You are on page 1of 9

Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

Contents lists available at ScienceDirect

Journal of Photochemistry & Photobiology, A: Chemistry


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

Facile and novel synthesis of nitrogen doped TiO2/Acid soluble


collagen-polyvinyl pyrrolidone (ASC-PVP) hybrid nanocomposite for rapid
and effective photodegradation of naphthalene from water under visible
light irradiation
Elaheh Hosseini Doabi a, Fatemeh Elmi a, *, Maryam Mitra Elmi b
a
Department of Marine Chemistry, Faculty of Marine & Environmental Sciences, University of Mazandaran, Babolsar, Iran
b
Research Center of Cellular and Molecular Biology, Health Research Center, Babol University of Medical Sciences, Babol, Iran

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

Keywords: A highly efficient, eco-friendly, and visible-light-driven photocatalyst, namely nitrogen doped titanium dioxide
N-TiO2/ASC-PVP hybrid nanocomposite (N-TiO2)/collagen-polyvinyl pyrrolidone (ASC-PVP) hybrid nanocomposite is prepared via facile sol–gel method.
Naphthalene The ASC-PVP steadily covers the N-TiO2. The XPS data confirmed the oxygen vacancies and nitrogen interstitial-
Photodegradation
type doping in N-TiO2 prepared by the sol–gel route. Furthermore, the O 1s and N 1S XPS data prove that both
Visible light
nitrogen and oxygen atoms of ASC-PVP are participated in hydrogen bonds with N-Ti-O and/or Ti-O-N bonds in
N-TiO2 nanoparticles. Moreover, the N-TiO2/ASC-PVP is shown an obvious red shift of absorption band edge and
reduced recombination rate of electron-hole in comparison with N-TiO2 nanoparticles. N-TiO2/ASC-PVP was
used to study photocatalytic degradation of a typical polycyclic aromatic hydrocarbon (PAHs), namely naph­
thalene in water under visible light irradiation. N-TiO2/ASC-PVP possessed high photocatalytic activity for the
naphthalene degradation (86%) within 100 min under visible light. The inhibitory effect on the naphthalene
photodegradation indicated that O−2 . and h+⋅are the major reactive species. Also, the N-TiO2/ASC-PVP exhibited
very good visible light photocatalytic efficiency for the removal of naphthalene at very low concentrations from
seawater.

1. Introduction and thermal stability, low cost, availability, and non-toxicity [6]. The
heterogeneous photocatalytic process is an advanced oxidation process
Currently, water scarcity is one of the most important global chal­ for water treatment that takes place by directing sunlight or ultraviolet
lenges. Population growth, industrialization of communities, pollution light onto the surface of a semiconductor. The heterogeneous photo­
of existing freshwater resources, and climate change have made it catalysis reactions are a promising technology for the degradation of
difficult to provide safe drinking water [1]. One of the most important toxic organic and inorganic contaminants from water, and hydrogen
water pollutants is polycyclic aromatic hydrocarbons (PAHs), which are production [7–11].
classified as toxic, carcinogenic, and immunotoxic pollutants [2,3]. The The dependence of TiO2 photocatalytic activity on UV light, which
simplest member of PAHs is naphthalene (C10H8), which is one of the contains only 4% of sunlight, has limited its use [6]. In addition, the
polycyclic compounds in crude oil and is found in insecticides, cleaners, possibility of electron-hole pair recombination also reduces the effi­
and dyes [4]. In recent years, the use of photocatalyst in advanced ciency of TiO2 [12]. To solve this problem, photocatalyst modification
oxidation processes (AOP) has attracted the attention to remove water strategies such as surface chemical modification and doping are used
pollution because they do not cause secondary pollution and have the [13–18]. Doping of TiO2 with metal elements such as Au, Fe, and Ag
ability to convert organic pollutants into harmless substances such as increases the charge transfer property and the stability of the electron-
H2O and CO2 [5]. One of the most important photocatalysts is titanium hole pair in the photocatalyst [19,20]. The doping of TiO2 with
dioxide, which is used due to its high photocatalytic activity, chemical different types of metallic or non-metallic ions is done as substitution or

* Corresponding author.
E-mail address: f.elmi@umz.ac.ir (F. Elmi).

https://doi.org/10.1016/j.jphotochem.2021.113677
Received 5 August 2021; Received in revised form 10 November 2021; Accepted 19 November 2021
Available online 24 November 2021

1
E.H. Doabi et al. Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

interstitial. In recent years, the use of non-metallic elements such as C, software 2.3.23 (Casa Software Ltd., UK). All spectra were corrected by
N, and S has been considered due to the reduction of energy bandgap moving the position of the primary C 1s core level to 284.6 eV. The
(Eg), recombination rate, and increase of photocatalytic activity in the curves were fit with sum Gaussian-Lorentzian line shapes following
visible region [21–23]. Among non-metallic elements, N is important Shirley’s background subtraction.
due to its low ionization energy and high stability [24]. Since the cat­
alytic activity depends not only on the type of material but also on its 2.1. Synthesis of N-TiO2 nanoparticles
particle size, the use of nanoscale photocatalysts increases the surface-
to-volume ratio and improves the catalytic properties. It is necessary The sol–gel method was used to synthesize N-TiO2. In this method,
to obtain the optimal semiconductor particle size to improve the process. TiO2 and urea were combined in mass ratios of 1:2, 1:3, and 1:5. For the
Reducing the particle size below the optimum value causes surface synthesis of 1:2, 1.2 g of TiO2 was added to 10 ml of distilled water and
recombination to happen faster than the interfacial charge transfer placed in an ultrasonic bath at room temperature for 10 min. Then, it
procedure. The optimum size for TiO2 as a photocatalyst is about 10 nm was added to 2.4 g of urea in 10 ml of distilled water. The sample was
[25]. stirred at room temperature for 1 h and was kept in the dark for 24 h.
Organic-inorganic hybrid nanocomposite offers a novel route of Next, the sample was placed in a vacuum oven at 350 ◦ C for 1 h. Finally,
nanomaterials with preferred applications in various fields such as the cream-colored powder was washed with distilled water to reach pH
health, optics, energy storage, and the environment. Hybrid nano­ = 7 [28]. The same procedure was used to synthesize N-TiO2 nano­
materials display better properties in comparison to their counterparts, particles with other mass ratios.
wherein inorganic material performs several important roles such as
enhancing the thermal and mechanical stability, providing an accessible
2.2. Preparation of N-TiO2/ASC-PVP hybrid nanocomposite
porosity through an interconnected porous network for catalyst or
sensing, and chemical or electrochemical properties [26]. Organic ma­
0.5 g of N-TiO2 was dissolved in 10 ml of 10 × PBS. Then, 0.1 g of
terials give opportunities to improve mechanical properties allowing the
PVP in 5 ml of 10 × PBS was added and stirred at room temperature for
preparation of better fibers and films, controlling the porosity and
1 h. ASC was extracted from white fish scales by the acidic method [29].
porous network connectivity, and modifying the hydrophilic-
0.05 g of ASC in 5 ml of 0.5 M acetic acid added to N-TiO2/PVP. The pH
hydrophobic balance. Organic matter can play a role in a particular
of the sample was adjusted to 7.6 using 0.1 M NaOH. The sample was
chemical or physical substance, such as electrical or optical behaviors,
placed in a shaker incubator for 1 week at 10 ◦ C. The milky solid phase
electrochemical properties, and biochemical or chemical reactions [26].
was separated from the liquid using filter paper. The sample was dried at
In this study, a new hybrid nanocomposite comprising acid-soluble
room temperature.
collagen (ASC), polyvinyl pyrrolidone (PVP), and nitrogen doped tita­
nium dioxide (N-TiO2) is fabricated to serve as a novel photocatalyst for
naphthalene degradation under visible light. It should be noted that PVP 2.3. Photocatalytic activity of N-TiO2/ASC-PVP
is an important cross-linker to increase the strength of N-TiO2 and ASC
bonds as well as a non-toxic polymer. The use of PVP prevents the N-TiO2/ASC-PVP (4 mg) was added to 40 ml of naphthalene (15 mg/l
agglomeration of nanoparticles, improves the electro-optical properties, in distilled water/ethanol, 1: 1 V/V), pH = 6.5 ± 0.1. It was then sub­
and reduces the Eg in the UV and visible regions [27]. The goal of this jected to ultrasonic for 5 min and then stirred in the dark at room
study is to investigate the structural properties of N-TiO2/ASC-PVP temperature. Samples (3 ml) were collected at 10, 20, 30, 40, 50, 60, 70,
hybrid nanocomposite and to demonstrate its photocatalytic activity 80, 90, and 100 min intervals. After sample collection, the nano­
toward the removal of naphthalene contaminant in an aqueous medium. composite was separated under centrifugation at 10000 rpm. The
absorbance of the samples was read at 276 nm by UV–Vis spectroscopy.
2. Materials and methods The equilibrium adsorption capacity (qe) of naphthalene onto photo­
catalysts was calculated according to the following equation:
The materials used in this study have the highest degree of purity. (c0 − ce ) V
FT-IR spectroscopy (Model Tensor 27, Bruker, Germany) was used to qe = (1)
m
determine the functional groups of N-TiO2 and N-TiO2/ASC-PVP. The
denaturation temperature (Td) of the hybrid nanocomposite was deter­ where c0 is the initial concentration (mg/l), ce is the final concentration
mined using differential scanning calorimetry (DSC) (200-F3Maia, (mg/l) of naphthalene, V is the volume of the solution (l), and m is the
NETZSCH, Germany). Morphology and particle sizes of N-TiO2 and N- mass of the photocatlyst (g). The rate of naphthalene removal was
TiO2/ASC-PVP were characterized using a transmission electron mi­ determined according to Eq. (2):
croscope (TEM) (Zeiss EM900, Germany) and field emission-scanning
(c0 − ce )
electron microscope (FE-SEM) (Mira III, Czech Republic). The R(%) = × 100 (2)
c0
Brunauer-Emmett-Teller (BET) analysis (BEL, Belsorp II, Japan) was
used to determine the specific surface area, mean diameter, and total Before irradiation, the sample was stirred for 60 min in the dark at
volume of pores. XRD studies on N-TiO2 and N-TiO2/ASC-PVP (XRD- room temperature to achieve adsorption–desorption equilibrium. Then,
PW1730) were recorded with Cu lamp irradiation (40 kV and 30 mV it was stirred under visible light illumination using a 150 W xenon lamp
current) at 2θ angles equal to 10–80◦ . XRD analysis was explained in by a cut-off filter (λ > 420 nm). During the irradiation period at each
Supporting information (Fig. s1). The fate of the photogenerated interval of 10 min (10–80 min), the solution was isolated from the
electron-hole pairs for N-TiO2 and the N-TiO2/ASC-PVP were studied by sample. Then, N-TiO2/ASC-PVP was separated by centrifugation at
photoluminescence (PL) spectroscopy (Avaspec 2048 TEC, Netherlands) 10000 rpm. The absorbance of the solution was measured by UV–Vis
at the excitation wavelength of 320 nm. The diffuse reflectance spec­ spectroscopy at λmax = 276 nm.
troscopy (DRS) (S-2014, Scinco, South Korea) was used to calculate the The photocatalytic degradation efficiency (η) was calculated using
band gap for N-TiO2 and N-TiO2/ASC-PVP. Two-beam optical spec­ Eq. (3):
troscopy (UV–VIS-T80+, PG Instrument Limited, UK) was used to study c0 − c
the photodegradation of naphthalene. The x-ray photoelectron spec­ η(%) = × 100 (3)
c0
troscopy (XPS) was measured using XPS (BESTEC, Germany) with a
monochromatic Al Kα x-ray source (1486.6 eV) operating at ultrahigh where c0 is the initial naphthalene concentration at irradiation time of
vacuum (10-7 Pa). Peak fitting was performed using Casa XPS analysis zero, and c is the concentration of the naphthalene at various irradiation

2
E.H. Doabi et al. Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

time intervals. Furthermore, the kinetics photodegradation rate of Ti–O–Ti and Ti–O. The absorption peak in the range of 3650–3000
naphthalene over the as-prepared photocatalysts was calculated using cm− 1 is related to the vibrations of the OH groups. The sharp and strong
the Langmuir–Hinshelwood model Eq. (4): peak at 1700 cm− 1 is related to the stretching vibrations of C–
– O of urea
− 1 − 1
(c ) precursor. The peaks of 1452 cm and 1051 cm are attributed to the
(4)
0
ln
c
= tKapp asymmetric stretching vibrations of N–O and bending vibrations of
O–N–O bonds, respectively [30]. In the FT-IR spectrum for N-TiO2/
where kapp is the apparent pseudo-first-order rate constant; c0 and ct ASC-PVP nanocomposite, the peaks appearing in the range of 700–450
denote the initial and at time t irradiation time concentrations of cm− 1 belong to N–TiO2. The band that appeared at 1290 cm− 1 corre­
naphthalene, respectively. sponds to symmetric stretching vibrations of N–O. The absorption peak
due to asymmetric stretching vibrations of N–O (1452 cm− 1) to a lower
2.4. Photocatalytic activity of N-TiO2/ASC-PVP in the presence of frequency (1421 cm− 1) can be due to the participation of this bond in
scavengers hydrogen bonding with functional groups of a homogenous blend of ASC
and PVP polymer. The band at 1620 cm− 1 is associated with amide I
First, 4 mg of N-TiO2/ASC-PVP was added to 40 ml of naphthalene (stretching of C– – O weakly coupled with N–H bending and C–N
(15 mg/l). Then, 2 mg of EDTA was added as a hole scavenger. The stretching vibrations). The changes in vibrational signatures are an
sample was subjected to ultra-sonication at pH = 6.5 ± 0.1 for 5 min. indication of N–TiO2 and ACS-PVP interactions.
Then, the sample was moderately stirred on a hot plate in the dark for
60 min. Subsequent steps were performed as described in section 2.3. 3.2. DSC analysis
Similar experiments were performed in the presence of 0.2 ml of 2-prop­
anol (hydroxyl radical scavenger), 2 mg of p-benzoquinone (superoxide According to the DSC analysis (Fig. s2), the melting temperature of
radical scavenger), 2 mg of KI (the adsorbed hydroxyl radical and N-TiO2/ASC-PVP is 51.5 ◦ C. The denaturation temperature of ASC
valence band hole scavenger), and 2 mg of AgNO3 (electron scavenger). extracted from white fish scales has been reported to be 19 ◦ C [29].
Strong interactions between N-TiO2 nanoparticles with collagen fibrils
strengthen the ASC structure. In addition, the presence of PVP increases
2.5. The effect of N-TiO2/ASC-PVP on the photocatalytic removal of
the thermal resistance by increasing the activation energy. The area
naphthalene from seawater
below the thermogram curve showed the total enthalpy of melting ΔH
= 194.33 J/g.
Caspian Seawater was collected from the clean area close to Sari,
Mazandaran province. The sample was immediately transferred to the
3.3. BET analysis
marine science laboratory. The sample was filtered using sterile syringe
filters (FilterBio® 0.45 µm). Next, 4 mg of N-TiO2/ASC-PVP was added
The BET plots for N-TiO2 and the N-TiO2/ASC-PVP hybrid nano­
to 40 ml of naphthalene (120 μg/l in seawater/ethanol 1:1, V/V). The
composite are shown in Fig. s3. Both of these diagrams correspond to the
pH was adjusted to 6.5 ± 01. The sample was ultra-sonicated for 5 min
type IV isothermal diagram based on the IUPAC classification, which
and then moderately stirred for 60 min in the dark. Subsequently, it was
belongs to the mesoporous materials. In addition, the hysteresis loop for
exposed to visible light from a xenon lamp for 100 min. Every 10 min, 3
both samples is H3 type, indicating slit-shaped pores. The specific sur­
ml of sample was collected, and 4 ml of dichloromethane was added to
face area (SBET), average pore diameter (Dp), and the pore volume (Vp)
the vial. The resulting cloud solution was centrifuged, and the organic
obtained were represented in Table s1. The results show that the samples
phase absorption was recorded at λmax = 276 nm.
have low porosity.

3. Results and discussion


3.4. FE-SEM images

3.1. FT-IR analysis


Fig. 2 shows the FE-SEM images for N-TiO2 and N-TiO2/ASC-PVP
with different magnifications. As can be seen, N-TiO2 nanoparticles in
The FT-IR spectra for N-TiO2 and N-TiO2/ASC-PVP are shown in
agglomerates are pseudo spherical in shape. Fig. 2c and 2d show a liner
Fig. 1. In the N-TiO2 spectrum, the absorption band at 540 cm− 1 is
of ASC-PVP polymer covering the entire surface of N-TiO2 nanoparticles.
related to the bending vibrations of the Ti-O bond. The peaks in the
range of 1900–600 cm− 1 are the result of the stretching vibrations of
3.5. TEM images

The TEM images at different magnifications show that the N-TiO2


nanoparticles are stacked together, and the ASC-PVP homogenous
polymer covers all of the surfaces of N-TiO2 nanoparticles, Fig. s4.

3.6. DRS analysis

The UV–Vis DRS results of N-TiO2(I), N-TiO2(II), and N-TiO2(III)


samples with TiO2 to urea weight ratios of 1:2, 1:3, and 1:5, respectively,
is shown in Fig. 3a. According to Fig. 3a, the shift of absorption band
edge to higher wavelength is observed for N-TiO2/ASC-PVP compared to
N-TiO2 nanoparticles. The Tauc plots show that the Eg of N-TiO2(I), N-
TiO2(II), and N-TiO2(III) are equal to 3.04 eV, 2.86 eV, and 2.81 eV,
respectively (Fig. 3b). The Eg of the TiO2 anatase is 3.2 eV requires UV
light irradiation [31]. The location of nitrogen doping in the TiO2 crystal
lattice is crucial for its behavior as a photocatalyst, but it is debatable
which position is more advantageous. Asahi et al. reported that nitrogen
Fig. 1. FT-IR spectra of N-TiO2 nanoparticles and N-TiO2/ASC-PVP hybrid substitutional-type doping is effective for the narrowing of the bandgap
nanocomposite. of TiO2 anatase phase based on the mixing of O 2p with N 2p states in the

3
E.H. Doabi et al. Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

Fig. 2. FE-SEM images of (a) N-TiO2 with 200 kx magnification, and (b) N-TiO2 with 40 kx magnification, (c) N-TiO2/ASC-PVP with 200 kx magnification, and (d) N-
TiO2/ASC-PVP with magnification 50 kx.

valence band due to spin-restricted local density approximation calcu­


lations [32]. Another approaches confirmed that nitrogen interstitial-
type doping was associated with the photothreshold energy decrease,
which caused the formation of localized N 2p energy states within the
bandgap lying above the top of the valence band, facilitating the oxygen
vacancies generation [22,33].
Many experiment results confirm that during nitrogen doping, both
substitutional and interstitial types occur in the TiO2 network. The
presence of nitrogen can alter the band structure or suppress the
recombination of photogenerated electron-hole pairs. As a result, it in­
creases the photocatalytic activity of TiO2 in the blue-violet region of the
visible light spectrum (3.26–2.63 eV) [23]. According to Fig. 3b, N-
TiO2(III) nanoparticles possessed the smallest Eg in comparison to N-
TiO2(I) and N-TiO2(II) nanoparticles. Therefore, N-TiO2(III) was used to
synthesize N-TiO2/ASC-PVP hybrid nanocomposite. The Eg of N-TiO2/
ASC-PVP is equal to 2.77 eV (Fig. 3b). Light absorption intensity and
photocatalytic performance of N-TiO2/ASC-PVP are increased compared
to N-TiO2 under visible light.

3.7. Analysis of PL spectra

Intrinsic luminescence emissions (blue/violet) for the anatase phase


are proportional to the structural disorders in the hexacoordinated
structure of TiO6 [34,35]. Extrinsic luminescence (yellow/green) has
been attributed to surface disorders and oxygen vacancies [36,37]. Lee
et al. have shown that the intensity of nitrogen-doped TiO2 photo­
luminescence is lower than that of TiO2. This indicates that the electron-
hole recombination rate is slower in N-TiO2 than pure TiO2 [38]. As
shown in Fig. 4, the excitation wavelength is λexc = 320 nm (3.87 eV),
which is larger than the Eg values of N-TiO2 and N-TiO2/ASC-PVP. The
luminescence spectra were fitted with Lorentzian band shapes. The
emission peaks of N-TiO2 are centered at 367 nm (3.38 eV), 368 nm
(3.37 eV), 371 nm (3.34 eV), 380 nm (3.26 eV), 382 nm (3.24 eV), 455
nm (2.72 eV), 459 nm (2.70 eV), and 475 nm (2.61 eV), Fig. 4a. These
emission spectra in the region of UV-A light (3.26–3.23 eV) and violet
Fig. 3. (a) UV–Vis absorption spectra DRS spectra, and (b) Tauc plots of N- (3.26–2.85 eV) are corresponding with the intrinsic luminescence
TiO2(I), N-TiO2(II), N-TiO2(III), and N-TiO2/ASC-PVP. emissions of N-TiO2 [35,39].
There are emission peaks of N-TiO2/ASC-PVP are located at 394 nm

4
E.H. Doabi et al. Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

the photocatalytic properties. The XPS spectra of N-TiO2 and N-TiO2/


ASC-PVP samples are shown in Fig. 5. The surfaces of both samples are
composed of Ti, N, O, and C atoms. The N-TiO2/ASC-PVP is synthesized
in 10 × PBS, so sodium element is also detected by XPS. Fig. 6 displays N
1s peaks of the corresponding XPS spectra. In N-TiO2 spectra, one peak
with high intensity centered at 400.3 eV is detected. The N 1s peak
located at binding energies above 400 eV has been variously associated
with different interstitial N species such as N2, NO2– and NHx [43–46].
It is out of the scope of the present study to define the exact chemical
identity of the interstitial N species; however, in accordance with several
reports, the N 1s peak at 400.3 eV can be attributed to interstitial ni­
trogen with possible Ti-O-N and/or Ti-N-O linkages [47–48]. The N 1s
binding energy of N-TiO2/ASC-PVP is located at 398.6 eV. The N 1s
binding energies of pure PVP and pure collagen are located at 399.6 eV
and 398 eV, respectively [49,50]. It is considered that the shift of the
binding energy of nitrogen atoms is due to the formation of hydrogen
bonds in the N–H and C– – O bonds of ASC-PVP polymer and N–Ti–O
and/ or Ti–O–N bonds as shown in the FT-IR results.
Fig. s5 shows the high-resolution spectra of Ti 2p3/2 and Ti 2p1/2
regions of prepared samples. The XPS spectrum of the Ti 2p core level,
obtained from N–TiO2 (N–TiO2/ASC-PVP), exhibited a Ti 2p3/2 peak at
458 (458.2 eV) and a Ti 2p1/2 peak at 463.6 (463.9 eV). The difference in
binding energy between the two peaks 5.6 (5.7 eV) corresponds to the
Ti4+ oxidation state in the octahedral environment of the anatase TiO2
[51,52]. Also, the shoulder Ti 2p1/2 at a binding energy of 460.6 (461.4)
eV is related to Ti3+, indicating a small contribution of Ti2O3 [53]. The
presence of Ti3+ in TiO2 lattice is usually due to the oxygen vacancies
[54]. It is reported that the presence of nitrogen dopant notably facili­
tates oxygen vacancies formation in TiO2 [55]. Batalović et al. showed
that the combined effect of interstitial nitrogen and oxygen vacancies
induces bandgap reduction and enhances the optical properties of
anatase TiO2 [56].
The single O 1s peak in the N-TiO2 spectrum can be divided into

Fig. 4. PL spectra of (a) N-TiO2, and (b) N-TiO2/ASC-PVP.

(3.15 eV), 405 nm (3.06 eV), 417 nm (2.97 eV), 432 nm (2.87 eV), 442
nm (2.80 eV), 452 nm (2.74 eV), 468 nm (2.65 eV), 488 nm (2.54 eV),
512 nm (2.42 eV), 530 nm (2.34 eV), and 550 nm (2.25 eV) Fig. 4b.
Compared to the N-TiO2 PL spectrum, the emission intensity is higher in
the N-TiO2/ASC-PVP hybrid nanocomposite, and peak emissions have
centered at larger wavelengths. This suggests that more surface states
are available in N-TiO2/ASC-PVP than N-TiO2. The increase in emission
intensity can be due to a blend of ASC-PVP polymer, charge transfer
between ASC-PVP molecular orbital energy levels and deep, and shallow
traps in N-TiO2.
Research has shown that PVP has a maximum PL emission wave­
length at 394 nm with λexc = 325 nm, which is related to the electron
transfer from the lowest unoccupied molecular orbital (LUMO) to the
highest molecular orbital (HOMO) [40]. This radiative relaxation of the
electrons could be involved in the formation of a strong emission band at
394 nm for the N-TiO2/ASC-PVP hybrid nanocomposite. When collagen
is exposed to light, it emits fluorescence due to the amino acid residues
of tyrosine, phenylalanine, and tryptophan [41]. Reports in this field
indicate that phenylalanine amino acid residues play a major role in the
emission PL spectrum of collagen in the visible light region with a
maximum PL emission wavelength at 417 nm (2.97 eV) at λexc = 337 nm
(3.68 eV) [42]. Thus, collagen may play a role in the formation of an
emission band at 417 nm for N-TiO2/ASC-PVP.

3.8. XPS analysis

XPS analysis of samples is carried out to gain useful information


about surface chemical bonds that affects the optical band gap and thus Fig. 5. XPS spectra of N-TiO2 and N-TiO2/ASC-PVP.

5
E.H. Doabi et al. Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

Fig. 7. Photodegradation activities of N-TiO2 (4 mg), and N-TiO2/ASC-PVP (4


mg) under visible light irradiation (λ > 420 nm). Experimental condition:
naphthalene: 15 mg/l, temperature: 25 ± 1 ◦ C, pH = 6.5 ± 0.1, and time period:
60 min (dark) and 100 min (light on).

and reusability of the N-TiO2/ASC-PVP, Fig. 8. After four cycle tests, the
photocatalyst exhibits a high photodegradation rate and photo­
degradation efficiency of 81.4%, Table s2. This shows that N-TiO2/ASC-
PVP can be used as an efficient photocatalyst with excellent cycle per­
formance to degrade naphthalene in aqueous solutions. Fig. s7 shows
FT-IR spectrum of N-TiO2/ASC-PVP nanocomposite after four recycling
test. The peaks in the range of 700–450 cm− 1 are related to N–TiO2. The
band at 1290 cm− 1 corresponds to symmetric stretching vibrations of
Fig. 6. High resolution XPS spectra of N 1s region of N-TiO2 and N-TiO2/ASC- N–O. The absorption peak at 1450 cm− 1 is associated with asymmetric
PVP samples. stretching vibrations of N–O. The absorption peak in the range of
3650–3000 cm− 1 is due to the vibrations of the hydroxyl groups. The
three peaks (Fig. s6). The low binding one located at 530.1 eV is assigned band at 1730 cm− 1 is associated with the carbonyl stretching vibrations.
to the oxygen of Ti–O bonds in TiO2 [57]. The other one at 531.2 eV is The main amide I absorption is shifted from 1620 cm− 1 to 1640 cm− 1, a
typical for the oxygen of Ti-OH bonds [57]. The third peak appears at shift indicative of weakening the intermolecualr hydrogen bonds be­
532.1 eV. This feature was first reported by Saha and Tompkins [57]. tween N–TiO2 and blend of polymers after 4 cycles.
György et al. was first characterized it by a depth profiling study on TiN
surfaces [58]. They attributed this feature to the oxidation of Ti–N,
3.10. Proposed mechanism of naphthalene photodegradation
which leads to the Ti–O–N bond. Chen and Burda proposed that this
peak appearance is in accordance with the nitrogen substitution in the
The proposed scheme for naphthalene degradation by N-TiO2/ASC-
TiO2 matrix and reveals the O–Ti–N bond formation [59].
PVP presented in Fig. 9 can be summarized as follows: when a photon of
We also suggest that the appearance of this peak can be attributed to
energy equal to or greater than the Eg hits the photocatalyst an electron
the Ti–O–N and/or Ti–N–O linkages mainly in the surface, which is
in valence band (VB) gets excited to the conduction band (CB), leading
associated with N 1s at 400.3 eV. The high-resolution O 1s XPS spectra of
to the simultaneous production of a hole in the VB (h+) and excess
N-TiO2/ASC-PVP could be illustrated into two characteristic peaks at
529.5 eV and 531.3 eV. The O 1s peak with its binding energy at 531.3
eV could be assigned to chemisorbed water molecules on N-TiO2/ASC-
PVP. The O 1s binding energy at 529.5 eV could be assigned to ASC-PVP
blend polymer chemisorbed on the N-TiO2 surface. This indicates that
the O atoms in ASC-PVP blend polymer are involved in the chemisorp­
tion on the N-TiO2 surface.

3.9. Photocatalytic degradation of naphthalene by N-TiO2 and N-TiO2/


ASC-PVP

Fig. 7 shows that the removal percentage of naphthalene on N-TiO2


and N-TiO2/ASC-PVP under the adsorption process in the dark is 41.9%
and 43.4%, respectively. The adsorption capacity (qe) of naphthalene
onto N-TiO2 and N-TiO2/ASC-PVP is 62.81 mg/g and 63.22 mg/g,
respectively. The photocatalytic degradation efficiency (η%) of N-TiO2
and N-TiO2/ASC-PVP is 47.10% and 86%, respectively. The surface Fig. 8. Cycling test for the photodegradation of naphthalene with N-TiO2/ASC-
modification of N-TiO2 with ASC-PVP polymer greatly improved the PVP (4 mg) under visible light irradiation (λ > 420 nm). Experimental condi­
photocatalytic activity under visible light. tion: naphthalene: 15 mg/l, temperature: 25 ± 1 ◦ C, pH = 6.5 ± 0.1, and time
Furthermore, cycling experiments were done to evaluate the stability period: 60 min (dark) and 100 min (light on).

6
E.H. Doabi et al. Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

Fig. 9. Schematic diagram of photocatalytic degradation of naphthalene by N-TiO2/ASC-PVP.

electrons in the CB (e− ) (Eq. (5)).

N − TiO2 /ASC − PVP


(5)

× → N − TiO2 /ASC − PVP(h+ −
VB ) + N − TiO2 /ASC − PVP(e )

These photogenerated holes (h+) possess adequately positive po­


tential to produce OH. radicals from H2O molecules adsorbed on surface
of the photocatalyst (Eq. (6)) and/or from hydroxyl ions (Eq. (7)) which
can subsequently oxidize naphthalene (Eqs. (8) and (9)) [60].

N − TiO2 /ASC − PVP(h+ .


VB ) + H2 Oads →OH + H
+
(6)

N − TiO2 /ASC − PVP(h+ −


VB ) + OH →OH
.
(7)

N − TiO2 /ASC − PVP(h+


VB ) + naphthalen e→oxidation of naphthalen e (8)

OH . + naphthalene→degradation of naphthalene (9)


The presence of oxygen in the conduction band can produce O−2 . (Eq.
(10)), that can generate organic peroxide (Eq. (11)) and hydrogen
peroxide (H2O2) (Eq. (12)) that also can produce OH. radicals (Eq. (13))
Fig. 10. Effect of different scavengers on photocatalytic degradation of naph­
[60].
thalene under visible light. Experimental conditions: 15 mg/l initial naphtha­
N − TiO2 /ASC − PVP(e−CB ) + O2 →O−2 . (10) lene concentration, photocatalyst dosage: 4 mg, temperature: 25 ± 1 ◦ C, pH =
6.5 ± 0.1, and time period: 100 min (light on).
O−2 . + naphthalene→naphthalene − O − O. (11)

O−2 . + HO.2 + H + →H2 O2 + O2 (12) Table 1


Apparent pseudo-first order constants and η(%) in the presence of different
N − TiO2 /ASC − PVP(e−CB ) + H2 O2 →OH . (13) scavengers. Experimental conditions: 15 mg/l initial naphthalene concentration,
photocatalyst dosage: 4 mg, time period: 100 min (light on) at 25 ± 1 ◦ C.

3.11. Evaluation the photocatalytic degradation of naphthalene in the η(%) Kapp (min− 1) Scavengers
presence of scavengers 59.1 0.0089 Iso-propanol (200 µl)
42.4 0.0051 AgNO3 (2 mg)
It seems that reactive species such as OH. , O−2 . , h+, and e- may be 56.5 0.0045 KI (2 mg)
33 0.0039 EDTA (2 mg)
involved in photocatalytic processes in the degradation of organic pol­
31 0.0036 P-BQ (2 mg)
lutants. To investigate major photoactive species involved in the pho­ 86 0.0115 None
tocatalytic procedure under visible light irradiation, radical trapping
experiments are carried out. In the present work, the main reactive
species responsible for degrading of naphthalene are determined by 3.12. Evaluation of the photocatalytic degradation of naphthalene in
using EDTA, KI, AgNO3, parabenzoquinone (P-BQ), and isopropanol as seawater by N-TiO2/ASC-PVP
the scavengers of h+, adsorbed OH. and valence band holes, e-,⋅O−2 . ,
and⋅OH. , respectively. The results show that the maximum photodegradation efficiency of
P-BQ and EDTA possess a more remarkable inhibitory effect on the N-TiO2/ASC-PVP under visible light irradiation in seawater is 78%. The
naphthalene photodegradation than other scavengers, indicating that results indicate that the N-TiO2/ASC-PVP hybrid nanocomposite has a
O−2 . and h+• is the major reactive species (Fig. 10 and Table 1). There are reasonable efficiency for the removal of crude oils, the most prominent
only a few OH. generated during the photocatalytic reaction. Based on of which is naphthalene, Fig. 11. However, it seems salinity suppressed
these results, the possible photocatalytic mechanism of N-TiO2/ASC- the photocatalytic performance of N-TiO2/ASC-PVP, which was majorly
PVP is that the aromatic rings of naphthalene can be readily adsorbed on caused by the presence of cations (Na+, Ca2+, Mg2+,…) and anions (Cl-,
photocatalyst due to the π-π stacking adsorption between benzene rings SO42-, Br-, CO32–, HCO3–,…) in seawater. These ions may trap O−2 . and h+
of naphthalene and pyrrole rings of ASC-PVP polymer. Under visible reactive species. So, there exist fewer reactive species for the photo­
light irradiation, the excited electrons react with oxygen molecules degrading process.
producing O−2 . . Both h+ and O−2 . can decompose naphthalene molecules.

7
E.H. Doabi et al. Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

Acknowledgements

This work was supported by the Department of Marin Chemistry,


Faculty of Marine & Environmental Sciences, University of Mazandaran,
Babolsar, Iran.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.


org/10.1016/j.jphotochem.2021.113677.

References

[1] M. Elimelech, W.A. Phillip, The future of seawater desalination: energy,


technology, and the environment, Science 333 (6043) (2011) 712–717, https://
doi.org/10.1126/science:1200488.
[2] E.D. Eweis, J. B., Ergas, S. J., Chang, D. P. Y., & Schroeder, Bioremediation
principles , McGraw-Hil, McGraw-Hill, Toronto. 1998.
[3] K. Srogi, Monitoring of environmental exposure to polycyclic aromatic
Fig. 11. Photocatalytic degradation of naphthalene pollutant in seawater under
hydrocarbons: a review, Environ. Chem. Lett. 5 (2007) 169–195, https://doi.org/
visible light. Experimental conditions: 120 μg/l initial naphthalene concentra­ 10.1007/s10311-007-0095-0.
tion, photocatalyst dosage: 4 mg, time period: 100 min (light on) at 25 ± 1 ◦ C. [4] S.K. Samanta, O.V. Singh, R.K. Jain, Polycyclic aromatic hydrocarbons:
environmental pollution and bioremediation, Trends Biotechnol. 20 (6) (2002)
243–248, https://doi.org/10.1016/S0167-7799(02)01943-1.
4. Conclusions [5] C.h. Kutal, N. Serpone, Photosensitive metal—organic systems: Mechanistic
principles and applications, Advances in Chemistry, American Chemical Society,
Marine petroleum pollution caused by operational effluent dis­ Washington, DC, 1993.
[6] R. Daghrir, P. Drogui, D. Robert, Modified TiO2 for environmental photocatalytic
charges has been considered a challenging problem. More research ef­ applications: a review, Ind. Eng. Chem. Res. 52 (10) (2013) 3581–3599, https://
forts are needed to improve effective treatment methods for removing doi.org/10.1021/ie303468t.
dissolved organic matters, particularly PAHs, and to understand the [7] A.O. Ibhadon, P. Fitzpatrick, Heterogeneous photocatalysis: recent advances and
applications, Catalysts 3 (1) (2013) 189–218, https://doi.org/10.3390/
related mechanisms. This study targeted photocatalytic degradation of a catal3010189.
typical PAH, namely naphthalene in water under visible light irradia­ [8] M.Y. Ghaly, T.S. Jamil, I.E. El-Seesy, E.R. Souaya, R.A. Nasr, Treatment of highly
tion. We have demonstrated a facile and novel approach to synthesize N- polluted paper mill wastewater by solar photocatalytic oxidation with synthesized
nano TiO2, Chem. Eng. J. 168 (1) (2011) 446–454, https://doi.org/10.1016/j.
TiO2/ASC-PVP for visible light photocatalytic applications. The sol–gel cej.2011.01.028.
method leads to surface functionalization of N-TiO2 nanoparticles with [9] X. Wang, J. Liu, S. Leong, X. Lin, J. Wei, B. Kong, Y. Xu, Z.-X. Low, J. Yao, H. Wang,
enhancing their photocatalytic activity. Furthermore, the N-TiO2/ASC- Rapid construction of ZnO@ZIF-8 heterostructures with size-selective
photocatalysis properties, ACS Appl. Mater. Interfaces 8 (14) (2016) 9080–9087,
PVP hybrid nanocomposite is shown an obvious shift of absorption band
https://doi.org/10.1021/acsami.6b00028.
edge to higher wavelength and decreased recombination rate of [10] F.-J. Zhang, X. Li, X.-Y. Sun, C. Kong, W.-J. Xie, Z.h. Li, J. Liu, Surface partially
electron-hole in comparison with N-TiO2 nanoparticles. oxidized MoS2 nanosheets as a higher efficient cocatalyst for photocatalytic
hydrogen production, Appl. Surf. Sci. 487 (2019) 734–742, https://doi.org/
The XPS data confirmed the oxygen vacancies and nitrogen
10.1016/j.apsusc.2019.04.258.
interstitial-type doping in prepared N-TiO2 by sol–gel route. Further­ [11] K. Zhang, M. Ma, P. Li, D.H. Wang, J.H. Park, Water splitting progress in tandem
more, the O 1s and N 1S XPS data prove that both nitrogen and oxygen devices: moving photolysis beyond electrolysis, Adv. Energy Mater. 6 (15) (2016)
atoms of ASC-PVP participated in hydrogen bonds with N-Ti-O and/ or 1600602, https://doi.org/10.1002/aenm.201600602.
[12] O. Carp, C.L. Huisman, A. Reller, Photoinduced reactivity of titanium dioxide,
Ti-O-N bonds in N-TiO2 to form a stable hybrid structure. N-TiO2/ASC- Prog. Solid State Chem. 32 (2004) 33–177, https://doi.org/10.1016/j.
PVP exhibited great photocatalytic activity for the naphthalene degra­ progsolidstchem.2004.08.001.
dation (86%) within 100 min under visible light irradiation. More [13] J. Wang, G. Wang, B. Cheng, J. Yu, J. Fan, Sulfur-doped g-C3N4/TiO2 S-scheme
heterojunction photocatalyst for Congo Red photodegradation, Chinese J. Catal. 42
importantly, the N-TiO2/ASC-PVP represents good photocatalytic effi­ (1) (2021) 56–68, https://doi.org/10.1016/S1872-2067(20)63634-8.
ciency for the removal of naphthalene pollutants at very low concen­ [14] Y.K. Hashemi, M. Tavakkoli Yaraki, S. Ghanbari, L. Heidarpoor Saremi, M.
trations from seawater. The reasonable photocatalytic efficiency and H. Givianrad, Photodegradation of organic water pollutants under visible light
using anatase F, N co-doped TiO2/SiO2 nanocomposite: semi-pilot plant
ecofriendly make the N-TiO2/ASC-PVP hybrid nanocomposite a good experiment and density functional theory calculations, Chemosphere 275 (2021)
candidate for photocatalytic applications. 129903, https://doi.org/10.1016/j.chemosphere.2021.129903.
[15] M. Malika, S.S. Sonawane, Statistical modelling for the Ultrasonic
photodegradation of Rhodamine B dye using aqueous based Bi-metal doped TiO2
CRediT authorship contribution statement
supported montmorillonite hybrid nanofluid via RSM, Sustain. Energy Technol.
Assess. 44 (2021) 100980, https://doi.org/10.1016/j.seta.2020.100980.
Elaheh Hosseini Doabi: Conceptualization, Methodology, Formal [16] Y. Wang, X. Liu, L. Guo, L. Shang, S.h. Ge, G. Song, N. Naik, Q. Shao, J. Lin, Z.
h. Guo, Metal organic framework-derived C-doped ZnO/TiO2 nanocomposite
analysis, Investigation, Visualization, Validation, Writing – original
catalysts for enhanced photodegradation of Rhodamine B, J. Colloid Interface Sci.
draft, Writing – review & editing. Fatemeh Elmi: Supervision, Valida­ 599 (2021) 566–576, https://doi.org/10.1016/j.jcis.2021.03.167.
tion, Conceptualization, Methodology, Visualization, Formal analysis, [17] H.V. Bao, N.M. Dat, N.T.H. Giang, D.B. Thinh, L.T. Tai, D.N. Trinh, N.D. Hai, N.A.
Investigation, Writing – review & editing. Maryam Mitra Elmi: Vali­ D. Khoa, L.M. Huong, H.M. Nam, M.T. Phong, N.H. Hieu, Behavior of ZnO-doped
TiO2/rGO nanocomposite for water treatment enhancement, Surf. Interfaces 23
dation, Conceptualization, Methodology, Visualization, Formal analysis, (2021) 100950, https://doi.org/10.1016/j.surfin.2021.100950.
Investigation, Writing – review & editing. [18] F.u. Zhang, C.-L. Zhang, H.-Y. Peng, H.-P. Cong, H.-S. Qian, Near-infrared
photocatalytic upconversion nanoparticles/TiO2 nanofibers assembled in large
scale by electrospinning, Part. Part. Syst. Charact. 33 (5) (2016) 248–253, https://
Declaration of Competing Interest doi.org/10.1002/ppsc.v33.510.1002/ppsc.201600010.
[19] R.M. Mohamed, M.A. Barakat, Enhancement of photocatalytic activity of ZnO/SiO2
The authors declare that they have no known competing financial by nanosized Pt for photocatalytic degradation of phenol in wastewater, Int. J.
Photoenergy 2012 (2012), https://doi.org/10.1155/2012/103672.
interests or personal relationships that could have appeared to influence [20] A.L. Linsebigler, G. Lu, J.T. Yates, Photocatalysis on TiOn surfaces: principles,
the work reported in this paper. mechanisms, and selected results, Chem. Rev. 95 (3) (1995) 735–758, https://doi.
org/10.1021/cr00035a013.

8
E.H. Doabi et al. Journal of Photochemistry & Photobiology, A: Chemistry 425 (2022) 113677

[21] H. Irie, Y. Watanabe, K. Hashimoto, Carbon-doped anatase TiO2 powders as a [41] D. Su, C. Wang, S. Cai, C. Mu, D. Li, W. Lin, Influence of palygorskite on the
Visible-light sensitive photocatalyst, Chem. Lett. 32 (8) (2003) 772–773, https:// structure and thermal stability of collagen, Appl. Clay Sci. 62–63 (2012) 41–46,
doi.org/10.1246/cl.2003.772. https://doi.org/10.1016/j.clay.2012.04.017.
[22] H. Irie, Y. Watanabe, K. Hashimoto, Nitrogen-concentration dependence on [42] S.E. Kumekov, N.K. Saitova, E.O. Syrgaliyev, Migration of optical excited states of
photocatalytic activity of TiO2-xNx powders, J. Phys. Chem. B 107 (3) (2003) the modified chromium complexes of collagen, Mater. Sci. Technol. Creat. New
5483–5486, https://doi.org/10.1021/jp030133h. Mater. 14 (2) (2017) 63–66. http://rep.ksu.kz//handle/data/6016.
[23] T. Umebayashi, T. Yamaki, H. Itoh, K. Asai, Band gap narrowing of titanium [43] S. Livraghi, M.R. Chierotti, E. Giamello, G. Magnacca, M.C. Paganini,
dioxide by sulfur doping, Appl. Phys. Lett. 454 (2002) (2012) 3–6, https://doi.org/ G. Cappelletti, C.L. Bianchi, Nitrogen-doped titanium dioxide active in
10.1063/1.1493647. photocatalytic reactions with visible light: a multi-technique characterization of
[24] S.A. Ansari, M.M. Khan, M.O. Ansari, M.H. Cho, Nitrogen-doped titanium dioxide differently prepared materials, J. Phys. Chem. C. 112 (2008) 17244–17252,
(N-doped TiO2) for visible light photocatalysis, New J. Chem. 40 (4) (2016) https://doi.org/10.1021/jp803806s.
3000–3009, https://doi.org/10.1039/C5NJ03478G. [44] Z. Zhang, X. Wang, J. Long, Q. Gu, Z. Ding, X. Fu, Nitrogen-doped titanium dioxide
[25] R.M. Mohamed, D.L. McKinney, W.M. Sigmund, Enhanced nanocatalysts, Mater. visible light photocatalyst : spectroscopic identification of photoactive centers,
Sci. Eng. R. 73 (1) (2012) 1–13, https://doi.org/10.1016/j.mser.2011.09.001. J. Catal. 276 (2) (2010) 201–214, https://doi.org/10.1016/j.jcat.2010.07.033.
[26] S.H. Mir, L.A. Nagahara, T. Thundat, P. Mokarian-Tabari, H. Furukawa, A. Khosla, [45] S.H. Cheung, P. Nachimuthu, A.G. Joly, M.H. Engelhard, M.K. Bowman, S.
Review—organic-inorganic hybrid functional materials: an integrated platform for A. Chambers, N incorporation and electronic structure in N-doped TiO2 (1 1 0)
applied technologies, J. Electrochem. Soc. 165 (8) (2018) B3137–B3156, https:// rutile, Surf. Sci. 601 (7) (2007) 1754–1762, https://doi.org/10.1016/j.
doi.org/10.1149/2.0191808jes. susc.2007.01.051.
[27] K.M. Koczkur, S. Mourdikoudis, L. Polavarapu, S.E. Skrabalak, [46] M. Delfino, J.A. Fair, D. Hodul, Xray photoemission spectra of reactively sputtered
Polyvinylpyrrolidone (PVP) in nanoparticle synthesis, Dalt. Trans. 44 (41) (2015) TiN X-ray photoemission spectra of reactively sputtered TIN, J. Appl. Phys. 71 (12)
17883–17905, https://doi.org/10.1039/C5DT02964C. (1992) 6079–6085, https://doi.org/10.1063/1.350465.
[28] J. Yuan, M. Chen, J. Shi, W. Shangguan, Preparations and photocatalytic hydrogen [47] B. Viswanathan, K.R. Krishanmurthy, Nitrogen incorporation in TiO2: Does it make
evolution of N-doped TiO2 from urea and titanium tetrachloride, Int. J. Hydrogen a visible light photo-active material ? Int. J. Photoenergy article ID 2012 (2012)
Energy. 31 (10) (2006) 1326–1331, https://doi.org/10.1016/j. 1–10, https://doi.org/10.1155/2012/269654.
ijhydene.2005.11.016. [48] Cristiana Di Valentin, Emanuele Finazzi, Gianfranco Pacchioni, Annabella Selloni,
[29] F. Elmi, M.M. Elmi, F.N. Amiri, Thermodynamic parameters and influence of Stefano Livraghi, Maria Cristina Paganini, Elio Giamello, N-doped TiO2: theory and
kinetic factors on the self-assembly of acid-soluble collagen nanofibrils, Food experiment, Chem. Phys. 339 (1-3) (2007) 44–56, https://doi.org/10.1016/j.
Biophys. 12 (3) (2017) 365–373, https://doi.org/10.1007/s11483-017-9492-5. chemphys.2007.07.020.
[30] J. Gomes, J. Lincho, E. Domingues, R.M. Quinta-Ferreira, R.C. Martins, N-TiO2 [49] Junyang Xian, Qing Hua, Zhiquan Jiang, Yunsheng Ma, Weixin Huang, Size-
photocatalysts: a review of their characteristics and capacity for emerging dependent interaction of the poly(N-vinyl-2-pyrrolidone) capping ligand with Pd
contaminants removal, Water 11 (2) (2019) 373–377, https://doi.org/10.3390/ nanocrystals, Langmuir. 28 (17) (2012) 6736–6741, https://doi.org/10.1021/
w11020373. la300786w.
[31] J. Zhang, P. Zhou, J. Liu, J. Yu, New understanding of the difference of [50] Arghya K. Bishal, Cortino Sukotjo, Christos G. Takoudis, Room temperature TiO2
photocatalytic activity among anatase, rutile and brookite TiO2, Phys. Chem. atomic layer deposition on collagen membrane from a titanium alkylamide
Chem. Phys. 16 (38) (2014) 20382–20386, https://doi.org/10.1039/C4CP02201G. precursor, J. Vac. Sci. Technol. A. 35 (1) (2017) 01B134, https://doi.org/10.1116/
[32] R. Asahi, T. Morikawa, T. Ohwaki, K. Aokiand, Y. Taga, Visible-light photocatalysis 1.4972245.
in nitrogen-doped titanium oxides, Science 293 (5528) (2001) 269–271, https:// [51] Ulrike Diebold, The surface science of titanium dioxide, Surf. Sci. Rep. 48 (5-8)
doi.org/10.1126/science.1061051. (2003) 53–229.
[33] T. Lindgren, J.M. Mwabora, E. Avendaño, J. Jonsson, A. Hoel, C.-G. Granqvist, S.- [52] C.D. Wagner, A.V. Naumkin, A. Kraut-Vass, J.W. Allison, NIST standard reference
E. Lindquist, Photoelectrochemical and optical properties of nitrogen doped database 20, version 3.4 (web version), (Http:/Srdata.Nist.Gov/ Xps/). (2003).
titanium dioxide films prepared by reactive DC magnetron sputtering, J. Phys. [53] I. Bertoti, M. Mohai, J.L. Sullivan, S.O. Saied, Surface characterisation of plasma-
Chem. B. 107 (24) (2003) 5709–5716, https://doi.org/10.1021/jp027345j. nitrided an XPS study titanium: An XPS sttudy, Appl. Surf. Sci. 84 (1995) 357–371,
[34] L.V. Saraf, S.I. Patil, S.B. Ogale, S.R. Sainkar, S.T. Kshirsager, Synthesis of https://doi.org/10.1016/0169-4332(94)00545-1.
nanophase TiO2 by ion beam sputtering and cold condensation technique, Int. J. [54] Cristiana Di Valentin, Gianfranco Pacchioni, Annabella Selloni, Stefano Livraghi,
Mod. Phys. B. 12 (1998) 2635–2647, https://doi.org/10.1142/ Elio Giamello, Characterization of paramagnetic species in N-doped TiO2 powders
S0217979298001538. by EPR spectroscopy and DFT calculations, J. Phys. Chem. B. 109 (23) (2005)
[35] Y. Lei, L.D. Zhang, G.W. Meng, G.H. Li, X.Y. Zhang, Y. Lei, L.D. Zhang, G.W. Meng, 11414–11419, https://doi.org/10.1021/jp051756t.
G.H. Li, X.Y. Zhang, C.H. Liang, W. Chen, Preparation and photoluminescence of [55] Lijie Zhu, Qipeng Lu, Longfeng Lv, Yue Wang, Yufeng Hu, Zhenbo Deng,
highly ordered TiO2 nanowire arrays preparation and photoluminescence of highly Zhidong Lou, Yanbing Hou, Feng Teng, Ligand-free rutile and anatase TiO2
ordered TiO2 nanowire arrays, Appl. Phys. Lett. 78 (2001) 1125–1127, https://doi. nanocrystals as electron extraction layers for high performance inverted polymer
org/10.1063/1.1350959. solar cells, RSC Adv. 7 (33) (2017) 20084–20092, https://doi.org/10.1039/
[36] J. Yu, C. Jimmy, Yu Jiaguo, Ho Wingkei, Effects of F− doping on the photocatalytic C7RA00134G.
activity and microstructures of nanocrystalline TiO2 powders, Chem. Mater. 14 (9) [56] K. Batalović, N. Bundaleski, J. Radaković, N. Abazović, M. Mitrić, R.A. Silva,
(2002) 3808–3816. M. Savić, J. Belošević-Čavor, Z. Rakočević, C.M. Rangel, Modification of N-doped
[37] N. Wang, H. Lin, J.L. Ã, X. Yang, L. Zhang, Photoluminescence of TiO2 : Eu TiO2 photocatalysts using noble metals (Pt, Pd) - a combined XPS and DFT study,
nanotubes prepared by a two-step approach, J. Lumin. 123 (2007) 889–891. doi: Phys. Chem. Chem. Phys. 19 (10) (2017) 7062–7071, https://doi.org/10.1039/
10.1016/j.jlumin.2006.01.318. C7CP00188F.
[38] H.U. Lee, S.C. Lee, S. Choi, B. Son, S.M. Lee, H.J. Kim, J. Lee, Efficient visible-light [57] N.C. Saha, H.G. Tompkins, Titanium nitride oxidation chemistry : An xray
induced photocatalysis on nanoporous nitrogen-doped titanium dioxide catalysts, photoelectron spectroscopy study Titanium nitride oxidation spectroscopy study
Chem. Eng. J. 228 (2013) 756–764, https://doi.org/10.1016/j.cej.2013.05.059. chemistry : an x-ray photoelectron, J. Appl. Phys. 72 (1992) 3072–3079, https://
[39] V.M. Longo, A.T. de Figueiredo, A.B. Campos, J.W.M. Espinosa, A.C. Hernandes, C. doi.org/10.1063/1.351465.
A. Taft, J.R. Sambrano, J.A. Varela, E. Longo, Different origins of green-light [58] E. György, A. Pérez del Pino, P. Serra, J.L. Morenza, Depth profiling
photoluminescence emission in structurally ordered and disordered powders of characterisation of the surface layer obtained by pulsed Nd:YAG laser irradiation of
calcium molybdate, J. Phys. Chem. A. 112 (38) (2008) 8920–8928, https://doi. titanium in nitrogen, Surf. Coatings Technol. 173 (2-3) (2003) 265–270, https://
org/10.1021/jp801587w. doi.org/10.1016/S0257-8972(03)00520-6.
[40] Bradley Steinfeld, Jennifer Scott, Gavin Vilander, Larry Marx, Michael Quirk, [59] Xiaobo Chen, Clemens Burda, Photoelectron spectroscopic Investigation of
Julie Lindberg, Kelly Koerner, The role of lean process improvement in nitrogen-doped titania nanoparticles, J. Phys. Chem. B. 108 (40) (2004)
implementation of evidence-based practices in behavioral health care, J. Behav. 15446–15449, https://doi.org/10.1021/jp0469160.
Heal. Serv. Res. 42 (4) (2015) 504–518, https://doi.org/10.1007/s11414-013- [60] M.J. García-Martínez, L. Canoira, G. Blázquez, I. Da Riva, R. Alcántara, J.F. Llamas,
9386-3. Continuous photodegradation of naphthalene in water catalyzed by TiO2 supported
on glass Raschig rings, Chem. Eng. J. 110 (1-3) (2005) 123–128.

You might also like