Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Microchemical Journal 133 (2017) 1–12

Contents lists available at ScienceDirect

Microchemical Journal

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

Enhanced self-cleaning properties of N-doped TiO2 coating for


Cultural Heritage
L. Bergamonti a, G. Predieri b, Y. Paz c, L. Fornasini d, P.P. Lottici d, F. Bondioli a,⁎
a
Department of Engineering and Architecture, University of Parma, Parco Area delle Scienze 181/a, 43124 Parma, Italy
b
Department of Chemistry, Life Sciences and Environmental Susteinability, University of Parma, Parco Area delle Scienze 17/a, 43124 Parma, Italy
c
Department of Chemical Engineering, Technion-Israel Institute of Technology, 32000 Haifa, Israel
d
Department of Mathematical, Physical and Computer Sciences, University of Parma, Parco Area delle Scienze 7/a, 43124 Parma, Italy

a r t i c l e i n f o a b s t r a c t

Article history: Nanostructured TiO2 is the most widely studied photocatalytic material and is used in self-cleaning coatings in
Received 5 January 2017 buildings and for Cultural Heritage applications. To extend the photocatalytic activity of nanocrystalline titania
Received in revised form 3 March 2017 into the visible spectral range, doping of titania by non-metal elements as N, C, F and S is currently widely inves-
Accepted 3 March 2017
tigated. Here we report tests on carbonatic Travertine stones coated with two self-cleaning coatings based on N-
Available online 6 March 2017
doped TiO2 nanoparticles. The two coatings, both obtained by sol-gel synthesis, differ by their titanium precur-
Keywords:
sors and nitrogen sources. The obtained nanoparticles were characterized by X-ray diffraction, UV–VIS absorp-
Photocatalytic TiO2 tion, Raman and X-ray photoelectron spectroscopies. The harmlessness of the coatings is demonstrated by
Self-cleaning colorimetric analysis and by capillary absorption measurements. The self-cleaning properties of the treated
Travertine stones were evaluated by the measurement of water contact angle and of organic dyes degradation as a function
Sol-gel of UV–VIS light irradiation time. The results confirm that the synthesized nanoparticles were successfully doped
N-doped TiO2 with nitrogen, allowing a photocatalytic activity, even in the visible range, comparable or better than that shown
by P25 Evonik commercial product. A comparison between coating where the dopants are introduced by an acid-
ic process and coating doped using urea as the nitrogen source revealed higher activity for the former. The impor-
tance of the nature of the dye in the evaluation of the self cleaning properties is discussed.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction and expensive intervention and the use of detergents could cause
changes in the aesthetic aspect and the formation of by-products that
Stone-based historic monuments, buildings and sculptures, are fac- could be harmful for the objects as well as to the environment [16,17].
ing deterioration processes of constituent materials that have a power- To limit these invasive interventions, it is necessary to perform preven-
ful impact on their cultural, social and economic values [1]. Because of tive and long-lasting actions. The possibility to obtain building materials
air pollution [2], acid rain [3], soluble salts [4,5] and biodegradation [6, with self-cleaning surfaces is a current topic of great interest, especially
7], these cultural heritage objects undergo the risk of being irreparably for exterior surfaces exposed to environmental pollutants, atmospheric
lost. The degradation processes of natural and artificial stone artifacts precipitations and human damages [18–20]. Self-cleaning behavior of a
depend on the nature of the constituent materials and on the character- solid surface refers to the capability to remove pollutants under natural
istics of the environment. The actions of the chemical agents, and in par- environments. There are two ways to obtain this result: to design hy-
ticular of the atmospheric pollutants, are very dangerous on limestones drophilic surfaces or to achieve hydrophobic structures. In the first
and calcarenites (mainly composed of carbonates), frequently used as case the water contact angle (WCA) is close to zero, hence water
building materials [8–11]. These agents cause high levels of sulfate forms a uniform surface film, which prevents the soil clinging to sub-
and nitrate depositions, with development of black crusts that acceler- strate [21–23]. In the second case the surface is highly hydrophobic
ate the stone decay and produce an aesthetically unacceptable staining (WCA N 90°), so that the water drops roll over the surface while sweep-
of the stone surfaces [12–15]. For these reasons, it is necessary to re- ing dirt away [23,24]. Self-cleaning surfaces can be also obtained by syn-
move soiling material to slow down the deterioration processes and to ergistic action of superhydrophilic and photocatalytic materials to
improve the appearance and the aesthetic quality of the buildings. It transform pollutants into harmless species [25–27].
should be emphasized, however, that stone cleaning is an irreversible Titanium dioxide (TiO2) is the most attractive photocatalyst for air
and wastewater purification and is extensively used for oxidative and
⁎ Corresponding author. reductive removal of organic contaminants [28–31]. The use of titanium
E-mail address: federica.bondioli@unipr.it (F. Bondioli). dioxide (TiO2) for photocatalytic self-cleaning of building materials has

http://dx.doi.org/10.1016/j.microc.2017.03.003
0026-265X/© 2017 Elsevier B.V. All rights reserved.
2 L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12

started already in the early 1990s [32], nevertheless, its use for this reason, the discussion about the photocatalytic activity on a
protecting objects of cultural heritage is relatively new [33–38], and is model pollutant should be made with due consideration [74]. The effi-
characterized by an increasing number of studies concerning environ- ciency of a photocatalyst has to be referred to the specific dye
mental sustainability and potential human health risks [39,40]. employed: the photodegradation kinetics cannot be generalized to dif-
TiO2 is present in nature in three crystalline phases: rutile (thermo- ferent dyes and the test is useful only for a comparison of the
dynamically the most stable), anatase and brookite. Both rutile and an- photodegradation efficiency for that particular dye and between differ-
atase phases show photocatalytic reactivity but anatase seems to ent coatings in specific experimental conditions [76].
exhibit better photocatalytic performances than rutile [29]. Briefly, Static water contact angle (CA) tests were carried out in order to de-
photocatalysis is based on the generation of electron–hole pairs in termine also the effects on TiO2 treated surfaces of the hydrophilicity in-
TiO2 semiconductor upon irradiation with light. Electrons are created duced by the light irradiation, while capillary water absorption and
in the conduction band, leaving holes in the valence band: both charges colorimetric study were employed to verify the harmlessness of the N-
participate in redox reactions with adsorbed organic species and TiO2 coatings on the physical and aesthetical properties of the treated
adsorbed molecular oxygen and water [41] through formation of active stones [77].
radicals. The main obstacle for practical application of TiO2 has to do
with the large band-gap of the photocatalyst (~ 3.0 eV for rutile and 2. Experiment
~3.2 eV for anatase), meaning that the processes require UV light irradi-
ation, that corresponds to only 4–8% of the incident solar flux [42,43]. 2.1. Synthesis of nitrogen-doped TiO2 nanoparticles
Many methodologies have been employed to increase the light absorp-
tion towards the visible range (longer wavelengths) and to improve the Chemicals and materials used for the synthesis were titanium(IV)
photocatalytic performances of TiO2: sensitization with organic dyes oxysulfate (TiOSO4, Aldrich), titanium (IV) tetraisopropoxide (TTIP, Al-
[44], deposition of noble metal ions on the TiO2 surface [45–49], doping drich), urea (NH2CONH2, Aldrich), nitric acid (HNO3, 65%, Carlo Erba),
with transition metals [50–53], doping with non-metals (such as N, P, C, ammonium hydroxide (NH4OH, 26%, Carlo Erba). All reagents and
S) [54–57]. The term “doping” is used loosely for describing the addition Aeroxide P25 (Evonik) were used as received without any further
of foreign atoms into lattice, surface or interstitial sites. Compared to the purification.
other non-metal elements, nitrogen seems to be the most attractive be- Two different N-doped TiO2 sols (0.1 M), henceforth called N-TiA
cause of its comparable atomic size with that of oxygen, small ionization and N-TiU, were obtained.
energy and stability [57–61]. Nitrogen can be used as a dopant in the al- N-TiA was synthesized starting from titanium(IV) oxysulfate,
teration of the band-gap of TiO2 introducing filled N2p states above the TiOSO4 (1 × 10−2 mol), hydrolyzed in 100 ml distilled water at 40 °C.
TiO2 valence band, which decrease the apparent gap of TiO2 down to After cooling, NH4OH was slowly added dropwise to the solution, up
≈ 2.5 eV [54]. In addition, N-doping may form oxygen vacancies that to pH ≈ 7. The obtained amorphous precipitate was separated by cen-
may red-shift absorption [62] or enhance the photocatalytic efficiency trifugation and washed repeatedly with distilled water to remove the
by electron trapping, slowing down the recombination of sulfate ions (as confirmed by BaCl2 test). The powder was suspended
photogenerated electrons and holes [63–65]. in 100 ml of distilled water, followed by peptization with HNO3 to ob-
The incorporation of nitrogen into TiO2 structure has been reported tain pH ≈ 1.5.
via substitutional (replacement of oxygen) or interstitial (addition of ni- N-TiU sol was synthesized using tetraisopropoxide, TTIP, as a precur-
trogen into the TiO2 lattice) solid solution formation [66,67]. To incorpo- sor and urea as a complexing agent and nitrogen source: 0.1 mol of TTIP
rate nitrogen into TiO2 various chemical and physical techniques have were hydrolyzed in 10 ml of isopropanol. After vigorous stirring for
been developed, such as ammonolysis, sol-gel, hydrothermal and 15 min, 40 ml of distilled water were slowly added to hydrolyze the pre-
solvothermal reactions, sputtering, ion implantation and high energy cursor. 0.01 mol of urea were solubilized in 50 ml of water and then
milling with an organic compound [43,68,69]. Among these, the sol- added to Ti solution under vigorous stirring. The slurry was stirred at
gel method is the most used to obtain the desired doping level and room temperature for 1 h to favor the complete hydrolysis of the alkox-
nanoparticle size distribution, thanks to the easy control of the synthesis ide. After these procedures, both N-TiA and N-TiU sols were stirred vig-
conditions (type of solvents, pH, temperature, etc.). As a nitrogen orously at room temperature for 30 min and refluxed at 100 °C for 3 h to
source, aliphatic amines, nitrate and ammonium salts, nitrides, urea, induce the crystallization. The resulted N-TiA sol had a pale-yellow color
ammonia have been used [54,70]. and was very stable at a final pH of 1.6, while the N-TiU suspension,
The present work aimed to synthesize N-doped TiO2 powders by a characterized by a final pH of 8.3, was light yellow and quite unstable.
sol–gel method and to extend the photocatalytic activity of TiO2 into
the visible light region, thus obtaining a better photocatalytic efficiency. 2.2. Characterization of the titania samples
Nano-sized doped titania samples were obtained using different N
sources (ammonia/nitric acid and urea). The self-cleaning photocatalyt- To characterize the titania powders, small amounts of both sols were
ic properties of the N-doped TiO2 coatings were verified on Travertine centrifuged at 1000 rpm, repeatedly washed with distilled water and fi-
samples, one of the most famous Italian stones, extensively used since nally dried at 30 °C for 24 h. Room-temperature X-ray powder diffrac-
the Roman age. For this purpose, the degradation of two organic dyes, tion patterns were collected in the range 10°–60° by Thermo ARL
usually employed in staining/cleaning tests over building materials X'TRA X-ray diffractometer with Si–Li detector, using CuKα radiation
[71–73] was studied. It is known that under UV–VIS light, dyes may un- at 40 kV and 40 mA, at 0.2° scan rate (in 2θ). Measurements were per-
dergo photolysis and also can sensitize TiO2, injecting an electron into formed on samples spread on a conventional glass sample holder. The
its conduction band. The self-sensitization of the organic dyes in the vis- reflections of powdered crystalline silicon were used for 2θ calibration.
ible region undoubtedly complicates the interpretation of the X-ray diffraction patterns were compared with the standard anatase
photodegradation mechanisms [74,75], but, for sake of comparison, and brookite diffractograms (reference of anatase JCPDS card #21-
we followed the tests which are commonly employed for building ma- 1272 and brookite JCPDS card #29-1360 and rutile JCPDS card #76-
terials. The use of organic dyes is simplified with respect to other 1940).
“test” pollutants (e.g. stearic acid) and enables to verify by simple color- The average diameter of the crystallites was estimated by the
imetric methods the photodegradation of the stains on the stones up to Scherrer equation from the line broadening of the main diffraction
the restoration of the original aesthetical appearance. In addition, dyes peaks:
have complex photodegradation mechanistic pathways and an accurate
determination of photocatalytic efficiencies is a very difficult task. For D ¼ kλ=βcosθ ð1Þ
L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12 3

where D is the average crystal diameter (nm), k is the shape factor (usu- also by measuring the apparent density (2270 ± 54 kg/m3). The tests
ally taken = 0.89 for spherical particles), λ = 0.15406 nm is the Cu-Kα were performed on the parallel face to the plane of sedimentation.
X-ray wavelength, β is the full width at half-maximum (FWHM) inten- The N-TiA and N-TiU sols and the P25 0.1 M dispersion were applied
sity of the diffraction peak and θ is the Bragg diffraction angle [78]. on stones by brush. The amount of titania per unit surface (mg/cm2)
Non-polarized Raman scattering measurements were carried out in was calculated from the dried weight of the stone samples before and
a nearly backscattered geometry using a Jobin Yvon LabRAM micro- after the application of the coating: 0.19 ± 0.01 mg/cm2 for N-TiA,
spectrometer (300 mm focal length spectrograph) equipped with an in- 0.26 ± 0.01 mg/cm2 for N-TiU and 0.22 ± 0.1 mg/cm2 for P25.
tegrated Olympus BX40 microscope. Spectra were recorded at room To evaluate the changes of the stone surface appearance due to the
temperature. The powder samples were irradiated with 632.8 nm line TiO2 coatings according to UNI EN 15886:2010 norms [82], colorimetric
(nominal 15 mW He–Ne laser excitation). The spectral resolution was analysis was performed by a Techkon Spectrodens colorimeter using
about 1.5–2 cm−1. The Rayleigh radiation was blocked by an edge filter D65 illuminant (daylight standard) and 10° viewing angle geometry.
and the backscattered Raman light was dispersed by a 1800 grooves/ The results on nine regions of a few mm2 of each specimen before and
mm holographic grating on a Peltier cooled CCD, consisting of an array after treatment were averaged on five samples. Color differences were
of 1024/256 pixels. The entrance-slit width was fixed at 100 μm. The measured in the CIELab and CIELCH spaces, that better represent
laser power on the sample was adjusted at b2 mW by means of a series human sensibility [83]. In the CIELab space L∗ values range from 0 to
of density filters to avoid any damage to the samples or uncontrolled +100 (black-white) and represents the lightness, while a* (opponent
thermal effects. Spectra were collected using both ×100 and long work- red-green) and b* (opponent yellow-blue) are the chromatic coordi-
ing distance × 50 microscope objectives. Typical exposures were nates. In the CIELCH space, the attributes of chroma, C* =
10–60 s, repeated three to five times. The system was regularly calibrat- (a*2 + b*2)1/2, and hue, h* = tan−1(b*/a*), represent the color purity
ed using the 520.6 cm−1 Raman band of silicon or by means of reference or saturation and the hue of the saturated color, respectively.
emission lines of Ar or Cd light sources. The data analysis was performed The colorimetric changes ΔE*, ΔC* and ΔH*, due to TiO2 applications
by LabSpec built-in software. with respect to the uncoated stone, were evaluated by:
Specific surface areas were determined by the classical Brunauer–
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Emmett–Teller (BET) procedure with a Micromeritics Pulse Chemisorb 2
ΔE ¼ ΔL2 þ Δa2 þ Δb
2705 analyzer using nitrogen at 77 K (single point method). Prior to pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð2Þ
measuring, the samples (150 mg) were outgassed under helium flow ΔH ¼ ΔE2 −ΔC 2 −ΔL2
at 473 K for 2 h. ΔC  ¼ C r− C s
UV–VIS diffuse reflectance spectroscopy (DRS) measurements on
TiO2 powdered samples were carried out by a Varian 2390 spectropho- where C∗r and C∗s are the chroma of uncoated and coated samples,
tometer equipped with an integrating sphere. The spectra were record- respectively.
ed in reflectance mode at room temperature, with BaSO4 as reference, Changes in C* and h* are more sensitive to changes of a* or b* coor-
scanning in the 250–800 nm wavelength range. Absorption coefficients dinates, depending on the original color of the material. Travertine is a
and energy gaps were calculated based on Kubelka–Munk treatment yellowish stone characterized by b* values higher than a*: in our case,
[79] and Tauc's plots [80]. L* = 81.6, a* = 1.9, b* = 9.6. In the evaluation of the treatment's effects
TiO2 samples were investigated by XPS using a Thermo Scientific it should be taken into account that C* is strongly influenced by the co-
Sigma Probe, hemispherical energy analyzer and a monochromatic Al ordinate b*, whereas h* is very sensitive to changes in a*.
Kα source (hν = 1486.6 eV). The powders were attached onto a carbon The microstructure of the stone surfaces before and after the coat-
tape. The signal from the C1s peak at 285 eV was taken as internal con- ings was examined by a Scanning Electron Microscope Jeol JSM 6400
trol, to account for charging effects. The samples were first scanned in equipped with an Oxford Instruments Link Analytical Si(Li) Energy Dis-
fast mode, where the signal is collected in a wide angle (from 24° to persive System detector (SEM-EDS). The data analysis was performed
84° relative to the sample's normal). Then specific elements were by INCA built-in software. The microanalyses were carried out with re-
scanned more precisely in a slower mode. Here, the signal was collected spect to calcium as reference.
from a narrower window (from 24° to 38° degrees relative to the nor- For the water capillary absorption tests, five samples for each coating
mal). This corresponds to a bulk-emphasizing measurement. A step were employed, following UNI EN 15801:2010 recommendation [84].
size of 1 eV was used for survey scanning, whereas for the acquisition After coating, the samples were dried at 60 °C in a ventilated oven,
of specific peaks the step size was 0.05 eV. The lateral size of the beam until reaching mass difference b 0.1% between two measurements, at
was 200 μm by 400 μm, which is large enough to average over local var- 24 h intervals. The specimens were then placed on blotting paper satu-
iations among particles. rated with demineralized water, with the TiO2 coated side facing down-
wards. The water level was maintained constant throughout the test.
2.3. Stone characterization and sols application Specimens were weighed at regular time intervals and at subsequent

Travertine stone is a sedimentary stone formed by deposition of


CaCO3 generated from thermal sources and directly from the upper
mantle [13]. The stone is characterized by parallel layer, typical struc-
ture of the sedimentary rock. The samples used in this study (Fig. 1)
present layers with massive and laminated structures. The massive
type is characterized by small pores irregularly distributed while the
laminated type presents large and irregular pores parallel to bedding
[81]. Travertine is a highly heterogeneous stone, and a single stone
block can contain several different facies. The porosity is mainly inter-
crystalline and intergranular and, according to the technical specifica-
tion, in our samples is 8.3% on average. Macropores up to centimeter
size give to the stones the classical vacuolar structure.
Due to the heterogeneity of Travertine stone, the samples were cho-
sen so as to have, for each coating, the same number of samples with Fig. 1. Typical Travertine structure in the direction perpendicular to the sedimentation
massive and layered structure. The selection of the samples was done plane: the difference between a) massive and b) laminated structures is evident.
4 L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12

24 h intervals until t = tf, when the difference between two consecutive The photocatalytic oxidation reactions were performed under the
weighing was b1% of the initial weight of the specimen. The amount of Osram ultra-Vitalux lamp. The test was made also under visible light ir-
absorbed water Qi at time ti was calculated by Q i ¼ Mi −M radiation, adding a low-energy 400 nm cutoff filter. A further test was
A , M0 being the
0

mass of the dry specimen, Mi the mass at time ti and A the area in contact made with a LED at 405 nm. In Fig. 2 are reported the emission spectra
with the wet paper. of the light sources and the absorption spectra of MeO and RhB.
The calculated values of Qi were reported (in kg/m2) as a function of Three coverings of the dyes' solution (about 1.5 ml, 1 mM) were ap-
the square root of time t1/2
i . The absorption coefficient (AC) and the rel-
plied by a brush on the surface of the stone samples. After each covering
ative capillary index (RCI) were also reported. AC (in kg/(m2s1/2) was step, the stones were kept to dry in ventilated oven at 50 °C for 60 min.
evaluated as the slope of the curve Qi at low times (b30 min), where The coating degradation activity on the dyes was evaluated by color-
the amount of the absorbed water vs. t1/2 should be approximately lin- imetric measurements on nine control points for each stained sample
ear according to the most used model for the capillarity absorption [85]. surface (three samples for each type) using the Techkon Spectrodens
The RCI parameter is defined [38,86] as the ratio between the area colorimeter. Assuming as reference the starting value of the color pa-
under the Qi curve for the treated stone, (Qi)tr, and for the untreated rameters measured on the stained stone, the color variation was moni-
one, (Qi)unt: tored over the irradiation time. A normalized chroma change, ΔCn, at
different exposure times, t, was obtained by the chromatic coordinates
a*(t) and b*(t) as follow:
t
∫ 0f ðQ i Þtr dt
RCI ¼ ð3Þ vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t
∫ 0f ðQ i Þunt dt u 
uða ðt Þ−a ð0ÞÞ2 þ ðb ðt Þ−b ð0ÞÞ2
ΔC n ¼ t  2    2 ð4Þ
aC −a ð0Þ þ bC −b ð0Þ
representing the water absorption capability of a given coated sample
versus the control.
where a∗C and b∗C are measured on clean stones before the staining with
dyes [38]. This function should reach 1 after the complete dye
2.4. Photoinduced properties
degradation.

The wetting degree of the stone surface and its light induced changes
3. Results and discussion
were assessed by the values of water contact angle (WCA). Measure-
ments of the static contact angles were performed by OCA 20 apparatus
3.1. TiO2 powders characterization
(DataPhysic Instrument GmbH) using the sessile drop method, in accor-
dance to UNI EN 15802:2010 [87] standard rules. A water drop with vol-
The specific surface area of the powder, as determined by the BET
ume of 3 μl was placed on the surface by a needle: CA measurements
isotherm, was 260 m2/g and 230 m2/g for N-TiA and N-TiU, respectively,
were made on the drop profile 15 s after its deposition, every 30 min,
more than four times larger than that of commercial TiO2 powder
up to 2 h. The reported values are an average of 12 + 12 measurements
(Evonik, Aeroxide P25) [88], generally considered one of the most active
taken on two stone specimens for each treatment (N-TiA, N-TiU, P25)
TiO2-based products for dye photodegradation. These results of specific
and for the untreated reference.
surface areas confirm those obtained previously for undoped titania
Contact angles measurements were repeated also on samples irradi-
powders prepared by the authors using a similar sol-gel method [37].
ated by a UV–VIS Osram ultra-Vitalux (300 W, 280–800 nm spectral
The XRD patterns of P25, N-TiA and N-TiU are shown in Fig. 3. The
range emission). The distance between the lamp and the stone samples
anatase, rutile and brookite percentage in the powders was calculated
was kept at 14 cm to achieve an irradiance power in the UV range of
by the integrated diffraction peak intensities, according to the method
about 3.5 mW/cm2.
proposed by Zhang and Banfield [89]. As expected, P25 is characterized
Methyl orange (MeO) and rhodamine B (RhB), extensively
by the presence of anatase, as main phase, and rutile, as secondary
employed in various applications, were used to evaluate the photocata-
phase (Fig. 3a). The rutile weight fraction, WR, was determined by:
lytic efficacy of the N-doped TiO2 coatings on stained stone samples.
MeO is an anionic azo-dye frequently used in titration as pH indicator:
at pH b 3.1 is red, until pH 4.4 is orange (pKa = 3.8) and at pH N 4.4 is AR
WR ¼ ð5Þ
clearly yellow. When dissolved in water exhibits a broad peak at ðkA AA þ AR Þ
about 465 nm, typical of a substituted azo-benzene.
RhB is a cationic red dye of the xanthene class, often used as a tracer where AA and AR are the integrated intensities of the (101) anatase main
dye and in biology as a staining fluorescent dye. The pH value of a water peak and of the (110) rutile peak, respectively, and kA = 0.884 is a coef-
solution of RhB is 4.3. Rhodamine B exists, in the 1–13 pH range, in two ficient determined by calibration. In the insert of Fig. 3a the
principal forms, i.e. cationic (RhB+) or zwitterionic (RhB±), with its deconvolution employed to determine the rutile WR fraction is report-
absorption maximum close to 551–553 nm. ed: in good agreement with the commonly reported values WR =

Fig. 2. (a) Emission spectra of the blue LED and of the Osram ultra-Vitalux lamp, with and without UV filter; (b) absorption spectra of methyl-orange and rhodamine B dyes.
L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12 5

0.21 ± 0.02 [88]. A rough estimation of the nanocrystals size by the


Scherrer formula gives 17 ± 4 nm for anatase and 25 ± 4 nm for rutile.
In the case of N-TiA and N-TiU, the observed diffraction peaks can be
perfectly indexed by diffraction planes of tetragonal anatase (marked
with A). Besides, a small diffraction peak of the brookite phase (marked
with B), is observed at 2θ = 30.81°. The main diffraction peak of anatase
(101) at 2θ = 25.28° overlaps with the (120) and (111) reflections of
brookite at 2θ = 25.34° and 2θ = 25.69°, respectively. The weight frac-
tion of brookite phase, WB, was calculated using the integrated intensi-
ties AA of the (101) anatase peak and AB of the (121) brookite peak, by:

kB AB
WB ¼ ð6Þ
ðkA AA þ kB AB Þ

where kB is equal to 2.721 [89]. In the inserts of Fig. 3b and c the


deconvolution employed to determine the anatase/brookite phase con- Fig. 4. Raman spectra of TiO2 nanopowders: the wavenumbers of the main anatase peaks
are shown. A, B and R indicate features of anatase, brookite and rutile phases, respectively.
tent were reported. In the fitting procedure with lorentzian or pseudo-

Fig. 5. UV–VIS absorption spectra of N-TiA and N-TiU doped TiO2 and of the reference P25.
Fig. 3. XRD pattern of P25 (a), N-TiA (b), N-TiU (c): (A) anatase. (R) rutile, (B) brookite. In (a) Log(1/R) versus λ (R is the reflectance); (b) Kubelka-Munk F(R) function vs. photon
the inserts: (a) fit of the main anatase-rutile and (b, c) anatase-brookite peaks. energy, hv; (c) Tauc's plot (hνF(R))1/2 vs. photon energy, hv.
6 L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12

Voigt functions, it was assumed that (120), (111), and (121) brookite obtained from the output of the spectrometer. The band gap energy
peaks have the same peak broadening, while their intensities, according was estimated by the Kubelka–Munk (KM) function: the absorption co-
to the JCPDS card no. 29-1360, were fixed by the ratios 100/80/90, re- efficient, K, and the scattering coefficient, S, are related at each wave-
spectively. The percentage fraction of anatase and brookite were esti- length to the KM function by the following [92]:
mated 82% and 18% for N-TiA and 92% and 8% for N-TiU, respectively.
 
Finally, the line broadening of the main diffraction peak (101) of anatase
K 1−R2
was used to determine the crystallite size of the particles: 4.5 ± 1.0 nm F ðRÞ ¼ KM ¼ ¼ ð7Þ
and 8.5 ± 1.2 nm in N-TiA and N-TiU, respectively, significantly lower S 2R
than that of P25.
Raman spectroscopy is very sensitive to the crystallinity and micro- Under the assumption that S does not depend on the wavelength,
structure of TiO2 and is largely applied to discriminate the local order the KM function can be taken as proportional to the absorption coeffi-
characteristics of TiO2. In Fig. 4 the Raman spectra of both N-TiA and cient. Eg can be roughly calculated by the Tauc's plot, (hνK)n ∝
N-TiU nanopowders were compared with that of P25. The spectra of (hνF(R))n ∝ (hν–Eg), assuming an indirect band gap (i.e. n = 1/2), as
both N-doped powders show five high-intensity Raman peaks, ob- known for anatase, by fitting the linear part of the curve and extrapolat-
served at 151–153 cm−1, 204 cm− 1, 398 cm− 1, 512 cm−1, and ing the hν value for F(R) = 0.
635 cm−1, corresponding to the characteristic Eg, Eg, B1g, A1g(B1g), and In Fig. 5(a–c) the plots obtained by UV–VIS Diffuse Reflectance Spec-
Eg Raman modes of the TiO2 anatase structure, respectively. Moreover, troscopy (DRS) for the N-doped TiO2 powders and the reference P-25
some features from the brookite phase, also evidenced by XRD data, are reported. Both N-TiA and N-TiU showed a light absorption increase
are observed at 246–318/324–366 cm−1. The Raman spectrum of P25, towards the visible range relative to P25. The band gaps obtained from
a mixture of anatase (~80%) and rutile (~20%) according to the XRD re- Tauc's plots were 3.25, 3.10 and 3.02 eV for P25, N-TiU, N-TiA, respec-
sults, shows weak features of rutile, hardly visible on the intense ana- tively. The absorption enhancement in the visible-light range is in
tase spectrum. It is well-known that the most intense mode at good agreement with the yellowish color of both samples.
143 cm−1 in crystalline anatase shows increasing broadening and a XPS measurements performed on both N-TiA and N-TiU powders
shift towards higher wavenumbers, as the particle size decreases [90]. showed some differences between the two powders. Both samples re-
The broadening (FWHM: 14 cm−1 in P25, 26 cm− 1 in N-TiU and vealed the Ti2p doublet (2p3/2, 2p1/2), at binding energies of 458 eV
33 cm−1 in N-TiA) and the shift (from 143 cm−1 to 150 cm− 1 up to and 464 eV, which are typical for Ti4+ species in TiO2 (Fig. 6a, c) [93].
155 cm−1) observed in these N-doped samples, which are only margin- Nevertheless, the N-TiA powder revealed also the presence of a second
ally due to the brookite contribution, are compatible with anatase Ti2p doublet (Fig. 6c) of binding energy lower by 1.5 eV than that of
nanocrystals of 5–10 nm size [91] in good agreement with the XRD re- Ti4+ species in pristine TiO2. This suggests that unlike in the N-TiU pow-
sults, where N-TiA shows larger linewidths than N-TiU. der, where all the titanium atoms are in the same environment as in
The UV–VIS absorption of a semiconductor is described by the ener- TiO2, part of the titanium in the N-TiA samples experience a different
gy gap value, Eg, corresponding to a strong increase in the absorbance, environment, leading to weaker binding of the core electrons to the ti-
due to the excitation of valence-band electrons into the conduction tanium nucleus. The Ti peak in TiN was reported as located around
band. Since the UV–VIS spectra of the powdered catalysts were record- 455 eV [94]: such a shift might indicate the presence of substitutional
ed in diffuse-reflection mode, the diffuse reflectance, R, was directly nitrogen, provided that there is evidence also from the position of the

Fig. 6. XPS peaks of Ti2p and N1s in N-TiU (A and B) and in N-TiA powders (C and D).
L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12 7

nitrogen peak. Despite the inherent low intensity of the N1s peak, a presence of nitrate. An interesting XPS finding was the ratio between
clear difference between N-TiU and N-TiA is observed. In the N-TiU sam- the atomic concentration of oxygen to that of titanium, which was con-
ples, nitrogen peak is comprised of a single peak, centered at binding en- siderably higher than 2 in all samples. This may be originated from the
ergy of 400 eV (Fig. 6b). In contrast, the nitrogen peak in the N-TiA presence of adsorbed oxygen-containing molecules at the surface. How-
samples is broader and comprised of two peaks of equal contribution, ever, in that case, one could expect that the oxygen/titanium ratio in the
one at 400 eV and a second at 401 eV (Fig. 6d). Former work has claimed surface mode of operation would be higher than in the bulk mode, in
that the binding energy of electrons in interstitial N in titania should be contrast to our measurements that showed almost no difference be-
around 400–401 eV and that, if nitrogen is in both interstitial and substi- tween the two modes of operation.
tutional position, the two peaks should be well resolved [95]. It should In conclusion, it is possible to say that the changes in the optical ab-
be noted, however, that the N1s peak in TiN is 397 eV [94] and that sorption of the N-doped photocatalysts can be ascribed to the introduc-
this value is slightly shifted to lower energies (by 0.5–1 eV), in the pres- tion of localized N2p doping states near the valence band edge of TiO2.
ence of oxygen, i.e. in TiOxNy [96]. Hence, the location of the nitrogen While doping may introduce nitrogen atoms in both interstitial and
peak seems to suggest the nitrogen is in interstitial positions in both substitutional positions, our XPS results, corroborated with the absorp-
types of samples, with larger variety in the N-TiA samples. The exact na- tion results, suggest that the nitrogen dopants in the N-TiA are likely to
ture of the interstitial peak of nitrogen in titanium oxide is not clear and reside in positions that affect the immediate environment of titanium
was hypothesized as N2, NO− 3 or N\\H [97]. Indeed, another N1s peak atoms, whereas nitrogen dopants in the N-TiU samples are mostly inter-
was found at 408 eV in some of the N-TiA samples, indicating the stitial and do not affect the local environment of the titanium atoms. At

Fig. 7. SEM images and EDS spectra on: (a) untreated, (b) N-TiA coated, (c) N-TiU coated and (d) P25 coated Travertine.
8 L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12

Fig. 8. Partial and total color differences of the Travertine coated with TiO2 with respect to the raw stone, without coatings.

any case, it should be noted that whether substitutional positioning of (RCI) and the capillary absorption coefficient AC30, evaluated as the
the nitrogen is a prerequisite for visible light activity [98] or not [99] is slope of the linear fit to the Qi curve vs. t1/2 in the first 30 min.
still a matter of controversy. Capillary test measurements are affected by experimental uncer-
tainty due to the non homogeneity of the stone samples of the order
3.2. Stone characterization of 2–4%: nevertheless, the overall capillary rise seems to be slightly im-
proved by both titania treatments with respect to untreated samples, as
The microstructure of the travertine samples, before and after the ti- shown in Fig. 9 for long-time test as well as evidenced by the RCI values.
tania treatments, was determined by SEM-EDS measurements (Fig. 7). The AC30 coefficient shows a significant improvement of capillary rise
The stones mainly consist of calcium carbonate (CaCO3) in calcite only for N-TiU coating compared to the untreated surface, while for N-
form. As clearly visible in Fig. 7a, the untreated stone shows both TiA and P25 coatings the decrease of capillary absorption was not clearly
micritic (crystals b 2 μm) and crystalline texture. The hexagonal crystals observed, suggesting similar behavior between treated and untreated
of calcite, grown inside the pores, are distinguishable. A homogeneous surfaces in short time.
and crack-free coating is observed for the N-TiA-coated surface
(Fig. 7b). In contrast, the N-TiU-coated surface (Fig. 7c) and P25- 3.3. Photoinduced activity of TiO2 coatings
coated surface (Fig. 7d) are less uniform and reveal micro crystalline ag-
gregates of TiO2. The wettability, measured by the water contact angle, is a very im-
The EDS spectra were recorded from the corresponding micrograph portant property depending on the chemical composition and the struc-
areas, showing the presence of titania in all the treated samples. SEM- ture of the solid surfaces. The static contact angle values measured
EDS measurements on the cross sections, not reported here, reveal a before light irradiation show (Fig. 10) significant differences between
limited (0.3 b μm) penetration depth for both treatments. treated and untreated stones. In particular, the N-TiA coating induces
To verify the preservation of the original appearance, color measure-
ments were performed on the surface of the stone samples before and
after the application of the treatments. In Fig. 8 the total (ΔE*) and par-
tial (ΔL*, ΔC* and ΔH*) color differences and the colorimetric coordinate
differences (Δa*, Δb*) are reported. The total color difference ΔE* after
the application of the N-TiA and N-TiU treatments is lesser than 3, a
value which is well below the acceptable limit in the field of cultural
heritage [71]. The commercial P25 coating gives a higher color differ-
ence (~3) and a decrease of color saturation with respect to the untreat-
ed stone, probably due to the larger particles' size and the higher
agglomeration degree as observed by SEM analysis. Analyzing the be-
havior of the color coordinates, it can be concluded that the color change
is due to the decrease of b* parameter, indicating a slight yellowing of
the stone.
The effects of the coating on the behavior of the stones in the pres-
ence of water were evaluated by comparative measurements of water
absorption by capillarity (Qi) on untreated and titania coated samples,
according to standard tests [84]. The Qi values (kg/m2) as a function of Fig. 9. Capillarity water absorption tests carried on Travertine samples treated with P25,
square root of time (s1/2) are shown in Fig. 9 prior to and after coating, N-TiA and N-TiU. Open circles: untreated stone; full circles: treated stone. Error bars are
for each type of samples. Table 1 reports the relative capillary index shown on the legend symbols.
L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12 9

Table 1 The high photo-activity of the N-TiA coating is maintained also


Capillary absorption coefficient (AC30) and relative capillary index (RCI). blocking the UV portion by a filter, while both P25 and N-TiU do not in-
AC30 AC30 RCI (±0.02) duce appreciable changes in MeO discoloration with respect to the un-
Before coating After coating treated stone, as evident in Fig. 11c.
P25 0.016 ± 0.001 0.015 ± 0.002 0.93 The ΔCn curves for blue LED irradiation are reported in Fig. 11e: N-
N-TiA 0.014 ± 0.002 0.012 ± 0.001 0.96 TiA is again the most effective catalyst, photodegrading N 50% of the
N-TiU 0.009 ± 0.001 0.006 ± 0.001 0.84 MeO dye in 90 min, in spite of the slower kinetics under irradiation at
lower power with respect to the UV–VIS irradiation.
The effects of the photocatalysis on RhB are shown in Fig. 11b, d, f.
an increase of the water contact angle approaching hydrophobicity (al- Under UV–VIS Osram irradiation (Fig. 11b) the degradation of the dye
most 100°). This behavior is probably due to the homogeneous, smooth is similar for all the treated surfaces and higher than for untreated
film formed by this treatment, as previously observed by SEM analysis. ones: 80% of discoloration is reached in 3 h.
The well-known UV photoinduced hydrophilicity of TiO2 coatings is Using UV filter (Fig. 11d) and LED (Fig. 11f) the photoactivity effi-
a fundamental side effect of photocatalytic self-cleaning surfaces [34]. ciency of all TiO2 catalysts is moderate. The photo-cleaning induced on
During light irradiation, the contact angle of N-TiA and P25 treated stained surfaces by catalysts is faster during the first hour of irradiation
stones decreased while the CA of N-TiU and untreated stone (UT) with respect to UT stones. With the UV filtered lamp the 40% of the RhB
remained roughly the same. In particular, for N-TiA coating, after 2 h cleaning is obtained only after about 3 h, whereas the same degradation
of irradiation the contact angle fell drastically to b30°, well below the efficiency with the blue LED is obtained in about 7 h. On the untreated
P25 values. The change during irradiation in the shape of water droplets stone the self-fading of the dye doesn't exceed the 15–20%.
on a Travertine sample, coated with N-TiA sol, is shown in Fig. 10b. Unlike MeO behavior, where the N-TiA coating, under visible light ir-
To compare the self-cleaning ability of the different titania coatings, radiation, is clearly active, for RhB the different kinds of TiO2 seem to be
we have chosen to evaluate the photocatalytic activity on stained stones not particularly efficient.
by means of colorimetric measurements of the photodegradation of The two dyes investigated in the photocatalytic tests behave in a
both methyl orange (MeO) and rhodamine B (RhB) dyes under expo- quite different manner during the discoloration. The difference of the
sure to UV–VIS and VIS illumination as explained in the experimental reactivities of the dyes could be due to their degree of adsorption on
section. TiO2, which depends both on the surface pH and on the ionization
The normalized change in the Chroma ΔCn is reported as a function state of the organic molecules [101]. It is known that at pH value less
of the irradiation time (in minutes): Fig. 11 shows the discoloration of than the Point of Zero Charge (PZC) of TiO2 (≈6.5) the surface is posi-
the dyes on untreated (UT) and treated Travertine samples, under dif- tively charged [78–100]. Therefore, one should expect that the methyl
ferent light sources, enabling a comparison between the N-doped TiO2 orange anionic dye is better adsorbed on acid TiO2 (N-TiA) than on neu-
coatings N-TiA and N-TiU with commercial P25. On the untreated tral (P25) or basic (N-TiU) TiO2 [78]. This could explain the higher activ-
stone, slow photodegradation of the dyes takes place even in the ab- ity of the acid N-TiA compared to that of P25 and N-TiU in the photo-
sence of catalyst under the different irradiation sources: 20–30% for degradation of MeO.
MeO and 15–30% for RhB, after about 3 h, with higher discoloration Rhodamine B is a cationic/zwitterionic dye and the adsorption most-
under UV–VIS Osram lamp. This may be due to the natural self-fading ly occurs at the basic/neutral TiO2 surfaces (negatively charged), like N-
of the dyes when exposed to light irradiation [100]. TiU and P25 that indeed exhibit good photocatalytic performances, in
Under UV–VIS irradiation the efficiency of MeO decomposition sig- particular under UV–VIS irradiation. Nevertheless, also N-TiA has
nificantly improved in the presence of the photocatalysts (Fig. 11a). proved to be very effective in RhB discoloration: this could be due to
The fastest and most effective decomposition was obtained by N-TiA the presence of the zwitterionic form of RhB, whose deprotonated car-
treatment that, at the end of the experiment, decomposed almost 90% boxyl groups should make possible dye adsorption on N-TiA particles
of the dye. A 40% and 60% MeO discoloration was obtained by N-TiU [102].
and P25 catalysts, respectively, in both cases higher than the self- Generally, in both dyes discoloration, the N-TiA revealed higher ac-
fading of the dye on the untreated stone. tivity: this result is correlated with the shifted absorption spectrum

Fig. 10. (a) Evolution of the contact angle as a function of the irradiation time (b) Photographs of the water droplets over a Travertine sample coated by N-TiA, taken at 0, 60 and 120 min
(top to bottom).
10 L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12

and with the XPS results that indicated N doping in positions that direct- titanium), whereas doping using urea as a nitrogen source usually yields
ly affect the vicinity of the titanium atoms. The N atoms into TiO2 lattice nitrogen in interstitial positions.
can form intermediate energy levels, leading to the narrow band gap of The coatings do not alter the aesthetic properties of Travertine stone
TiO2. and slightly improve the resistance to the capillary water absorption. Both
In any case, the specimens treated with N-TiA or N-TiU show always treatments exhibit good photocatalytic activity under UV–VIS irradiation
higher dye degradation than that obtained by P25, probably due to the in the discoloration of organic dyes such as methyl orange and rhodamine
absorption at higher wavelengths. B used as contaminants of the Travertine, even in the presence of potential
dye photosensitization effects. In particular, the photodegradation of MeO
by N-TiA is more efficient with respect to P25 and N-TiU whereas for RhB
4. Conclusions no appreciable differences are found between the different treatments.
A more effective adsorption of the anionic MeO dye on the N-TiA
Sol-gel N-doped TiO2-based photocatalytic coatings, prepared with nanoparticles could explain this behavior.
different Ti-precursors and N sources (N-TiA and N-TiU), were synthe- By UV filtering or by using a blue-LED irradiation, the effectiveness of
sized, characterized and tested to evaluate their efficiency as self- N-TiA coating for MeO degradation is maintained, whereas, in the RhB
cleaning treatments for Travertine stone. The obtained TiO2 crystals test, all treatments show a reduced activity.
(anatase as main crystalline phase) have nanometric size. The presence The increase of the absorption in the visible range for the N-doped
of nitrogen, that creates an enhancement in the visible-light absorption, TiO2 seems to be effective only for N-TiA with MeO. The results for rho-
was confirmed by XPS analysis, which indicates that doping by an acidic damine B cannot definitely interpreted: this fact confirms the intrinsic
process (N-TiA samples) may yield nitrogen in substitutional positions difficulties found when using dyes as model compounds for the deter-
(or at least in positions that affect the immediate environment of mination of the photocatalytic activity [75].

Fig. 11. Photocatalytic activity vs. time of P25, N-TiA and N-TiU coatings on methyl orange and rhodamine B under UV–VIS (a and b), UV-filter (c and d) and Blue LED irradiation (e and
f) measured by chromatic changes ΔCn of the stained stone.
L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12 11

The UV photoinduced hydrophilicity, measured by the static water of nanocrystalline TiO2 with application as photoactive coating on stones, Environ.
Sci. Pollut. Res. 21 (2014) 13264–13277, http://dx.doi.org/10.1007/s11356-013-
contact angle as a function of irradiation time, is evident only in the 2136-5.
case of N-TiA coating that enhanced the wettability of the stone surfaces [21] C. Kapridaki, P. Maravelaki-Kalaitzaki, TiO2-SiO2-PDMS nano-composite hydropho-
bic coating with self-cleaning properties for marble protection, Prog. Org. Coat. 76
improving the self-cleaning properties. (2013) 400–410, http://dx.doi.org/10.1016/j.porgcoat.2012.10.006.
The results, due to the synergic effect of the anatase phase structure, [22] P.N. Manoudis, A. Tsakalof, I. Karapanagiotis, I. Zuburtikudis, C. Panayiotou, Fabrica-
small crystallite size, high specific surface area and N-doping, highlight tion of super-hydrophobic surfaces for enhanced stone protection, Surf. Coat.
Technol. 203 (2009) 1322–1328, http://dx.doi.org/10.1016/j.surfcoat.2008.10.041.
the compatibility of the proposed treatments on the studied carbonatic [23] J.P. Youngblood, N.R. Sottos, Bioinspired materials for self-cleaning and self-
stone. Of the two tested treatments, N-TiA coating displays good self- healing, MRS Bull. 33 (2008) 732–741, http://dx.doi.org/10.1557/mrs2008.158.
[24] C. Garlisi, Self-cleaning coatings activated by solar and visible radiation, J. Adv.
cleaning properties (hydrophilicity and photocatalytic activity), open-
Chem. Eng. 05 (2015) 5–7, http://dx.doi.org/10.4172/2090-4568.1000e103.
ing interesting possibilities for building applications and in the cultural [25] R. Wang, K. Hashimoto, A. Fujishima, M. Chikuni, E. Kojima, A. Kitamura, M.
heritage field. Shimohigoshi, T. Watanabe, Light-induced amphiphilic surfaces, Nature 388
(1997) 431–432, http://dx.doi.org/10.1038/41233.
[26] A. Fujishima, X. Zhang, D. Tryk, TiO2 photocatalysis and related surface phenome-
na, Surf. Sci. Rep. 63 (2008) 515–582, http://dx.doi.org/10.1016/j.surfrep.2008.10.
Acknowledgments 001.
[27] S. Banerjee, D.D. Dionysiou, S.C. Pillai, Self-cleaning applications of TiO2 by photo-
induced hydrophilicity and photocatalysis, Appl. Catal. B Environ. 176-177
Support from MAECI (project “NANO4HER, Nanotechnology at the
(2015) 396–428, http://dx.doi.org/10.1016/j.apcatb.2015.03.058.
service of cultural heritage preservation”, Italy-Israel Scientific and Tech- [28] M.B. Fisher, D.A. Keane, P. Fernández-Ibáñez, J. Colreavy, S.J. Hinder, K.G. McGuigan,
nological Cooperation, Ministry of Science) is gratefully acknowledged. S.C. Pillai, Nitrogen and copper doped solar light active TiO2 photocatalysts for
water decontamination, Appl. Catal. B Environ. 130–131 (2013) 8–13, http://dx.
doi.org/10.1016/j.apcatb.2012.10.013.
References [29] A. Fujishima, T.N. Rao, D.A. Tryk, Titanium dioxide photocatalysis, J Photochem
Photobiol C: Photochem Rev 1 (2000) 1–21, http://dx.doi.org/10.1016/S1389-
[1] R. Douglas-Jones, J.J. Hughes, S. Jones, T. Yarrow, Science, value and material decay 5567(00)00002-2.
in the conservation of historic environments, J. Cult. Herit. (2016)http://dx.doi.org/ [30] S.O. Obare, G.J. Meyer, Nanostructured materials for environmental remediation of
10.1016/j.culher.2016.03.007. organic contaminants in water, J. Environ. Sci. Heal. Part A- Toxic/Hazardous Subst.
[2] C.R. Esterkin, A.C. Negro, O.M. Alfano, A.E. Cassano, Air pollution remediation in a Environ. Eng. 39 (2004) 2549–2582, http://dx.doi.org/10.1081/LESA-200027010.
fixed bed photocatalytic reactor coated with TiO2, AIChE J. 51 (2005) 2298–2310, [31] Y. Paz, Application of TiO2 photocatalysis for air treatment: patents' overview,
http://dx.doi.org/10.1002/aic.10472. Appl. Catal. B Environ. 99 (2010) 448–460, http://dx.doi.org/10.1016/j.apcatb.
[3] S. Eyssautier-Chuine, B. Marin, C. Thomachot-Schneider, G. Fronteau, A. Schneider, 2010.05.011.
S. Gibeaux, P. Vazquez, Simulation of acid rain weathering effect on natural and ar- [32] J. Chen, C. Poon, Photocatalytic construction and building materials: from funda-
tificial carbonate stones, Environ. Earth Sci. 75 (2016) 748, http://dx.doi.org/10. mentals to applications, Build. Environ. 44 (2009) 1899–1906, http://dx.doi.org/
1007/s12665-016-5555-z. 10.1016/j.buildenv.2009.01.002.
[4] N. Tsui, R.J. Flatt, G.W. Scherer, Crystallization damage by sodium sulfate, J. Cult. [33] E. Quagliarini, F. Bondioli, G.B. Goffredo, A. Licciulli, P. Munafò, Smart surfaces for
Herit. 4 (2003) 109–115, http://dx.doi.org/10.1016/S1296-2074(03)00022-0. architectural heritage: preliminary results about the application of TiO2-based
[5] C. Rodriguez-Navarro, E. Doehne, E. Sebastian, How does sodium sulfate crystal- coatings on travertine, J. Cult. Herit. 13 (2012) 204–209, http://dx.doi.org/10.
lize? Implications for the decay and testing of building materials, Cem. Concr. 1016/j.culher.2011.10.002.
Res. 30 (2000) 1527–1534, http://dx.doi.org/10.1016/S0008-8846(00)00381-1. [34] E. Quagliarini, F. Bondioli, G.B. Goffredo, A. Licciulli, P. Munafò, Self-cleaning mate-
[6] P. Tiano, Biodeterioration of stone monuments a worldwide issue, Open Conf. Proc. rials on Architectural Heritage: compatibility of photo-induced hydrophilicity of
J. 7 (2016) 29–38, http://dx.doi.org/10.2174/2210289201607020029. TiO2 coatings on stone surfaces, J. Cult. Herit. 14 (2013) 1–7, http://dx.doi.org/10.
[7] S. Eyssautier-Chuine, N. Vaillant-Gaveau, M. Gommeaux, C. Thomachot-Schneider, 1016/j.culher.2012.02.006.
J. Pleck, G. Fronteau, Efficacy of different chemical mixtures against green algal [35] M.F. La Russa, S.A. Ruffolo, N. Rovella, C.M. Belfiore, A.M. Palermo, M.T. Guzzi, G.M.
growth on limestone: a case study with Chlorella vulgaris, Int. Biodeterior. Crisci, Multifunctional TiO2 coatings for Cultural Heritage, Prog. Org. Coat. 74
Biodegrad. 103 (2015) 59–68, http://dx.doi.org/10.1016/j.ibiod.2015.02.021. (2012) 186–191, http://dx.doi.org/10.1016/j.porgcoat.2011.12.008.
[8] S. Eyssautier-Chuine, M. Gommeaux, C. Moreau, C. Thomachot-Schneider, G. [36] D. Colangiuli, A. Calia, N. Bianco, Novel multifunctional coatings with photocatalytic
Fronteau, J. Pleck, B. Kartheuser, Assessment of new protective treatments for po- and hydrophobic properties for the preservation of the stone building heritage,
rous limestone combining water-repellency and anti-colonization properties, Q. J. Constr. Build. Mater. 93 (2015) 189–196, http://dx.doi.org/10.1016/j.conbuildmat.
Eng. Geol. Hydrogeol. 47 (2014) 177–187, http://dx.doi.org/10.1144/qjegh2013- 2015.05.100.
026. [37] L. Bergamonti, F. Bondioli, I. Alfieri, A. Lorenzi, M. Mattarozzi, G. Predieri, P.P. Lottici,
[9] E. Franzoni, E. Sassoni, Correlation between microstructural characteristics and Photocatalytic self-cleaning TiO2 coatings on carbonatic stones, Appl. Phys. A
weight loss of natural stones exposed to simulated acid rain, Sci. Total Environ. Mater. Sci. Process. 122 (2016) 124, http://dx.doi.org/10.1007/s00339-015-9560-y.
412-413 (2011) 278–285, http://dx.doi.org/10.1016/j.scitotenv.2011.09.080. [38] L. Bergamonti, I. Alfieri, A. Lorenzi, A. Montenero, G. Predieri, G. Barone, P.
[10] E. Sassoni, E. Franzoni, Influence of porosity on artificial deterioration of marble Mazzoleni, S. Pasquale, P.P. Lottici, Nanocrystalline TiO2 by sol–gel: characterisa-
and limestone by heating, Appl. Phys. A Mater. Sci. Process. 115 (2014) 809–816, tion and photocatalytic activity on Modica and Comiso stones, Appl. Surf. Sci.
http://dx.doi.org/10.1007/s00339-013-7863-4. 282 (2013) 165–173, http://dx.doi.org/10.1016/j.apsusc.2013.05.095.
[11] O. Rudic, D. Rajnovic, D. Cjepa, S. Vucetic, J. Ranogajec, Investigation of the durabil- [39] F. Grande, P. Tucci, Titanium dioxide nanoparticles: a risk for human health?
ity of porous mineral substrates with newly designed TiO2-LDH coating, Ceram. Int. Mini-Rev. Med. Chem. 16 (2016) 762–769, http://dx.doi.org/10.2174/
41 (2015) 9779–9792, http://dx.doi.org/10.1016/j.ceramint.2015.04.050. 1389557516666160321114341.
[12] A. Bonazza, C. Sabbioni, N. Ghedini, Quantitative data on carbon fractions in inter- [40] A.M. Ferrari, M. Pini, P. Neri, F. Bondioli, Nano-TiO2 Coatings for Limestone: which
pretation of black crusts and soiling on European built heritage, Atmos. Environ. 39 Sustainability for cultural heritage? Coatings 5 (2015) 232–245, http://dx.doi.org/
(2005) 2607–2618, http://dx.doi.org/10.1016/j.atmosenv.2005.01.040. 10.3390/coatings5030232.
[13] Á. Török, Black crusts on travertine: factors controlling development and stability, [41] J.-M. Herrmann, Photocatalysis fundamentals revisited to avoid several misconcep-
Environ. Geol. 56 (2008) 583–594, http://dx.doi.org/10.1007/s00254-008-1297-x. tions, Appl. Catal. B Environ. 99 (2010) 461–468, http://dx.doi.org/10.1016/j.
[14] C. Grossi, R. Esbert, F. Díaz-Pache, F. Alonso, Soiling of building stones in urban en- apcatb.2010.05.012.
vironments, Build. Environ. 38 (2003) 147–159, http://dx.doi.org/10.1016/S0360- [42] C. Burda, Y. Lou, X. Chen, A.C.S. Samia, J. Stout, J.L. Gole, Enhanced nitrogen doping
1323(02)00017-3. in TiO2 nanoparticles, Nano Lett. 3 (2003) 1049–1051, http://dx.doi.org/10.1021/
[15] G. Fronteau, C. Schneider-Thomachot, E. Chopin, V. Barbin, D. Mouze, A. Pascal, nl034332o.
Black-crust growth and interaction with underlying limestone microfacies, Geol. [43] L. Rojas-Blanco, M.D. Urzúa, R. Ramírez-Bon, F.J. Espinoza Beltrán, Photocatalytic
Soc. Lond. Spec. Publ. 333 (2010) 25–34, http://dx.doi.org/10.1144/SP333.3. thin films containing TiO2:N nanopowders obtained by the layer-by-layer self-
[16] C. Vazquez-Calvo, M. Alvarez de Buergo, R. Fort, M.J. Varas-Muriel, The measure- assembling method, Appl. Surf. Sci. 258 (2012) 2103–2106, http://dx.doi.org/10.
ment of surface roughness to determine the suitability of different methods for 1016/j.apsusc.2011.03.150.
stone cleaning, J. Geophys. Eng. 9 (2012) S108–S117, http://dx.doi.org/10.1088/ [44] C. Guarisco, G. Palmisano, G. Calogero, R. Ciriminna, G. Di Marco, V. Loddo, M.
1742-2132/9/4/S108. Pagliaro, F. Parrino, Visible-light driven oxidation of gaseous aliphatic alcohols to
[17] F. Valentini, A. Diamanti, M. Carbone, E.M. Bauer, G. Palleschi, New cleaning strat- the corresponding carbonyls via TiO2 sensitized by a perylene derivative, Environ.
egies based on carbon nanomaterials applied to the deteriorated marble surfaces: a Sci. Pollut. Res. 21 (2014) 11135–11141, http://dx.doi.org/10.1007/s11356-014-
comparative study with enzyme based treatments, Appl. Surf. Sci. 258 (2012) 2546-z.
5965–5980, http://dx.doi.org/10.1016/j.apsusc.2012.01.076. [45] C. Liu, D. Yang, Y. Jiao, Y. Tian, Y. Wang, Z. Jiang, Biomimetic synthesis of TiO2 –SiO2
[18] P. Munafò, G.B. Goffredo, E. Quagliarini, TiO2-based nanocoatings for preserving ar- –Ag Nanocomposites with enhanced visible-light photocatalytic activity, ACS Appl.
chitectural stone surfaces: an overview, Constr. Build. Mater. 84 (2015) 201–218, Mater. Interfaces 5 (2013) 3824–3832, http://dx.doi.org/10.1021/am4004733.
http://dx.doi.org/10.1016/j.conbuildmat.2015.02.083. [46] Y. Guo, S. Yang, X. Zhou, C. Lin, Y. Wang, W. Zhang, Enhanced photocatalytic activ-
[19] E. Franzoni, A. Fregni, R. Gabrielli, G. Graziani, E. Sassoni, Compatibility of photocat- ity for degradation of methyl Orange over silica-titania, J. Nanomater. 2011 (2011)
alytic TiO2-based finishing for renders in architectural restoration: a preliminary 1–9, http://dx.doi.org/10.1155/2011/296953.
study, Build. Environ. 80 (2014) 125–135, http://dx.doi.org/10.1016/j.buildenv. [47] K. Loganathan, P. Bommusamy, P. Muthaiahpillai, M. Velayutham, The syntheses,
2014.05.027. characterizations, and photocatalytic activities of silver, platinum, and gold
[20] L. Bergamonti, I. Alfieri, M. Franzò, A. Lorenzi, A. Montenero, G. Predieri, M. doped TiO2 nanoparticles, Environ. Eng. Res. 16 (2011) 81–90, http://dx.doi.org/
Raganato, A. Calia, L. Lazzarini, D. Bersani, P.P. Lottici, Synthesis and characterization 10.4491/eer.2011.16.2.81.
12 L. Bergamonti et al. / Microchemical Journal 133 (2017) 1–12

[48] S.A. Ansari, M.M. Khan, M.O. Ansari, M.H. Cho, Silver nanoparticles and defect- [74] M. Rochkind, S. Pasternak, Y. Paz, Using dyes for evaluating photocatalytic proper-
induced visible light photocatalytic and photoelectrochemical performance of ties: a critical review, Molecules 20 (2015) 88–110, http://dx.doi.org/10.3390/
Ag@m-TiO2 nanocomposite, Sol. Energy Mater. Sol. Cells 141 (2015) 162–170, molecules20010088.
http://dx.doi.org/10.1016/j.solmat.2015.05.029. [75] B. Ohtani, Preparing articles on photocatalysis—beyond the illusions, misconceptions,
[49] K. Mogyorósi, Á. Kmetykó, N. Czirbus, G. Veréb, P. Sipos, A. Dombi, Comparison of and speculation, Chem. Lett. 37 (2008) 3, http://dx.doi.org/10.1246/cl.2008.216.
the substrate dependent performance of Pt-, Au-and Ag-doped TiO2 photocatalysts [76] N. Shaham-Waldmann, Y. Paz, Away from TiO2: a critical minireview on the develop-
in H2-production and in decomposition of various organics, React. Kinet. Catal. ing of new photocatalysts for degradation of contaminants in water, Mater. Sci.
Lett. 98 (2009) 215–225, http://dx.doi.org/10.1007/s11144-009-0052-y. Semicond. Process. 42 (2016) 72–80, http://dx.doi.org/10.1016/j.mssp.2015.06.068.
[50] J. Medina-Valtierra, C. Frausto-Reyes, J. Ramírez-Ortiz, G. Camarillo-Martínez, Self- [77] P. Munafò, E. Quagliarini, G.B. Goffredo, F. Bondioli, A. Licciulli, Durability of nano-
cleaning test of doped TiO2 -coated glass plates under solar exposure, Ind. Eng. engineered TiO2 self-cleaning treatments on limestone, Constr. Build. Mater. 65
Chem. Res. 48 (2009) 598–606, http://dx.doi.org/10.1021/ie8008555. (2014) 218–231, http://dx.doi.org/10.1016/j.conbuildmat.2014.04.112.
[51] T.A. Segne, S.R. Tirukkovalluri, S. Challapalli, Studies on characterization and photo- [78] L. Bergamonti, I. Alfieri, A. Lorenzi, A. Montenero, G. Predieri, R. Di Maggio, F.
catalytic activities of visible light sensitive TiO2 nanocatalysts Co-doped with mag- Girardi, L. Lazzarini, P.P. Lottici, Characterization and photocatalytic activity of
nesium and copper, Int. Res. J. Pure Appl. Chem. 1 (2011) 84–103, http://dx.doi. TiO2 by sol–gel in acid and basic environments, J. Sol-Gel Sci. Technol. 73 (2015)
org/10.9734/IRJPAC/2011/453. 91–102, http://dx.doi.org/10.1007/s10971-014-3498-y.
[52] T. Umebayashi, T. Yamaki, H. Itoh, K. Asa, Analysis of electronic structures of 3d [79] R. López, R. Gómez, Band-gap energy estimation from diffuse reflectance measure-
transition metal-doped TiO2 based on band calculations, J. Phys. Chem. Solids 63 ments on sol–gel and commercial TiO2: a comparative study, J. Sol-Gel Sci. Technol.
(2002) 1909–1920, http://dx.doi.org/10.1016/S0022-3697(02)00177-4. 61 (2012) 1–7, http://dx.doi.org/10.1007/s10971-011-2582-9.
[53] B. Thirupathi, P.G. Smirniotis, Co-doping a metal (Cr, Fe, Co, Ni, Cu, Zn, Ce, and Zr) [80] K. Madhusudan Reddy, S.V. Manorama, A. Ramachandra Reddy, Bandgap studies
on Mn/TiO2 catalyst and its effect on the selective reduction of NO with NH3 at on anatase titanium dioxide nanoparticles, Mater. Chem. Phys. 78 (2003)
low-temperatures, Appl. Catal. B-Environ. 110 (2011) 195–206, http://dx.doi.org/ 239–245, http://dx.doi.org/10.1016/S0254-0584(02)00343-7.
10.1016/j.apcatb.2011.09.001. [81] M.Á. García-del-Cura, D. Benavente, J. Martínez-Martínez, N. Cueto, Sedimentary
[54] S.A. Ansari, M.M. Khan, M.O. Ansari, M.H. Cho, Nitrogen-doped titanium dioxide structures and physical properties of travertine and carbonate tufa building
(N-doped TiO2) for visible light photocatalysis, New J. Chem. 40 (2016) stone, Constr. Build. Mater. 28 (2012) 456–467, http://dx.doi.org/10.1016/j.
3000–3009, http://dx.doi.org/10.1039/C5NJ03478G. conbuildmat.2011.08.042.
[55] X. Chen, C. Burda, The electronic origin of the visible-light absorption properties of [82] UNI EN 15886, Conservation of cultural property—test methods— colour measure-
C-, N- and S-doped TiO2 nanomaterials, J. Am. Chem. Soc. 130 (2008) 5018–5019, ment of surfaces, 2010, UNI Ente Nazionale Italiano di Unificazione, 2010.
http://dx.doi.org/10.1021/ja711023z. [83] C.M. Grossi, P. Brimblecombe, R.M. Esbert, F.J. Alonso, Color changes in architectur-
[56] J. Schneider, M. Matsuoka, M. Takeuchi, J. Zhang, Y. Horiuchi, M. Anpo, D.W. al limestones from pollution and cleaning, Color. Res. Appl. 32 (2007) 320–331,
Bahnemann, Understanding TiO2 photocatalysis: mechanisms and materials, http://dx.doi.org/10.1002/col.20322.
Chem. Rev. 114 (2014) 9919–9986, http://dx.doi.org/10.1021/cr5001892. [84] EN 15801, Conservation of Cultural Property - Test Methods - Determination of
[57] V. Vaiano, O. Sacco, D. Sannino, P. Ciambelli, Nanostructured N-doped TiO2 coated Water Absorption by Capillarity, 2009.
on glass spheres for the photocatalytic removal of organic dyes under UV or visible [85] E.W. Washburn, The dynamics of capillary flow, Phys. Rev. 17 (1921) 273–283,
light irradiation, Appl. Catal. B Environ. 170-171 (2015) 153–161, http://dx.doi. http://dx.doi.org/10.1103/PhysRev.17.273.
org/10.1016/j.apcatb.2015.01.039. [86] R. Peruzzi, T. Poli, L. Toniolo, The experimental test for the evaluation of protective
[58] V. Vaiano, O. Sacco, G. Iervolino, D. Sannino, P. Ciambelli, R. Liguori, E. Bezzeccheri, treatments: a critical survey of the “capillary absorption index”, J. Cult. Herit. 4
A. Rubino, Enhanced visible light photocatalytic activity by up-conversion phos- (2003) 251–254, http://dx.doi.org/10.1016/S1296-2074(03)00050-5.
phors modified N-doped TiO2, Appl. Catal. B Environ. 176-177 (2015) 594–600, [87] UNI EN 15802:2010, Conservation of Cultural Property—Test
http://dx.doi.org/10.1016/j.apcatb.2015.04.049. Methods—Determination of Static Contact Angle, UNI Ente Nazionale Italiano di
[59] O. Sacco, M. Stoller, V. Vaiano, P. Ciambelli, A. Chianese, D. Sannino, Photocatalytic Unificazione, Milano, 2010.
degradation of organic dyes under visible light on N-doped. photocatalysts, Int. J. [88] J. Yu, H. Yu, B. Cheng, M. Zhou, X. Zhao, Enhanced photocatalytic activity of TiO2
Photogr. 2012 (2012) 1–8, http://dx.doi.org/10.1155/2012/626759. powder (P25) by hydrothermal treatment, J. Mol. Catal. A Chem. 253 (2006)
[60] H. Li, Y. Hao, H. Lu, L. Liang, Y. Wang, J. Qiu, X. Shi, Y. Wang, J. Yao, A systematic 112–118, http://dx.doi.org/10.1016/j.molcata.2006.03.021.
study on visible-light N-doped TiO2 photocatalyst obtained from ethylenediamine [89] H. Zhang, J.F. Banfield, Understanding polymorphic phase transformation behavior
by sol–gel method, Appl. Surf. Sci. 344 (2015) 112–118, http://dx.doi.org/10.1016/ during growth of nanocrystalline aggregates: insights from TiO2, J. Phys. Chem. B
j.apsusc.2015.03.071. 104 (2000) 3481–3487, http://dx.doi.org/10.1021/jp000499j.
[61] H. Han, R. Bai, Buoyant photocatalyst with greatly enhanced visible-light activity [90] P.P. Lottici, D. Bersani, M. Braghini, A. Montenero, Raman scattering characteriza-
prepared through a low temperature hydrothermal method, Ind. Eng. Chem. Res. tion of gel-derived titania glass, J. Mater. Sci. 28 (1993) 177–183, http://dx.doi.
48 (2009) 2891–2898, http://dx.doi.org/10.1021/ie801362a. org/10.1007/BF00349049.
[62] R. Asahi, Visible-light photocatalysis in nitrogen-doped titanium oxides, Science [91] D. Bersani, P.P. Lottici, X.-Z. Ding, Phonon confinement effects in the Raman scat-
293 (2001) 269–271, http://dx.doi.org/10.1126/science.1061051. tering by TiO2 nanocrystals, Appl. Phys. Lett. 72 (1998) 73, http://dx.doi.org/10.
[63] S. Livraghi, M.C. Paganini, E. Giamello, A. Selloni, C. Di Valentin, G. Pacchioni, 1063/1.120648.
Origin of photoactivity of nitrogen-doped titanium dioxide under visible [92] H. Lin, C. Huang, W. Li, C. Ni, S. Shah, Y. Tseng, Size dependency of nanocrystalline
light, J. Am. Chem. Soc. 128 (2006) 15666–15671, http://dx.doi.org/10.1021/ TiO2 on its optical property and photocatalytic reactivity exemplified by 2-
ja064164c. chlorophenol, Appl. Catal. B Environ. 68 (2006) 1–11, http://dx.doi.org/10.1016/j.
[64] G. Barolo, S. Livraghi, M. Chiesa, M.C. Paganini, E. Giamello, Mechanism of the apcatb.2006.07.018.
photoactivity under visible light of N-doped titanium dioxide. Charge carriers mi- [93] W.E. Slink, P.B. DeGroot, Vanadium-titanium oxide catalysts for oxidation of bu-
gration in irradiated N-TiO2 investigated by electron paramagnetic resonance, J. tene to acetic acid, J. Catal. 68 (1981) 423, http://dx.doi.org/10.1016/0021-
Phys. Chem. C 116 (2012) 20887–20894, http://dx.doi.org/10.1021/jp306123d. 9517(81)90113-5.
[65] T. Ochiai, A. Fujishima, Photoelectrochemical properties of TiO2 photocatalyst and [94] P. Prieto, R.E. Kirby, X-ray photoelectron spectroscopy study of the difference be-
its applications for environmental purification, J Photochem Photobiol C: tween reactively evaporated and direct sputter-deposited TiN films and their oxi-
Photochem Rev 13 (2012) 247–262, http://dx.doi.org/10.1016/j.jphotochemrev. dation properties, J. Vac. Sci. Technol. A 13 (1995) 2819, http://dx.doi.org/10.1116/
2012.07.001. 1.579711.
[66] M. Ceotto, L. Lo Presti, G. Cappelletti, D. Meroni, F. Spadavecchia, R. Zecca, M. Leoni, [95] J. Lynch, C. Giannini, J.K. Cooper, A. Loiudice, I.D. Sharp, R. Buonsanti, Substitutional
P. Scardi, C.L. Bianchi, S. Ardizzone, About the nitrogen location in nanocrystalline or interstitial site-selective nitrogen doping in TiO2 nanostructures, J. Phys. Chem. C
N-doped TiO2: combined DFT and EXAFS approach, J. Phys. Chem. C 116 (2012) 119 (2015) 7443–7452, http://dx.doi.org/10.1021/jp512775s.
1764–1771, http://dx.doi.org/10.1021/jp2097636. [96] A. Glaser, S. Surnev, F.P. Netzer, N. Fateh, G.A. Fontalvo, C. Mitterer, Oxidation of va-
[67] C.W. Dunnill, I.P. Parkin, Nitrogen-doped TiO2 thin films: photocatalytic applica- nadium nitride and titanium nitride coatings, Surf. Sci. 601 (2007) 1153–1159,
tions for healthcare environments, Dalton Trans. 40 (2011) 1635–1640, http:// http://dx.doi.org/10.1016/j.susc.2006.12.010.
dx.doi.org/10.1039/C0DT00494D. [97] O. Diwald, T.L. Thompson, T. Zubkov, E.G. Goralski, S.D. Walck, J.T. Yates, Photo-
[68] C. Shifu, C. Lei, G. Shen, C. Gengyu, The preparation of nitrogen-doped chemical activity of nitrogen-doped rutile TiO2 (110) in visible light, J. Phys.
photocatalyst TiO2 − xNx by ball milling, Chem. Phys. Lett. 413 (2005) 404–409, Chem. B 108 (2004) 6004–6008, http://dx.doi.org/10.1021/jp031267y.
http://dx.doi.org/10.1016/j.cplett.2005.08.038. [98] H. Irie, Y. Watanabe, K. Hashimoto, Nitrogen-concentration dependence on photo-
[69] C.W. Dunnill, A. Kafizas, I.P. Parkin, CVD production of doped titanium dioxide thin catalytic activity of TiO2 − xNx powders, J. Phys. Chem. B 107 (2003) 5483–5486,
films, Chem. Vap. Depos. 18 (2012) 89–101, http://dx.doi.org/10.1002/cvde. http://dx.doi.org/10.1021/jp030133h.
201200048. [99] S. Livraghi, M.R. Chierotti, E. Giamello, G. Magnacca, M.C. Paganini, G. Cappelletti, C.L.
[70] T.C. Jagadale, S.P. Takale, R.S. Sonawane, H.M. Joshi, S.I. Patil, B.B. Kale, S.B. Ogale, N- Bianchi, Nitrogen-doped titanium dioxide active in photocatalytic reactions with vis-
doped TiO2 nanoparticle based visible light photocatalyst by modified peroxide ible light: a, multi-technique characterization of differently prepared materials, J.
sol−gel method, J. Phys. Chem. C 112 (2008) 14595–14602, http://dx.doi.org/10. Phys. Chem. C 112 (2008) 17244–17252, http://dx.doi.org/10.1021/jp803806s.
1021/jp803567f. [100] S. Bae, S. Kim, S. Lee, W. Choi, Dye decolorization test for the activity assessment of
[71] F. Gherardi, A. Colombo, M. D'Arienzo, B. Di Credico, S. Goidanich, F. Morazzoni, F. visible light photocatalysts: realities and limitations, Catal. Today 224 (2014)
Simonutti, L. Toniolo, Efficient self-cleaning treatments for built heritage based on 21–28, http://dx.doi.org/10.1016/j.cattod.2013.12.019.
highly photo-active and well-dispersible TiO2 nanocrystals, Microchem. J. 126 [101] C. Guillard, H. Lachheb, A. Houas, M. Ksibi, E. Elaloui, J.M. Herrmann, Influence of
(2016) 54–62, http://dx.doi.org/10.1016/j.microc.2015.11.043. chemical structure of dyes, of pH and of inorganic salts on their photocatalytic deg-
[72] L. Bergamonti, I. Alfieri, A. Lorenzi, G. Predieri, G. Barone, G. Gemelli, P. Mazzoleni, radation by TiO2 comparison of the efficiency of powder and supported TiO2, J.
S. Raneri, D. Bersani, P.P. Lottici, Nanocrystalline TiO2 coatings by sol–gel: photocat- Photochem. Photobiol. A Chem. 158 (2003) 27–36, http://dx.doi.org/10.1016/
alytic activity on Pietra di Noto biocalcarenite, J. Sol-Gel Sci. Technol. 75 (2015) S1010-6030(03)00016-9.
141–151, http://dx.doi.org/10.1007/s10971-015-3684-6. [102] F. Chen, J. Zhao, H. Hidaka, Highly selective deethylation of rhodamine B: adsorp-
[73] E. Quagliarini, F. Bondioli, G.B. Goffredo, C. Cordoni, P. Munafò, Self-cleaning and tion and photooxidation pathways of the dye on the TiO2/SiO2 composite
de-polluting stone surfaces: TiO2 nanoparticles for limestone, Constr. Build. photocatalyst, Int. J. Photogr. 5 (2003) 209–217, http://dx.doi.org/10.1155/
Mater. 37 (2012) 51–57, http://dx.doi.org/10.1016/j.conbuildmat.2012.07.006. S1110662X03000345.

You might also like