchased from Sinopharm foam brush. The coatings were then cured under ambient conditions for
Chemical Reagent Co., Ltd. γ-glycidoxypropyltrimethoxysilane (GPS) 3 weeks. The sample name includes the percent of TiO2 for example T33
was purchased from Dow Corning. MTMS was purchased from Zhejiang contains 33% TiO2.
Quzhou Zhengbang Organosilicon Co., Ltd (China). Titania nano-
particles (P25) were provided by Degussa. Latex coatings were 30 2.4. Characterization
pigment volume concentration (PVC)-White and 30PVC-Blue paints,
bound with a butyl acrylate/methyl methacrylate/Methacrylic acid Fourier transform infrared (FTIR) analysis was carried out using KBr
copolymer, were provided by Dow Chemical Company. All reagents discs with a Nicolet Nexus 470 FTIR spectrometer (Thermo Nicolet)
were used as received. De-ionized water was used in all experiments. over the wavenumber range from 4000 to 500 cm−1 with a resolution
of 2 cm−1 and an accumulation of 32 scans.
2.2. Preparation of hybrid binder The morphology of the TiO2 dispersion was observed by TEM
(Hitachi H-600, Hitachi Corp) at an accelerating voltage of 100 kV. The
The hybrid binder was prepared according to the following process: samples were diluted with de-ionized water and directly dried on
84.9 g (1.84 mol) of ethanol, 86.0 g (0.41 mol) of TEOS and 14.0 g copper grids. The Z-average particle size and its distribution were de-
(0.10 mol) of MTMS were charged into a four-necked flask equipped termined using a Zetasizer Nano ZS90 instrument (Malvern, UK). X-ray
with a mechanical stirrer, a reflux condenser and a heating mantel. The diffraction analysis was accomplished using a Bruker X-ray dif-
molar ratio of TEOS/MTMS was about 8:2. The mixture was heated to fractometer (Siemens). The diffracted angle was recorded from 20 to
70 °C while stirring. Once the target temperature was acheived 45 g 90°.

16
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

Fig. 2. TEM images of (a) G-TiO2 and (b) G-TiO2


diluted by TM82. Inset is the particle size distribu-
tion obtained by dynamic laser scattering tests.

Fig. 3. SEM images of photocatalytic TiO2-based coatings: (a) T0, (b) T18, (c) T33, (d) T45, (e) T56.

The surface morphology of the coatings were observed using a field been cast onto glass slides (1.5 × 1.0 cm2) were pre-treated in an ac-
emission scanning electron microscope (FESEM, JSM-6701F, JEOL Co., celerated weathering tester for 24 h using the same procedure as that
Ltd, Japan), with an accelerating voltage of 10 kV. The specimens were used in accelerated weathering testing below. The treated samples were
sputter-coated with gold prior to FESEM imaging. The chemical com- then leaned against the wall of a 15 ml quartz vial containing 10 ml of
position of the coatings was measured by energy dispersive X-ray methyl blue solution with a concentration of 2×10−5 M which had
spectroscopy (EDX, QUANTAX 400, Bruker, Germany). been charged in advance. The vial was stored for 12 h in the dark to
UV–vis spectra were recorded using a UV–vis-NIR spectrometer (U- attain an equilibrated adsorption concentration. After that, it was illu-
4100, Hitachi Co., Japan), with air as a reference. The reflectivity of the minated by UVB light (intensity; 0.71 W/m−2) for 180 min. The ab-
coatings was determined using a reflectometer (C84-III, Shanghai sorbance of methyl blue solutions at 664 nm was periodically mon-
XianDai Environmental Engineering Technology Co., Ltd, China). The itored by the UV–vis-NIR spectrometer (U-4100, Hitachi Co., Japan).
reduced fraction of reflectivity (ΔR) was calculated with the formula, The contact angle was determined on an OCA15 contact angle
ΔR = (R0 − Rt)/R0 × 100%, where R0 and Rt correspond to the initial analyzer (Dataphysics, Germany) using a 3 μl droplet of deionized
reflection coefficient and the reflection coefficient of coatings at time t water. Average values from five parallel measurements were taken.
of outdoor exposure testing, respectively. The gloss at 60° of the latex The accelerated weathering testing (QUV) was conducted in a QUV
coating during accelerated weathering testing was measured using a tester (QUV/se, Q-PANEL Co., Ltd., USA). UV lamps with wavelength of
BYK-Gardner (AG-4601, BYK, Germany) gloss/hazer. Average values 310 nm were used and the accelerated weathering cycle was set as
from three independent measurements were adopted. The gloss reten- follows: UV irradiation with intensity of 0.71 W/m2 for 4 h at 60 °C, and
tion was calculated with the formula, Gt/G0 × 100%, where G0 and Gt condensation for 4 h at 50 °C. The gloss and color of the coatings were
correspond to the initial gloss and the gloss of the coating at time t in monitored during accelerated weathering testing. Color change (ΔE)
the accelerated weathering test. The color (L, a, b) of the coating was was then calculated according to Eq. (1):
measured using a spectrophotometer (CM-700d, Konica Minolta
Holdings, Inc., Japan). ΔE = (Lt − L0)2 + (at − a0)2 + (bt − b0)2 (1)
The photocatalytic properties of the coating were examined ac-
cording to reference [27]. Typically, the TiO2-based coatings which had where, (Lt, at, bt) represents the color of the coating after t hours of

17
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

Fig. 4. UV–vis transmittance spectra of photocatalytic TiO2-


based coatings. (film thickness: 600–900 nm).

Table 2 compatibility. Slight aggregation of TiO2 nanoparticles took place, as


Properties of latex coatings and photocatalytic clear coatings. shown in Fig. 2b. This indicates that G-TiO2 and TM82 are not com-
pletely compatible. This is reasonable because TM82 has relatively high
Sample Reflectivity ΔE Sample ΔE Gloss
non-polarity while the hydrolyzed GPS-encapsulated TiO2 nano-
30PVC-White 95.4 – 30PVC-Blue 47.3 particles are quite hydrophilic. Nevertheless, no agglomeration or
D-WT0 94.7 0.16 D-BT0 0.21 39.3 precipitates were found at any ratio of G-TiO2/TM82, thus enabling
D-WT18 93.2 0.64 D-BT18 0.17 37.2 adjustment of the TiO2 content in the photocatalytic coatings.
D-WT33 92.9 0.29 D-BT33 0.69 42.8
D-WT45 91.9 0.60 D-BT45 1.02 43.5
The photocatalytic TiO2-based coatings were cast onto quartz glass
D-WT56 91 0.83 D-BT56 1.83 54.5 slides in order to examine their morphology and optical properties.
Fig. 3 presents the SEM images of the surfaces of the coatings with
various TiO2 contents. The surface of the T0 sample (without TiO2
accelerated weathering testing and (L0, a0, b0) is the initial color of the nanoparticles) was very smooth (Fig. 3a). Protrusions were observed for
coating. the T18 coating (Fig. 3b) and these became more pronounced for the
T33 coating (Fig. 3c). These protrusions could be attributed to ag-
gregates of TiO2 nanoparticles. However, as the TiO2 content increased
3. Results and discussion
further, the surface roughness of the coatings declined. The T56 coating
had a homogenous surface (Fig. 3e), implying that the dispersion of
3.1. Preparation and properties of photocatalytic TiO2-based coatings
TiO2 nanoparticles was improved at this high ratio of TiO2 to TM82
binder.
Photocatalytic TiO2-based coatings were prepared by blending hy-
Fig. 4 shows the UV–vis transmittance spectra of the coatings with
brid binder TM82 with an aqueous TiO2 dispersion and drying at room
various TiO2 loadings. Pure TM82 coatings possessed almost the same
temperature. Fig. 1 presents the FTIR spectrum of the dried TM82. The
transmittance as glass. Introduction of TiO2 nanoparticles caused a
spectra of TEOS and MTMS are also included for comparison. The weak
slight decline in transparency. The transmittances of the coatings at
absorption band around 3000–2800 cm−1 indicated the nearly com-
550 nm decreased from 91.9% to 84.6% as the TiO2 loading increased
plete hydrolysis of TEOS and MTMS in the TM82 binder. The distinct
from 0 to 56%. Though slightly decreased, the transparencies of all
absorption peak at 1276 cm−1, due to SieCH3, clearly demonstrates the
TiO2-based coatings were still found to be acceptable. Fig. 4 also
participation of MTMS in the film-formation. Another strong absorption
showed that the photocatalytic TiO2-based coatings exhibit UV-
peak at around 1060 cm−1 corresponds to the antisymmetric stretching
shielding properties. The transmittances of the coatings of T0, T18, T33,
vibration of SieOeSi, suggesting the condensation of the precursors.
T45 and T56 at 350 nm were 90.4%, 67.2%, 66.8%, 59.7% and 40.1%,
Since the antisymmetric stretching vibration of the Si-O-Si from the
respectively. Basically, the higher the TiO2 loading is, the better the UV-
condensation product of pure TEOS is located at 1080 cm−1 [28], the
shielding performance of the coatings.
shift of this absorption band herein additionally indicates the co-con-
densation between TEOS and MTMS.
The aqueous TiO2 dispersion, G-TiO2, was first prepared from the 3.2. Application of photocatalytic TiO2-based coatings on latex coatings
P25 nanopowder by bead milling. As shown in Fig. 2a, the P25 nano-
powder was efficiently de-agglomerated. The particle size and PS dis- The photocatalytic TiO2-based coatings were easily applied over
tribution (PDI) of the G-TiO2 dispersion were measured to be 102 nm two types of latex coatings, 30PVC-White and 30PVC-Blue, using a
and 0.23, respectively. XRD analysis (Fig. 1s) indicates that the milled foam brush to form interlayer free photocatalytic clear topcoat/latex
TiO2 nanoparticles in G-TiO2 retained only about half of the crystal- base coat self-cleaning coatings. No defects were found for these two
linity of P25. A drop of G-TiO2 was added into TM82 to examine their layer coatings. Tables 2 and 3 summarize the properties of the double

18
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

Fig. 5. SEM images of cross section of 30PVC-Blue coating and double layer coatings: (a) 30PVC-Blue, (b) D-BT0, (c) D-BT18, (d) TD-BT33, (e) D-BT45 and (f) D-BT56.

layer coatings. Little color difference, ΔE, was seen for the 30 PVC white underneath. No voids were observed at the interface between the two
paint when the photocatalytic TiO2-based coating was cast on top over layers of coatings. The good adhesion may be partially due to the al-
the latex paint, despite the TiO2 content in the photocatalytic coating (it cohol in the TiO2-based coatings since alcohol functionality may serve
should be noted that ΔE≤1.5 is considered as no color change). The to improve wetting and to swell the polymer latex film.
reflectivity values of the two layer coatings were reduced slightly as the
TiO2 content in the photocatalytic coatings increased, likely due to the
3.3. Self-cleaning properties of photocatalytic clear layer coatings
lower reflectivity of TiO2 nanoparticles relative to the 300 nm TiO2
pigment in the latex coatings underneath. With respect to the 30PVC-
The photocatalytic activity of TiO2-based coatings was examined
Blue paint, the photocatalytic TiO2-based clear topcoat displayed only
using the methylene blue (MB) degradation test as shown in Fig. 6. MB
minimal influence on color and in this particular case the gloss. These
degraded with accelerated weathering for all TiO2-based coatings, but,
application results demonstrate that the photocatalytic TiO2-based
the degradation rate depended on the TiO2 content in the coating. After
coating does not impact the original appearance of the latex coating.
180 min of UV irradiation, only 40.8%, 33.4%, 30.3% and 28.5% of the
The cross-sections of the double layer coatings based on 30PVC-Blue
MB were remained for the T18, T33, T45, and T56 coatings, respec-
were observed via SEM, as shown in Fig. 5. double layer structures with
tively. In contrast, more than 90% of the MB remained in the solution
0.5–1.4 μm thick topcoats were clearly seen. The surface of the coatings
for bare glass and T0 coating. This result suggests that the TiO2-based
became smooth after casting a layer of clear topcoat. The TiO2-based
coatings have excellent photocatalytic activity.
clear coatings had similar morphologies over paint as were observed for
Different photo active clear layer coatings were further exposed
those cast on glass. Careful inspection of the cross-section indicated that
outdoors to determine their self-cleaning performance in a real world
the top clear coats were dense and well adhered to the latex coatings
environment. The reflection coefficient (△R) and water contact angle

19
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

Fig. 6. Degradation of methylene blue as function of UV irradiation time for the TiO2-
based coatings with different TiO2 contents.

Fig. 8. Optical images of the 30PVC-White latex coatings and clear layer coatings: (a)
before and (b) after outdoor exposure testing.

Fig. 7. The change of the reflection coefficient (△R) of the 30PVC-White latex coating
and the clear layer coatings during outdoor exposure testing.

(WCA) were monitored to help further our understanding. Fig. 7 shows


the △R of the clear layer coatings versus outdoor exposure time. The
△R of the latex coating increased considerably in the initial 100 days
of outdoor exposure, and then leveled off with additional exposure Fig. 9. Water contact angle (WCA) of the 30PVC-White latex coating and the clear layer
time. The profile of the △R illustrated that outdoor pollutants sig- coatings as a function of outdoor exposure time.
nificantly contaminate the surface of latex coatings quickly. The T0
clear layer coating displayed lower △R in comparison with the un-
performing best.
coated latex coating. This may partially be the result of the high
Fig. 9 shows the surface wettability of the coatings as a function of
hardness of the TM82 coating relative to the latex coating. Because the
outdoor exposure time. The initial WCAs are 71.1°, 61.2°, 73.1°, 66.2°,
TM82 binder contained a high content of inorganic content, its hard-
63.9° and 54.9° corresponding to the 30PVC-White latex and the clear
ness is obviously higher than that of standard latex coatings. Interest-
layer coatings, D-WT0, D-WT18, D-WT33, D-WT45 and D-WT56, re-
ingly, the △R was further improved after the introduction of TiO2
spectively. Upon outdoor exposure, the D-WT33, D-WT45 and D-WT56
nanoparticles. The higher the TiO2 content in the clear topcoat, the
coatings became super hydrophilic within one week. The D-WT18
lower the △R.
sample became super hydrophilic after one year. The 30PVC-White
The testing panels before and after one year’s outdoor exposure are
latex and the D-WTM82 samples maintained WCAs above 43.2° and
shown in Fig. 8. There is no obvious difference in appearance between
30.4°, respectively, during the entire duration of outdoor exposure
the latex coating and the clear top layer coatings before testing. After
testing. Sufficient loading of TiO2 nanoparticles (herein ≥33 wt.%) are
one year of exterior exposure, the latex coating (without the clear top
required to provide rapid conversion of the surface to super hydro-
coat) was very dirty and had turned gray while the clear layer coatings,
philicity during outdoor exposure.
especially with high TiO2 content, displayed a clean appearance. The
Therefore, even though all five of the clear layer coating samples
whiteness of the D-WT56 coating was similar to the original color.
show improved dirt pick up resistance, only T33, T45 and T56 can be
These results suggest that the clear layer coatings had good self-
designated as self-cleaning if super hydrophilicity is required.
cleaning performance, in general, with higher TiO2 content topcoats
Furthermore, the above results clearly indicate that the self-cleaning

20
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

photo-catalyzed decomposition of organic components in the clear


topcoats. Fig. 10a shows that the gloss of the latex coating continued to
significantly drop during the accelerated weathering test, such that it
had the lowest gloss of all samples after 1500 h of accelerated weath-
ering. The gloss retention of the latex coating dropped to 37% at 1500 h
of weathering testing; at least part of this gloss loss was due to surface
chalking of the latex coating. On the other hand, the gloss retention of
the clear top layer coatings did not show any obvious changes after the
initial drop during the first 500 h of accelerated weathering testing.
This result suggests that the morphological change of clear top layer
coatings happens primarily during the early stage of the weathering
test.
As shown in Fig. 10b, the color differences of the clear top layer
coatings (except D-BT0) were low (ΔE below 0.5) during the initial
500 h of the weathering test. However, over the next 500–1000 h of
weathering testing, the ΔE values rose to 1.4–3.5. The color retention
then remained in the range of 1.4–2.5 for up to 1800 h of total
weathering testing. Surprisingly, the ΔE of one sample, D-BT0, con-
tinued to increase throughout the entire accelerated weathering time
period. The ΔE values were high relative to standard coatings, pre-
sumably because of UV irradiation-induced yellowing of the resin in the
latex coating underneath. Because the TM82 film served as a barrier
layer, any yellow chemicals formed could not be washed away in the
condensation steps of weathering test. Nevertheless, the yellow che-
micals in the latex coatings without topcoats were easily washed off,
leading to small color difference during the entire weathering test. The
initial delay in color change for the photocatalytic clear layer coating
systems can be attributed to the UV-irradiation shielding activity of the
TiO2 in the clear coat. [30,31] Sample D-BT18 exhibited the worst
yellowing due to its insufficient UV shielding. The D-BT45 and D-BT56
clear layer coatings have better UV-shielding ability than the D-BT33
coating, but still possess high color changes after 1000 h of weathering
testing. This may be attributed to the photo-catalyzed degradation of
the latex coatings by the D-BT45 and D-BT56 clear layer coatings, as
evidenced from SEM observation below.
SEM was employed to analyze the cross section of the clear layer
Fig. 10. The (a) gloss retention and (b) color difference of the 30PVC-Blue latex coating
coating systems after 1755 h of accelerated weathering testing in an
and the clear layer coatings as a function of accelerated weathering time.
effort to further gain insight into the mechanism of protection and
degradation. Fig. 11 shows these micrographs. The latex coatings
coatings with higher TiO2 loading have better self-cleaning perfor- without topcoats displayed severely chalked surfaces due to the de-
mance. composition of the organic polymer. The chalking of the latex coating
was also clearly present for the D-BT0 and D-BT56 samples. The de-
3.4. Effect of photocatalytic topcoat on the durability of latex coatings composition of the latex coating in the D-BT0 sample is likely due to a
lack of UV-shielding property with this clear layer coat. EDX mapping
Although photocatalytic TiO2-based self-cleaning coatings have was typically conducted for the cross section of the D-TB56 coating as
been frequently reported, [6,8,19,29] few studies have measured the shown in Fig. 12. The C atoms in the latex coating near the photo-
degradation of organic substrates (or coatings) underneath a photo- catalytic top coat layer are low in concentration, directly suggesting
catalytic clear layer coating. Nevertheless, how the photocatalytic self- that the D-BT56 topcoat led to photocatalytic degradation of the
cleaning coatings impact the durability of organic substrates (or coat- basecoat. The basecoat under the D-BT45 clear coat showed only slight
ings) underneath is an extremely crucial factor for their practical ap- degradation while those under the D-BT18 and D-BT33 clear coats
plication. In this work, a 30PVC-Blue latex coating with gloss (60° gloss: showed no decomposition at all. We believe that this is because the
47.3) was used as the basecoat underneath the clear layer self-cleaning quantity of TiO2 nanoparticles is lower at the interface for the topcoat
coatings in order to fully understand the effect of top photocatalytic with lower TiO2 content. Furthermore, the TiO2 nanoparticles at the
TiO2-based coatings on the degradation of the latex coatings under- clear coat/latex interface have lower photocatalytic activity in com-
neath. The experiments were conducted in an accelerated weathering parison with those at the surface, due to the lack of moisture and the UV
chamber. The gloss and color were monitored during the accelerated adsorption which has occurred as the light passed through the bulk of
weathering test. the clear coat. Both moisture and UV are essential for high photo-
Fig. 10 shows the gloss retention and color difference of the coatings catalytic activity. [13] Another reason for reduced photocatalytic ac-
as a function of accelerated weathering time. In the initial 500 h of tivity may be the preferential occupation of hybrid binder at the in-
accelerated weathering testing, the clear layer coatings displayed a terface because of the better compatiblity of the hybrid binder with the
significant reduction in gloss while the gloss of the latex coating latex film. Evidence for this is the similar WCA values for the TM82 film
without a photocatalytic layer decreased only slightly. We speculate (61.2°) and the latex film (71.1°). In addition, the TM82 film was more
that the gloss reduction of the clear layer coatings may be the result of a homogeneous than that of the pure G-TiO2 film (Fig. 2s) when cast over
morphological change in the topcoat rather than degradation of the the latex film. Because the hybrid binder is present at the clear coat/
latex coating underneath. The morphological change during weathering latex interface, the latex film is insulated from the photocatalytic TiO2
may be caused by further condensation of the TM82 binder or by the nanoparticles in the clear layer. However, at high TiO2 content (i.e

21
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

Fig. 11. SEM images of cross section of (a) the 30PVC-Blue latex coating and clear layer coatings of (b) D-BT0, (c) D-BT18, (d) D-BT33, (e) D-BT45 and (f) D-BT56, after 1755 h of
accelerated weathering.

56%) the hybrid hinder cannot fully cover the clear coat/latex inter- commercial practicality of utilizing a photocatalytic self-cleaning
face, which leads to photocatalytic degradation of the basecoat. coating.
Fig. 11 also shows that the T18, T33 and T45 clear layer coats are a
bit loose and rough in nature after 1755 h of accelerated weathering.
We believe that this is caused by the photocatalytic degradation of 4. Conclusions
organic components (for example, GPS) in the coating. The thickness of
the topcoats did not appear to change with weathering (Fig. 4), in- Photocatalytic TiO2-based coatings were fabricated by simple
dicating the inherently good durability of the photocatalytic TiO2-based blending of a hybrid binder, synthesized from TEOS/MTMS, with TiO2
coatings. This is due to the siloxo-based binder which has the ability to nanoparticles which had been de-agglomerated using a GPS dispersant.
resist photocatalytic degradation. [16,32] When the coatings were directly cast onto latex coatings, they did not
Based on the gloss and color retention results, we can conclude that impact the original appearance of the basecoat because of the optical
if the TiO2 content is below 45%, the photocatalytic TiO2-based coat- transparency of the coatings, in spite of the TiO2 content. As revealed
ings will work to provide UV-shielding rather than as a photocatalytic from the methylene blue fade test and real outdoor exposure testing, the
degradation coating for the basecoat. This result is quite interesting TiO2-based clear layer coatings show photocatalytic activity with a
since it means that a durable photocatalytic self-cleaning coating for a dependence on TiO2 content. The coatings with TiO2 content above
painted surface can be fabricated, without an interlayer, by controlling 33% quickly became super hydrophilic and exhibited good self-cleaning
the nano TiO2 content of the clear layer coating to between 33 and performance. More interestingly, the TiO2-based clear layer coatings
45%. Eliminating the need for an interlayer greatly improves the with TiO2 content below 45% did not lead to photocatalytical de-
gradation of the latex coating underneath, but instead, protected the

22
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

Fig. 12. SEM images (a) and EDX mapping images (b) C atom, (c) Si atom and (d) Ti atom of the cross section of the D-BT56 coatings after 1755 h of accelerated weathering test.

latex coating from UV degradation. The non-degradation of the organic nanocomposite coating of TiO2 and polydimethylsiloxane, Colloids Surf. A 484
(2015) 471–477.
coating could be attributed to reduced UV transmission through the [6] L. Wu, Y. Yu, J. Zhi, Low cost and large-area fabrication of self-cleaning coating on
photocatalytic layer as well as reduced water permeation to the (latex polymeric surface based on electroless-plating-like solution deposition approach,
coating)/(self-cleaning coating) interface. The low quantity of TiO2 RSC Adv. 5 (2015) 10159–10164.
[7] R. Fateh, R. Dillert, D. Bahnemann, Self-cleaning properties, mechanical stability,
nanoparticles distributed at the clear coat/latex interface, due to the and adhesion strength of transparent photocatalytic TiO2-ZnO coatings on poly-
relatively low TiO2 content, and the preferential occupation of hybrid carbonate, ACS Appl. Mater. interfaces 6 (2014) 2270–2278.
binder at this interface, may further contribute to the low amount of [8] M. Langlet, Sol-Gel preparation of photocatalytic TiO2 films on polymer substrates,
J. Sol-Gel Sci. Technol. 25 (2002) 223–234.
photocatalytic degradation of the basecoat. Clear top layer coatings [9] N.S. Allen, M. Edge, A. Ortega, G. Sandoval, C.M. Liauw, J. Verran, J. Stratton,
with appropriate TiO2 content (33–45%) are crucial to the fabrication R.B. McIntyre, Degradation and stabilisation of polymers and coatings: nano versus
of photocatalytic self-cleaning coatings for painted surfaces. In addi- pigmentary titania particles, Polym. Degrad. Stab. 85 (2004) 927–946.
[10] Q. Mao, D. Zeng, K. Xu, C. Xie, Fabrication of porous TiO2?SiO2 multifunctional
tion, it is not necessary to have an interlayer as a barrier to photo-
anti-reflection coatings by sol–gel spin coating method, RSC Adv. 4 (2014)
catalytic activity to protect the latex coating below. We believe that the 58101–58107.
self-cleaning coatings described in this paper provide significant bene- [11] S.R. Meher, L. Balakrishnan, Sol–gel derived nanocrystalline TiO2 thin films: a
fits for application as photocatalytic self-cleaning clear layer topcoats promising candidate for self-cleaning smart window applications, Mater. Sci.
Semicond. Process. 26 (2014) 251–258.
over painted surfaces. [12] A. Sobczyk-Guzenda, B. Pietrzyk, H. Szymanowski, M. Gazicki-Lipman,
W. Jakubowski, Photocatalytic activity of thin TiO2 films deposited using sol–gel
Appendix A. Supplementary data and plasma enhanced chemical vapor deposition methods, Ceram. Int. 39 (2013)
2787–2794.
[13] Y. Zhao, Y. Liu, Q. Xu, M. Barahman, A.M. Lyons, Catalytic, self-cleaning surface
Supplementary data associated with this article can be found, in the with stable superhydrophobic properties: printed polydimethylsiloxane (PDMS)
online version, at http://dx.doi.org/10.1016/j.porgcoat.2017.08.005. arrays embedded with TiO2 nanoparticles, ACS Appl. Mater. Interfaces 7 (2015)
2632–2640.
[14] E.J. Park, H.S. Yoon, D.H. Kim, Y.H. Kim, Y.D. Kim, Preparation of self-cleaning
References surfaces with a dual functionality of superhydrophobicity and photocatalytic ac-
tivity, Appl. Surf. Sci. 319 (2014) 367–371.
[15] L. Pinho, M.J. Mosquera, Titania-Silica nanocomposite photocatalysts with appli-
[1] E. Allain, S. Besson, C. Durand, M. Moreau, T. Gacoin, J.P. Boilot, Transparent
cation in stone self-Cleaning, J. Phys. Chem. C 115 (2011) 22851–22862.
mesoporous nanocomposite films for self-Cleaning applications, Adv. Funct. Mater.
[16] L. Yang, S. Zhou, L. Wu, Preparation of waterborne self-cleaning nanocomposite
17 (2007) 549–554.
coatings based on TiO2/PMMA latex, Prog. Org. Coat. 85 (2015) 208–215.
[2] P. Munafò, E. Quagliarini, G.B. Goffredo, F. Bondioli, A. Licciulli, Durability of
[17] X. Ding, S. Zhou, L. Wu, G. Gu, J. Yang, Formation of supra-amphiphilic self-
nano-engineered TiO2 self-cleaning treatments on limestone, Constr. Build. Mater.
cleaning surface through sun-illumination of titania-based nanocomposite coatings,
65 (2014) 218–231.
Surf. Coat. Technol. 205 (2010) 2554–2561.
[3] E. Jimenez-Relinque, J.R. Rodriguez-Garcia, A. Castillo, M. Castellote,
[18] V.S. Smitha, K.B. Jaimy, P. Shajesh, J.K. Jeena, K.G. Warrier, UV curable hydro-
Characteristics and efficiency of photocatalytic cementitious materials: type of
phobic inorganic–organic hybrid coating on solar cell covers for photocatalytic self
binder, roughness and microstructure, Cem. Concr. Res. 71 (2015) 124–131.
cleaning application, J. Mater. Chem. A 1 (2013) 12641.
[4] C. Kapridaki, P. Maravelaki-Kalaitzaki, TiO2–SiO2–PDMS nano-composite hydro-
[19] R. Cai, G.M. Van, P.K. Aw, K. Itoh, Solar-driven self-cleaning coating for a painted
phobic coating with self-cleaning properties for marble protection, Prog. Org. Coat.
surface, C. R. Chimie 9 (2006) 829–835.
76 (2013) 400–410.
[20] R. Fateh, A.A. Ismail, R. Dillert, D.W. Bahnemann, Highly active crystalline meso-
[5] Y. Qing, C. Yang, Y. Sun, Y. Zheng, X. Wang, Y. Shang, L. Wang, C. Liu, Facile
porous TiO2 films coated onto polycarbonate substrates for self-Cleaning
fabrication of superhydrophobic surfaces with corrosion resistance by

23
F. Xu et al. Progress in Organic Coatings 113 (2017) 15–24

applications, J. Phys. Chem. C 115 (2011) 10405–10411. Pigm. Resin Technol. 39 (2010) 315–321.
[21] R. Fateh, R. Dillert, D. Bahnemann, Preparation and characterization of transparent [27] A. Mills, C. Hill, P.K.J. Robertson, Overview of the current ISO tests for photo-
hydrophilic photocatalytic TiO2/SiO2 thin films on polycarbonate, Langmuir 29 catalytic materials, Photochem. Photobiol. A 237 (2012) 7–23.
(2013) 3730–3739. [28] A. Fidalgo, L.M. Ilharco, The defect structure of sol-gel-derived silica/poly-
[22] M. Shimobukikoshi, K. Takeda, M. Sengoku, K. Takahashi, A. Shimai, Water-based tetrahydrofuran hybrid films by FTIR, J. Non-Cryst. Solids 283 (2001) 144–154.
Photocatalytic Hydrophilic Composition and Water-based Primer for Photocatalyst [29] Q.F. Xu, Y. Liu, F.J. Lin, B. Mondal, A.M. Lyons, Superhydrophobic TiO2-polymer
and Photocatalytic Hydrophilic Composite Material, (2001) (JP2001038219-A). nanocomposite surface with UV-induced reversible wettability and self-cleaning
[23] A. Shimai, K. Takeda, K. Takahashi, M. Sengoku, M. Shimobukikoshi, Aqueous properties, ACS Appl. Mater. Interfaces 5 (2013) 8915–8924.
Paint Composition Applicable to Organic Substrate and Coated Film Thereof Having [30] N. Abidi, L. Cabrales, E. Hequet, Functionalization of a cotton fabric surface with
Self Cleaning Action by Photocatalyst Contained, (2001) (JP2001031907-A). titania nanosols: applications for self-cleaning and UV-protection properties, ACS
[24] M. Hayakawa, J. Kameshima, K. Omoshiki, S. Kitazaki, T. Ikeda, M. Kanno, Appl. Mater. Interfaces 1 (2009) 2141–2146.
A. Shimai, Photocatalyst Coating Material for Exterior Material of Buildings, Has [31] K.D. Kim, H.O. Seo, C.W. Sim, M.G. Jeong, Y.D. Kim, D.C. Lim, Preparation of
Photocatalyst Layer Comprising Photocatalyst Particles, Inorganic Oxide Particles highly stable superhydrophobic TiO2 surfaces with completely suppressed photo-
and Binder on Base Material, (2012) (JP2012250133-A). catalytic activity, Prog. Org. Coat. 76 (2013) 596–600.
[25] K. Feng, L. Hou, B. Tang, P. Wu, A self-protected self-cleaning ultrafiltration [32] N.S. Allen, M. Edge, G. Sandoval, J. Verran, J. Stratton, J. Maltby, Photocatalytic
membrane by using polydopamine as a free-radical scavenger, J. Membr. Sci. 490 coatings for environmental applications, Photochem. Photobiol. 81 (2005)
(2015) 120–128. 279–290.
[26] G. Peter, Modified silica sols: titania dispersants and co-binders for silicate paints,

24

You might also like