Vous êtes sur la page 1sur 10

Applied Catalysis A: General 524 (2016) 163172

Contents lists available at ScienceDirect

Applied Catalysis A: General


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

Efcient photocatalytic degradation of acid orange 7 over N-doped


ordered mesoporous titania on carbon bers under visible-light
irradiation based on three synergistic effects
Youji Li , Ming Li, Peng Xu, Shaohua Tang, Chen Liu
College of Chemistry and Chemical Engineering, Jishou University, Hunan 416000, China

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

Article history: N-Doped ordered mesoporous titania on carbon bers (NOMT/CFs) has been prepared by a liquid-crystal
Received 5 October 2014 template (LCT) method with the aid of supercritical deposition. The NOMT/CFs product has been char-
Received in revised form 18 January 2015 acterized by X-ray diffraction analysis, scanning electron and transmission electron microscopies, X-ray
Accepted 31 January 2015
photoelectron spectroscopy, N2 sorption, UV/Vis diffuse reectance spectrophotometry, and photolumi-
Available online 7 June 2016
nescence. The photoactivity of the obtained product in the degradation of acid orange 7 (AO7) solution
under visible-light irradiation has been evaluated. A photocatalytic mechanism involving three syner-
Keywords:
gistic effects of NOMT/CFs is proposed, whereby the titania acts as an electron/hole generator with its
N-doped ordered mesoporous titania
Carbon ber
ordered mesostructure facilitating electronhole separation, N-doping enhances visible-light adsorption,
Synergy and the CFs serve to concentrate the pollutant around the active sites, favor electron transfer, and facil-
Supercritical deposition itate convenient recycling. Gas chromatography and mass spectrometry have been used to analyze the
Visible-light photoactivity intermediates generated in the conversion of AO7, allowing a tentative decomposition pathway to be
proposed. The effects of catalyst amount, pH, and initial AO7 concentration have been examined as oper-
ational parameters. The photocatalytic reactions follow pseudo-rst-order kinetics and are discussed in
terms of a modied LangmuirHinshelwood model.
2016 Elsevier B.V. All rights reserved.

1. Introduction TiO2 with pollutants is in the range 108 103 s [1]. Thus, the
rapid recombination of electrons and holes is one of the major
Titanium dioxide, a chemically stable, highly efcient, and reasons for the low photocatalytic activity. To solve this short-
nontoxic photocatalyst, has been widely used for water and air coming, some researchers have recently developed pore-structured
purication [13]. However, conventional TiO2 photocatalyst can titania to provide a high content of activated centers for the
only utilize light of wavelengths shorter than 388 nm (UV range) adsorbed pollutants, because the efcient oxidation of organics
due to its wide band gap (Eg 3.2 eV for anatase), which lim- is a surface-orientated process in photocatalysis [11,12]. Since its
its its practical application with solar light [47]. Many efforts rst preparation through a modied solgel process using a phos-
have been made in the last two decades with a view to over- phate surfactant by David and Jackie [13] in 1995, mesoporous TiO2
coming this limitation. The two main approaches have been dye has received increasing attention because of its porous structure.
sensitization [4] and doping with impurities [57]. It has been This property facilitates photocatalytic degradation of contami-
reported that the incorporation of nitrogen into the titania crys- nants such as humic substances, phenolic compounds, pesticides,
tal lattice can shift the activation energy of the TiO2 photocatalyst chlorinated compounds, and dyes [1419]. It has also been found
into the visible light region [6,8]. Thus, researchers have found that a porous support serves to concentrate the pollutant around
that nitrogen-doping technology is important for the application the active sites on mesoporous titania, and is effective in facili-
of visible-light-induced photocatalytic TiO2 [9,10]. On the other tating the separation of electronhole pairs [20]. In addition, the
hand, photo-induced electronhole pairs have a short recombi- separation of TiO2 powder from treated wastewater prior to dis-
nation time of the order of 109 s, whereas the reaction time for charge is a time-consuming and expensive process, which has
limited the development and application of TiO2 photocatalysis in
water treatment. To provide high pollutant concentrations in the
Corresponding author.
vicinity of catalytic centers and to permit convenient separation
E-mail address: bcclyj@163.com (Y. Li).
of the photocatalysts from the catalytic system, many efforts have

http://dx.doi.org/10.1016/j.apcata.2015.01.050
0926-860X/ 2016 Elsevier B.V. All rights reserved.
164 Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172

been made to immobilize photocatalysts on porous supports, such synthesize NOMT/CFs. Simultaneously, another portion of the aged
as activated carbon [21], zeolites [22], ceramics [23], glass bers sol was calcined at 500 C for 1 h to synthesize pure NOMT pow-
[24], optical bers [25], and steel bers [26]. Compared to pow- der for comparison. To further evaluate the synergistic relationship
der and particulate supports, ber supports are more amenable to between NOMT and CFs, and to determine the effect of NOMT on
convenient recycling of composite catalysts [24,26]. Among such the synergy, N-doped nanostructured TiO2 on CFs (NNT/CFs) was
ber supports, carbon bers (CFs) are ideal substrate materials also prepared by deposition in supercritical CO2 . The synthetic pro-
for applications in the photocatalysis of pollutants because they cesses used to obtain the NNT and NNT/CFs were similar to those
are long, thin strand materials composed mostly of carbon atoms, used for NOMT and NOMT/CFs, respectively, but without the LCT.
which endows them with advantages such as light weight, high
tensile strength, manufacturing exibility, heat resistance, and sta-
bility under corrosive conditions [2729]. Although CFs have been
2.3. Characterization of the composite photocatalyst
used as substrates for the preparation of photocatalysts, most cases
have been limited to the immobilization of TiO2 particles without
The prole of the NOMT/CF photocatalyst was observed by
ordered mesostructure or doping [27,28,30].
scanning electron microscopy (SEM) (Hitachi S3400N, Japan).
In this work, with the goals of effectively utilizing visible
High-resolution transmission electron microscopy/selected-area
light, high catalytic performance, and convenient separation of
electron diffraction (HRTEM/SAED) was performed with a JEOL
the photocatalysts from the water treatment system, NOMT with
(JEM 2100F) microscope. Small-angle X-ray scattering (SAXS) mea-
an ordered mesostructure and visible-light photoactivity has been
surements were made on a Nanostar U small-angle X-ray scattering
supported on CFs by a liquid-crystal template (LCT) method with
system (Bruker, Germany) using Cu-K radiation (40 kV, 35 mA).
the aid of supercritical deposition. To explore the possible syn-
Wide-angle X-ray diffraction (WAXRD) patterns were recorded on
ergy between N-doping, the ordered mesostructure, and the CF
a Bruker D4 X-ray diffractometer with Ni-ltered Cu K radia-
support on the photoactivity of TiO2 , the as-prepared photocata-
tion (40 kV, 40 mA). Nitrogen adsorptiondesorption isotherms
lysts have been characterized by various analytical techniques. The
were used to determine the BrunauerEmmettTeller (BET) sur-
performances of these new materials have been tested in the pho-
face area and pore size distribution (ASAP2010, Micromeritics Co.,
tocatalytic degradation of a dye under visible-light irradiation. The
USA) at 77 K. UV/Vis absorption spectra were recorded at 298 K
dye under consideration is acid orange 7 (AO7; C16 H11 N2 O4 SNa),
on a UV/Vis diffuse reectance spectrophotometer (Shimadzu UV-
which is a highly water-soluble, acid orange dye of the anionic
2100). X-ray photoelectron spectroscopy (XPS) data were obtained
monoazo class. It is widely used as a colorant in textiles, leather,
with an ESCALab220i-XL electron spectrometer from VG Scien-
and paper. Hence, the photodegradation of AO7 is important with
tic using 300 W Al K radiation. Photoluminescence (PL) emission
regard to the purication of dye efuents [3].
spectra were measured using a Spex 500 uorescence spectropho-
tometer with 325 nm radiation for excitation. The CF-supported
2. Experimental composite was ignited at 900 C and the TiO2 loading was calculated
in terms of ash weight by thermogravimetric analysis (WCT-2C,
2.1. Reagents Optical Instrument Factory, Beijing).

Raw carbon bers used as supports, prepared by activation


of polyacrylonitrile, were obtained from Henan (surface area:
1280 m2 /g, total pore volume: 0.19 cm3 /g). Tetrabutyl titanate, 2.4. Photocatalytic reactor and light source
cetyltrimethylammonium bromide, hydrochloric acid, urea, and
absolute ethanol were all analytical grade and were purchased from To examine the degradation of AO7 in aqueous solution under
Beijing Zhonglian Chemical Reagent company. The chemical struc- visible light, experiments were carried out using a cylindrical batch
ture of AO7 and its absorption spectrum are given in Supporting reactor tted with a water-cooled quartz jacket and an air-sparging
Information Fig. S1. Deionized water, puried with an Elga-Pure motor. A 500 W high-pressure Hg lamp with a UV lter was placed
water purication system, was used to prepare all solutions for the about 20 cm from the reactor. The photon ow entering the reactor
experiments. was 6 mW cm2 for visible light irradiation.

2.2. Preparation of photocatalysts


2.5. Photocatalytic activity test
NOMT/CFs was prepared by deposition in supercritical CO2 ,
using tetra-n-butyl titanate, a liquid crystal, and urea as the The experimental procedure involved placing 500 mL of a l-
precursor, soft template, and nitrogen source, respectively. First, tered suspension in the photocatalytic reactor. Air was bubbled
cetyltrimethylammonium bromide (5 g) was accurately weighed through the reaction solution at a rate of 560 mL min1 to ensure
and then completely dissolved in distilled water with stirring for a a constant concentration of dissolved oxygen. Thereafter, samples
specied period of time to form the liquid-crystal template (LCT) by of about 5 mL were withdrawn at specied time intervals dur-
a similar methodology to that described elsewhere [18]. Next, tetra- ing illumination and ltered through 0.45 mm lters. The AO7
butyl titanate (25 mL) and urea (2 g) were dissolved in ethanol, and concentration in the reaction mixture was monitored by a TU-
the mixture was magnetically stirred (100 rpm) for 2 h. Ambient 1810 spectrophotometer by measuring the absorption intensity
laboratory temperature was maintained while stirring. A solution at max = 480 nm in a 10 mm quartz cell to evaluate decoloriza-
containing HCl (3.3 mL) and the obtained liquid crystal was added tion. The photocatalysis was tested in triplicate and the results are
dropwise over 1 h until a liquid-crystal sol was obtained. After reported as mean values, with an error of less than 5%. The total
aging the sol for 12 h at room temperature, it was deposited on CFs organic contents (TOC) of the suspensions containing the photo-
in supercritical CO2 (at 7.6 MPa and 50 C). Subsequently, organic catalyst and the corresponding initial solutions of the dyes were
material in the resulting composite was extracted in a soxhlet appa- directly analyzed using a TOC analyzer (model 820, Sievers, USA) to
ratus for 48 h. The composite was dried for 30 min at 100 C in an evaluate mineralization. Dye degradation products were identied
oven and then calcined at 500 C for 1 h in a nitrogen atmosphere to by GCMS (Agilent 7890A5973N).
Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172 165

3. Results and discussion the mesostructure, as described elsewhere [19]. Although the
patterns of NOMT/CFs exhibit narrower and more intense diffrac-
3.1. Characterization of catalysts tion peaks than those of raw CFs, the coexisting TiO2 peaks for
NNT/CFs are more distinct. Enhanced anatase diffraction peaks of
3.1.1. Morphological and textural properties of NOMT/CFs the (101), (004), (200), (105), (211), and (204) planes for NNT/CFs
The morphology of the obtained product was rst examined were observed at 2 = 25.48, 37.98, 48.28, 54.28, 55.38, and 64.37 ,
by SEM and TEM. Compared with the smooth surface of pure CFs respectively (JCPDS No. 211272). Meanwhile, there were obvious
(Fig. 1a), an image of NOMT/CFs (Fig. 1b) revealed that a titania decreases in the intensity of the distinct diffraction peaks of the
layer had been immobilized on almost each ber. Although the CFs for the G(002) (2 = 24.47 ) and G(101) (2 = 44.29 ) planes,
titania layer had been extensively deposited on the bers, the suggesting the presence of numerous titania layers on the sur-
same spatial distribution of CFs was retained in the composite. face of the support. The crystallite sizes, as calculated from the
Therefore, adequate UV irradiation could penetrate the felt-form broadening of the (101) anatase peak by the Scherrer equation,
photocatalyst to a certain depth to reach a three-dimensional envi- were found to be 10.1 nm (NOMT/CFs), 16.3 nm (NOMT), 14.7 nm
ronment for the photocatalytic reaction. This feature differentiates (NNT/CFs), and 19.7 nm (NNT) (Table 1). The crystallite sizes of
the present photocatalyst from current immobilized photocata- NOMT/CFs and NOMT are thus smaller than those of NNT/CFs and
lysts such as TiO2 immobilized on the surfaces of glass [31], steel NNT, respectively. This may be attributed to hindering effects of
plate [32], and ceramic membranes [33], which only provide a carbon adsorption and LCT decomposition on the growth of the
two-dimensional surface for photodegradation. Additionally, the anatase nanocrystals.
CF substrates are uniformly and compactly covered by a large num-
ber of irregular-shaped aggregates with dimensions of 0.51 m 3.1.3. Chemical composition of NOMT/CFs
to form a rough surface, as shown in Fig. 1c. The thickness of the XPS analysis provided valuable insight into the chemical compo-
titania layer is around 200 nm (Fig. 1d). However, the aggrega- sition of the carbon-ber-supported N-doped OMT photocatalyst.
tion size of NOMT in the composites is smaller than that of pure An XPS survey spectrum of the photocatalyst exhibited prominent
NOMT, at about 2 3 m according to the observed morphology peaks of carbon, oxygen, and titanium, but a relatively feeble peak
(Fig. S2). TEM characterization provided additional information due to nitrogen, as shown in Fig. 3a. The O1 s spectrum can be
regarding the interior structure of these architectures. A typical resolved into two peaks, one at 530.6 eV and the other at 529.1 eV
TEM image is shown in Fig. 1e; the NOMT/CFs exhibited a well- (Fig. 3b). The former can be attributed to surface hydroxyl groups
ordered channel arrangement in large domains, corresponding to and the latter to the Ti-O-Ti lattice oxygen of TiO2 [35]. A high-
the SAXS pattern (see discussion below). The inset in Fig. 1e shows resolution XPS spectrum of C1 s on the composite photocatalyst
the SAED pattern, which also displays a well-resolved diffraction is shown in Fig. 3c, which can be tted by three peaks at bind-
ring and many diffraction spots, indicating high crystallinity of ing energies of 281.5, 284.1 and 285.9 eV, respectively. This implies
the pure anatase phase. An HRTEM image (Fig. 1f) shows a 2D that three different chemical environments of carbon existed in
mesostructure, containing ordered mesopores of about 3 nm in the products. The weak peaks at 281.5 and 284.1 eV are usually
diameter, and randomly oriented anatase nanocrystals with well- assigned to the C-Ti and C-C, and the strong peak at 284.6 eV is
dened crystallinity with a lattice spacing of 0.35 nm, consistent ascribed to oxygen-bound species C-O [30,36,37]. Due to surface
with the d-spacing of the (101) reections of pure anatase [14]. oxidation during the preparation process, the oxidized active sites
on the CF surface can contribute to the deposition and growth of
N-doped OMT. The presence of nitride dopant is conrmed by the
3.1.2. Crystalline structure of samples N1 s core level peaks. Two peaks at binding energies of 399.6 and
For NOMT and its composites calcined at 500 C, the ordered 401.6 eV are observed, as shown in Fig. 3d. The rst major peak can
pore structure was characterized on the basis of SAXS measure- be attributed to the nitrogen introduced into the OTiN structure
ments in the 2 range 0.21.5 (Fig. 2a). The SAXS patterns of [38], indicating that some O atoms were substituted by N atoms,
both NOMT and NOMT/CFs show characteristic Bragg peaks [(100) consistent with the visible-light activity of doped TiO2 .
reections] indicative of ordered pores in the samples. NNT and The latter peak can be assigned to the presence of intersti-
its composites do not show any diffraction peaks in the 2 range tial N, being characteristic of NO or NO2 in the N-doped TiO2
0.21.5 in their SAXS patterns because of their extremely large sample [39,40]. The molar ratio of N to Ti in all N-doped sam-
pore dimensions and lack of an ordered mesoporous structure ples was about 2.3% based on XPS analysis. Additionally, nitrogen
[19]. The characteristic (100) reection of pure NOMT shifted to adsorption desorption isotherms and the corresponding pore size
a 2 angle larger than that for NOMT/CFs, indicating a decrease distribution curves were measured to investigate the porous struc-
in the unit-cell parameter and shrinkage of the framework. This ture of the as-prepared samples (see Fig. S3). The obtained samples
change may be attributed to the close contact of the thin NOMT exhibited type IV isotherms characteristic of mesoporosity. The
layers to the CF surfaces, which exhibit hinder effect for titania NOMT/CFs products showed higher surface area (778 m2 /g) than
crystallization. The WAXRD pattern for the NOMT/CFs obtained that of NNT/CFs (195 m2 /g) due to the covering N-doped OMT on the
at a calcination temperature of 500 C indicated only the anatase surface of the CFs. Although the mean pore diameter of NOMT/CFs
phase (Fig. 2b), whereas anatase and rutile phases can coexist in (1.3 nm) was lower than that of NNT/CFs (1.7 nm), the total pore
pure NOMT obtained at the same calcination temperature. Gen- volume of NOMT/CFs (0.08 mL/g) was larger than that of NNT/CFs
erally speaking, the mesoporous TiO2 show low crystal transition (0.04 mL/g) owing to the generation of more pores by the associa-
temperature from the anatase to the rutile possible due to crys- tion of the titania layers with ordered mesoporous structure.
tal size, mesoporous morphology, doping, calcination time and so
on [3]. At the same calcination condition, crystal transition tem- 3.1.4. Optical absorption properties
perature of pure NOMT is lower than composite NOMT/CFs. This The absorption edge and band-gap energies of the as-prepared
can be attributed to the CFs suppressing the phase transformation samples were determined by UV/Vis diffuse reectance spec-
of mesoporous TiO2 from anatase to rutile form at high temper- troscopy. Diffuse reectance spectra of the N-doped samples are
atures. which is good agreent with reports[30,34]. Additionally, illustrated in Fig. 4a. All spectra clearly show a red-shift of the
compared with pure NNT without rutile, the dual-phase nature of absorption edge towards the visible region due to the lower energy
NOMT endows it with low thermal stability due to between the valence band and conduction band of TiO2 . Notably,
166 Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172

Fig. 1. SEM image of carbon bers (a), SEM images of NOMT/CFs at low and high magnications (bd), TEM image of NOMT/CFs at low magnication (e) with selected-area
electron diffraction (SAED) pattern and its high-magnication TEM image with the d-spacing (f).

Fig. 2. SAXS (a) and WAXRD (b) patterns of samples.

the absorption intensities of the composites were higher than that cating that a small-particle quantum size effect was absent in the
of pure N-doped titania at UV and visible wavelengths. This dif- composites. The pure NOMT showed a small shift toward longer
ference can mainly be attributed to the black hue of CFs, and wavelengths in comparison with NNT. This behavior of NOMT is
therefore their strong absorption over the entire range of wave- presumably due to its mesoporous structure and the signicant sur-
lengths employed. The absorption intensities of the NOMT/CFs face effects. The band gaps Eg were calculated using the following
were higher than those of NNT/CFs at UV and visible wavelengths, equation:
which may be attributed to the mesoporous morphology. There
was no shift toward longer wavelengths for the absorption band of
the composites in comparison with that of N-doped titania, indi- 1/2
(Ah) = h Eg (1)
Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172 167

Table 1
An overview of the texture parameters of NOMT and NNT supported on CFs in comparison with those of pure NOMT and NNT and their physical mixtures with CFs.

Samples Content of TiO2 Surface area Total pore Average pore Crystallite size Band-gap energy
(wt.%)a (m2 /g)b volume (cm3 /g)c diameter (nm)c (nm)d (eV)e

NOMT/CFs 6.5 778 0.08 1.3 10.1 2.84


NNT + CFs 6.5 1127 19.7
NOMT 100 137 0.03 2.5 16.3 2. 84
NNT/CFs 6.5 195 0.04 1.7 14.7 2.90
NOMT + CFs 6.5 1263 16.3
NNT 100 61 19.7 2.90
a
TiO2 content was calculated by TG with an error of less than 1%.
b
The BET surface area was determined by the multipoint BET method.
c
The total pore volume and average pore size were calculated by the Barrett Joyner Halenda method (adsorption branch).
d
The average crystallite size of TiO2 was determined by XRD using the Scherrer equation.
e
The band-gap energies were calculated using the equation: (Ahv)1/2 = hv-Eg .

Fig. 3. XPS spectrum of the as-prepared NOMT/CFs (a) and its high-resolution XPS spectra of O1 s (b), C1 s (c), and N1 s (d).

where A is the absorption intensity and h is the photon energy. 3.2. Photocatalysis
Band gaps Eg were obtained from the intersection of the tangents
to the straight portions of plots of (Ahv)1/2 against the energy of 3.2.1. Photocatalytic degradation of AO7 under visible-light
exciting light (hv), as shown in Fig. 4b. Band gaps of the obtained irradiation
samples are listed in Table 1. The anatase form of the N-doped sam- Prior to detailed evaluation of the photocatalytic activities of
ple exhibited a relatively low band gap compared with that of pure the obtained products, it was noted that the concentration of AO7
TiO2 , which has a band-gap energy of 3.2 eV [4]. This observation in aqueous solution decreased even without illumination because
clearly indicates that the TiO2 doped with urea could generate addi- of adsorption. However, the decrease with each catalyst was only
tional energy levels above or below the valence band of TiO2 [35]. 8 26% of the original AO7 concentration (Fig. S4). For comparison,
Although Ti C bonds were formed, as indicated by XPS analysis it can be seen in Fig. 5a that AO7 self-photodegradation was almost
(Fig. 3c), which is benecial for band-gap narrowing as in the case negligible under visible-light irradiation in the absence of any cata-
of carbon-doped TiO2 [10,41], the CF support did not further narrow lyst. Evidently, P25 showed negligible photocatalytic activity in the
the band gap of the as-prepared N-doped TiO2 . degradation of AO7. However, all of the N-doped samples exhibited
good catalytic efciency for AO7 degradation due to the enhanced
168 Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172

Fig. 4. UV/Vis diffuse reectance spectra of samples (a) and plots of (Ahv)1/2 versus photon energy (b).

1.0 Photolysis
b
P25 Visible light
CB
0.8 NNT e- e-
EC O2
UVlight e-

2.8eV
0.6 NOMT
A/A0

NNT-CFs CFs O2
0.4 EV
3.2eV
NOMT-CFs
NOMT, O2p h+
OH--
0.2 a NNT/CFs
EV
Anatase, O2p h+
NOMT/CFs OH
VB
0.0
0 30 60 90 120 150 180 210 240
Time (min)
1.0 3.5
y = 0.0092 x, R2 = 0.98
1st cycle 2nd cycle 3rd cycle 4th cycle 5th cycle
y = 0.0041x, R2= 0.99
d
3.0 NOMT/CFs
0.8 y = 0.0024x , R2= 0.98
c
Ln (TOC0/TOC)

2.5 y = 0.0013x, R2= 0.98


0.6 y = 0.0011x, R2= 0.99
2.0 NNT/CFs
y = 0.0005x, R2 = 0.99
A/A0

0.4 1.5

1.0 NOMT-CFs
0.2 NNT-CFs
0.5 NOMT
NNT
0.0 0.0
0 2 4 6 8 10 12 14 16 18 20 0 70 140 210 280 350 420
Time (hr) Time (min)

Fig. 5. The photocatalytic degradation of AO7 aqueous solution under visible light; its color removal rates (a), catalytic mechanism (b), cycling runs (c), and mineralization
degree (d) over NOMT/CFs; catalyst content = 1 g/L, pH 7, AO7 content = 1 mg/L.

visible-light adsorption. This is consistent with the photocatalytic due to extensive adsorption of the dye molecules on the CFs, facili-
mechanism of N-doped anatase under visible light reported in the tating their photocatalytic decomposition by TiO2 . Additionally, the
literature [9,10]. Although pure NNT showed a certain photocat- photocatalytic activity of NNT-CFs was evidently lower than that
alytic effect (25%) for AO7 degradation, the photocatalytic activity of NOMT-CFs, mainly due to the mesostructured titania of the lat-
(44%) of NOMT was greater than that of NNT because of its porous ter. Furthermore, the photocatalytic activity of the composite was
structure with a great number of activated centers, providing higher than that of a simple mixture of the components because of
effective separation of electronhole pairs [16]. The rate of AO7 the intimate contact between the CFs and NOMT, which is benecial
decolorization was further accelerated by NOMT-CFs and NNT-CFs for electron transfer [42]. The suppression of exciton recombina-
Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172 169

Fig. 6. A possible pathway for the degradation of AO7 in the visible-light photocatalysis system by using NOMT/CFs as catalyst.

tion in the presence of CFs is evidently also an important factor in tively improved. Meanwhile, many dye molecules are adsorbed on
the improved activity of titania besides the high adsorptivity of dye the CF support, leaving them suitably predisposed in the photocat-
molecules, which showed good agreement with their PL intensities alytic environment, thereby enhancing the photocatalytic activity.
(see Fig. S5). On the basis of the above analysis and the relevant band To verify the practical applicability of NOMT/CFs, recycling exper-
positions of pure and N-doped TiO2 , a photocatalytic mechanism iments on the photocatalytic decoloration of AO7 solution were
for the action of NOMT/CFs is proposed, as illustrated in Fig. 5b. As carried out under visible-light irradiation. As shown in Fig. 5c, the
mentioned above, the band gap of pure TiO2 is 3.2 eV, and excita- efciency of photocatalytic decoloration was maintained without
tion can only be achieved with UV light of wavelengths less than any signicant decline after ve cycles. Thus, NOMT/CFs displayed
388 nm. After N-doping, the band gap was obviously decreased and great potential for practical application in wastewater treatment
new impurity levels were introduced between the conduction and using solar light. The excellent reusability of NOMT/CFs may stem
valence bands of TiO2 . Therefore, NOMT/CFs has a narrower band from the good binding property between the NOMT layer and CFs.
gap with a change in the light absorption band from the near-UV to As shown in Fig. S6, two SEM images of NOMT/CFs after ve cycles
the visible-light range. The process of visible-light photocatalytic show no detectable differences in comparison with those of the
oxidation of AO7 is described as follows: corresponding fresh samples (Fig. 1b and c), which is benecial for
the stability of the photocatalytic performance. Additionally, the
NTiO2 + h hVB + + eCB (2) brous structure of the catalyst is very benecial for its separation
O2 + eCB O2 (3) from the treated wastewater. To further examine the photoactivity
of the obtained samples under visible light, the degree of AO7 min-
O2 + H2 O OH + OH (4) eralization was analyzed in the presence of various photocatalysts,
+ OH
as shown in Fig. 5d. Photodegradation of AO7 roughly followed
hVB + OH (5)
pseudo-rst-order kinetics. The rate constants for AO7 degradation
OH + AO7 CO2 + H2 O (6) with NOMT/CFs, NNT/CFs, NOMT-CFs, NNT-CFs, NOMT, and NNT
were found to be 9.2, 4.1, 2.4, 1.3, 1.1, and 0.5 103 min1 , respec-
Firstly, electrons and holes are generated under visible light tively. Among all of these catalysts, NOMT/CFs show the highest
irradiation. The electrons can then react with molecular O2 on the photocatalytic activity for AO7 degradation under visible-light irra-
surface of TiO2 to form a superoxide anion radical (O2 ). This O2 diation. This can be attributed to three synergistic effects between
can then interact with adsorbed H2 O to produce OH radicals, the enhanced visible-light absorption of N-doping, high adsorptivity
main species responsible for the degradation of pollutants, such as of dye molecules, and fast electron transport in CFs and effective
AO7 in the present case. Meanwhile, holes can be captured by the separation of electronhole pairs in OMT.
surface OH to also form OH radicals. From the above equations, it
can be inferred that OMT played an important role in preventing the
recombination of photogenerated electron and hole pairs, because
its ordered mesoporous structure provides a large number of acti- 3.2.2. Possible pathways for conversion of AO7
vated centers. This increases the likelihood of electron-hole pairs In general, for TiO2 under UV irradiation, the oxidative activity
on the surface of the catalyst reacting with adsorbed O2 and H2 O, of TiO2 particles is mostly
respectively, to form O2 and OH. Additionally, CFs also facilitate attributed to species formed through reactions initiated by
channeling of electrons due to their graphite structure [5]. Hence, photogenerated electronhole pairs, such as OH, other peroxyl rad-
the separation rates of photogenerated charge carriers are effec- icals, and valence-band holes [42]. The intermediates generated in
170 Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172

Table 2
Reaction intermediates of AO7 degradation identied by GCMS.

Compound Identied reaction Photocatalytic time Compound Identied reaction Photocatalytic time
intermediates (min) intermediates (min)

sulfanilic acid 25 1,2-naphthalenedione 125

1-amino-2-naphthol 50 ninhydrin 150

p-benzoquinone 75 2-formylbenzoic acid 175

2-naphthalenol 100 hexylacetic acid 200

10

9
Rate constant K103 (min)

5 [AO7] = 1mg/L
4 pH = 7

3
0.5 1 2 3

NOMT/CFs (g/L) Fig. 8. Effect of pH on the degradation rate constant of AO7.

Fig. 7. Effect of NOMT/CFs dosage on the degradation rate constant of AO7.

vation of activated molecules through collision with ground-state


the conversion of AO7 using NOMT/CFs under visible light were molecules. At dosages higher than 1 g L1 , NOMT/CF aggregation
identied by GCMS, as illustrated in Table 2. (particleparticle interactions) may commence, thereby lowering
Additionally, a possible pathway for AO7 photocatalytic degra- the effective surface area of the catalyst and thus the adsorption of
dation by NOMT/CFs is proposed in Fig. 6. A representative GCMS the reactant. Additionally, the turbidity of the mixture is increased,
trace is also presented in Fig. S7. Overall, the photocatalytic degra- which greatly reduces the amount of light transmitted through the
dation of AO7 by NOMT/CFs can be described in terms of a series solution.
of consecutive degradation steps. AO7 is rstly decomposed to aro-
matic intermediates, further oxidized to ring-opened products, and
nally mineralized to CO2 , H2 O, and inorganic salts [43]. 3.2.4. Effects of pH
Fig. 8 shows the variation in rate constant of AO7 mineraliza-
tion with pH. Interestingly, the order of reaction rates was pH
3.2.3. Effects of the NOMT/CFs dosage 7 > pH 9 > pH 5 > pH 11 > pH 3. Strong acid or alkali was not appro-
The reaction rate, which is a function of catalyst content, is an priate for decomposing AO7 because the amount of hydroxyl ion
important parameter [44]. It is well known that the absorption adsorbed on TiO2 is inuenced by the pH [45]. At various pH val-
of light and adsorption of the reactant on the titania surface are ues, the main pathways for producing OH are probably different;
limiting factors in such photocatalysis. For a certain light inten- therefore, the rates of OH production are different. An increase
sity, the amount of photons impinging on the photocatalyst is in pH can be expected to increase the number of OH ions on the
nite. At contents up to about 1 g L1 (or saturated adsorption TiO2 surface, and OH could be formed when OH combines with
of photons), the degradation rate increases with increasing cata- photogenerated holes [Eqs. (5) and (7)] [40]:
lyst content simply because more catalyst provides more activated
centers and can adsorb more of the reactant. However, a further +
hVB
+ + H2 O(ads) OH + H (7)
increase in catalyst dosage beyond 1 g L1 may result in deacti-
Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172 171

Fig. 9. Effect of initial AO7 concentration on kapp and 1/kapp (a) and the ks and kr values of different samples (b).

In acidic solution, the photocatalytic degradation of AO7 is substrate. Therefore, the better dye removal performance may be
probably due to the formation of OH according to the following attributed to the porous structure of the titania, which is consistent
reactions [Eqs. (3), (8)(10)]: with the variation in ks . By addition of CFs, both K and kr of mix-
O tures are obviously increased and are much higher than those of
2(ads)
+ H+ HO2 (8)
pure N-doped titania. This may be due to the fact that the CFs serve
2HO2 O2 + H2 O2 (9) to concentrate the AO7 molecules around the active sites and facil-
itate electron transfer. However, the positive impact of CFs on the
H2 O2 + O
2(ads)
OH + OH + O2 (10) photocatalytic efciency greatly depends on the combined char-
More efcient formation of hydroxyl radicals occurs in alkaline acteristics of CFs and titania. Although ks of composites is lower
solution via the shorter route than that of mixtures, K and kr of the former were higher than
[Eq. (7)]. On the other hand, AO7 has a sulfonate group in its those of the latter because close contact between CFs and titania is
structure, which is negatively charged under alkaline conditions; more benecial to electron transfer and migration of the adsorbed
therefore, AO7 may not be effectively adsorbed on the photocata- AO7 to the active sites of titania. Additionally, the high photoactiv-
lyst surface in alkaline solution [46]. The above ndings indicated ity (K = 9.2 103 min1 , kr = 0.015 mg L1 min1 ) and adsorption
that the most efcient photodegradation of AO7 probably occurred capacity (ks = 4.26 L mg1 ) of NOMT/CFs can be attributed to more
at neutral pH. activated centers, faster electron transfer, and easier migration of
the adsorbed AO7 as a result of the mesostructure and close contact.
3.2.5. Effects of the initial AO7 concentration and kinetics of
photocatalytic mineralization
The efciency of the mineralization process was found to
decrease with increasing concentration of AO7 (Fig. 9). This may
have been a result of blocking of the photocatalytically active
sites [47]. The modied LangmuirHinshelwood kinetic model [37] 4. Conclusions
is generally utilized for such photo-oxidation processes, and the
degradation rate is expressed by the following equation: With the aim of effectively utilizing visible light and conve-
nient application of the photocatalyst in water treatment systems,
1 1 C0
= + (11) NOMT/CFs has been successfully fabricated by a combined method
Kapp kr kS kr of supercritical deposition and liquid-crystal templating. The
where C0 is the initial concentration of the reactant (mg L1 ), kr is NOMT/CF composite showed high photocatalytic activity for AO7
the reaction rate constant (mg L1 min1 ), ks is the adsorption coef- degradation under visible-light irradiation due to three synergis-
cient of the reactant (L mg1 ), and Kapp is the observed apparent tic effects arising from effective separation of electronhole pairs,
pseudo-rst-order rate constant (min1 ). enhanced visible-light adsorption, and high adsorptivity of dye
It was found that the LangmuirHinshelwood kinetic model [Eq. molecules. Meanwhile, the strong binding between the NOMT layer
(11)] adequately describes the and the CFs leads to excellent recyclability of NOMT/CFs in practical
kinetic behavior of AO7 degradation by NOMT/CFs, because a applications for wastewater treatment using solar light. A photo-
linear plot of reciprocal of rate (1/kapp ) against initial AO7 concen- catalytic mechanism for the action of NOMT/CFs and a degradation
tration (C0 ) showed a high R2 of 0.98. The LangmuirHinshelwood pathway for AO7 have been proposed. Conditions of 1 mg L1 AO7,
equation could also be applied for the degradation of AO7 by other pH 7, and 1 g L1 NOMT/CFs were identied as optimal for AO7
catalysts, as shown in Fig. 9b. As indicated in Fig. 9b, all of the sam- degradation. The NOMT/CFs system has good potential for com-
ples showed photocatalytic activity, and the rate constants kr and mercial application. The photocatalytic degradation processes can
absorption equilibrium constants ks determined in this way were be described in terms of a modied LangmuirHinshelwood model
dependent on the nature of the N-doped titania and the synergy for the surface reaction between the dyestuff and the oxidizing
between CFs and titania. Values of kr with NOMT were higher than agent. The values of the adsorption equilibrium constants, ks , and
those with NNT. This can be attributed to the fact that more pores the rate constants, kr , were certainly dependent on the catalyst por-
are formed and more activated centers are located on the external perties. For NOMT/CFs with the highest rate constant, kr and ks were
surface of the catalyst, making them more accessible to the target 0.015 mg L1 min1 and 4.26 L mg1 , respectively.
172 Y. Li et al. / Applied Catalysis A: General 524 (2016) 163172

Acknowledgements [19] K. Zimny, T. Roques-Carmes, C. Carteret, M.J. Stebe, J.L. Blin, J. Phys. Chem. 116
(11) (2012) 65856594.
[20] E.P. Reddy, L. Davydov, P. Smirniotis, Appl Catal. B 42 (1) (2003) 111.
This work was supported by Natural Science Foundation of [21] M.H. Baek, W.C. Jung, J.W. Yoon, J.S. Hong, Y.S. Lee, J. Ind. Engineer Chem. 19
China (21476095,51172092) and Hunan Provincial Natural Science (2) (2013) 469477.
Fund for Distinguished Young (13JJ1023). In addition, we appre- [22] Sanly Liu, May Lim, Rose Amal, Chem. Engineer. Sci. 105 (2014) 4652.
[23] Sittidej Teekateerawej, Junichi Nishino, Yoshio Nosaka, J. Photochem
ciate the constructive suggestions of the anonymous reviewers, Photobiol A 179 (3) (2006) 263268.
which are invaluable in improving the quality of the manuscript. [24] C.H. Lin, J.W. Lee, C.Y. Chang, Y.J. Chang, Y.C. Lee, M.Y. Hwa, Surf. Coat.
Technol. 205 (2010) S341S344.
[25] E. Taboada, I. Angurell, J. Llorca, J. Catal. 309 (2014) 460467.
Appendix A. Supplementary data [26] H.F. Guo, M. Kemell, M. Heikkil, M. Leskel, Appl. Catal. B 95 (34) (2010)
358364.
Supplementary data associated with this article can be found, [27] K.D. Nilay, J. Mol. Catal. A Chem. 337 (1-2) (2011) 3338.
[28] P. Fu, Y. Luan, X. Dai, P.F. Fu, Y. Luan, X.G. Dai, J. Mol. Catal. A 221 (1) (2004)
in the online version, at http://dx.doi.org/10.1016/j.apcata.2015.01.
8188.
050. [29] G.H. Jiang, X. Li, Z. Wei, T.T. Jiang, X.X. Du, W.X. Chen, Powder Technol. 260
(2014) 8489.
[30] J.W. Shi, H.J. Cui, J.W. Chen, M.L. Fu, B. Xu, H.Y. Luo, Z.L. Ye, J. Colloid Interface
References
Sci. 388 (1) (2012) 201208.
[31] E. Aubry, J. Lambert, V. Demange, A. Billard, Surf. Coat. Technol. 206 (23)
[1] S. Mamoru, M. Nicholas, A. Anne, L. Vincent, P. Cesar, B. Oualid, B. Paul, (2012) 49995005.
J.Mater. Res. 28 (3) (2013) 354361. [32] L. Axel, G. Thierry, P. Sbastien, B.R. Elisabeth, Appl. Catal. A 391 (12) (2011)
[2] J. Choina, H. Kosslick, C. Fischer, G.U. Flechsig, L. Frunza, A. Schulz, Appl. Catal. 4351.
B 129 (5) (2013) 589598. [33] Y.Q. Zhu, Q. Xie, F.J. Chen, X.F. Fan, Y.J. Feng, Sci Adv. Mater. 4 (12) (2012)
[3] M. Faycal Atitar, Adel A. Ismail, S.A. Al-Sayari, Detlef Bahnemann, D. Afanasev, 11911199.
A.V. Emeline, Chem. Engineer. J. 264 (2015) 417424. [34] W.X. Guo, F. Zhang, C.J. Lin, Z.L. Wang, Adv. Mater. 24 (35) (2012) 47614764.
[4] M.M. Maitani, C.H. Zhan, D. Mochizuki, E. Suzuki, Y.J. Wada, Appl. Catal. B [35] X.W. Cheng, X.J. Yua, Z.P. Xing, Appl. Surf. Sci. 258 (2012) 32443248.
140141 (2013) 406411. [36] Q. Peng, Y. Li, X. He, H. Lv, P. Hu, Y. Shang, C. Wang, R. Wang, T. Sritharan, S.
[5] P. Chen, L. Gu, X.D. Xue, M.J. Li, X.B. Cao, Chem. Commun. 46 (2010) Du, Compos. Sci. Technol. 74 (2013) 3743.
59065908. [37] L. Gu, J. Wang, H. Cheng, Y. Zhao, L. Liu, X. Han, ACS Appl. Mater. Interfaces 5
[6] N. Pugazhenthiran, S. Murugesan, S. Anandan, J. Hazard. Mater. 263 (2) (2013) (2013) 30853086.
541549. [38] N.C. Saha, H.G. Tompkins, J. Appl. Phys. 72 (1) (1992) 30723079.
[7] F.T. Li, Y. Zhao, Y.J. Hao, X.J. Wang, R.H. Liu, D.S. Zhao, D.M. Chen, J. Hazard. [39] J.M. Du, G.Y. Zhao, Y.F. Shi, H. Yang, Y.X. Li, G.G. Zhu, Y.J. Mao, R.J. Sa, W.M.
Mater 239240 (2012) 118127. Wan, Appl. Surf. Sci. 273 (2013) 278286.
[8] X.W. Cheng, X.J. Yu, Z.P. Xing, Appl. Surf. Sci. 258 (7) (2012) 32443248. [40] T.M. Breault, B.M. Bartlett, J. Phys. Chem. C 116 (2012) 59865994.
[9] S.Z. Hu, F.Y. Li, Z.P. Fan, J. Hazard. Mater. 196 (2011) 248254. [41] M.E. Hassan, L.C. Cong, G.L. Liu, D.W. Zhu, J.B. Cai, Appl. Surf. Sci. 294 (2014)
[10] D. Dolat, N. Quici, E. Kusiak-Nejman, A.W. Morawski, G. Li Puma, Appl. Catal. B 8994.
115116 (2012) 8189. [42] B. Wang, R. Karthikeyan, X.Y. Lu, J. Xuan, M.K.H. Leung, J. Hazard. Mater 263
[11] F.D. Mai, W.L.W. Lee, J.L. Chang, S.C. Liu, C.W. Wu, C.C. Chen, J. Hazard. Mater. (2013) 659669.
177 (13) (2010) 864871. [43] H.Z. Zhao, Y. Sun, L.N. Xu, J.R. Ni, Chemosphere 78 (2010) 4651.
[12] M.R. Bayati, A.Z. Moshfegh, F. Golestani-Fard, Appl. Catal. A 389 (12) (2010) [44] J. Choina, H. Kosslick, C. Fischer, G.U. Flechsig, L. Frunza, A. Schulz, Appl. Catal.
6067. B 129 (2) (2013) 589598.
[13] M.A. David, Y.Y. Jackie, Angew Chem. 107 (18) (1995) 22022206. [45] N. Daneshvar, M.H. Rasoulifard, A.R. Khataee, F. Hosseinzadeh, J. Hazard.
[14] C.X. He, B.Z. Tian, J.L. Zhang, Microporous Mesoporous Mater. 126 (12) Mater. 143 (12) (2007) 95101.
(2009) 5057. [46] A. Troupis, T.M. Triantis, E. Gkika, A. Hiskia, Appl. Catal. B 86 (12) (2009)
[15] J.G. Yu, G.H. Wang, B. Cheng, M.H. Zhou, Appl. Catal. B 69 (34) (2007) 98107.
171180. [47] A. Idris, N. Hassan, R. Rashid, A.F. Ngomsik, J. Hazard. Mater. 186 (1) (2011)
[16] D. S.Kim, S. J.Han, S.Y. Kwak, J. Colloid Interface Sci. 316 (1) (2007) 8591. 629635.
[17] T.C. An, J.K. Liu, G.Y. Li, S.Q. Zhang, H.J. Zhao, X.Y. Zeng, G.Y. Sheng, J. Fu, Appl.
Catal. A 350 (2) (2008) 237243.
[18] Anders E.C. Palmqvist, Curr. Opin Colloid Interface Sci. 8 (2003) 145155.

Vous aimerez peut-être aussi