Accepted Manuscript Adsorption-photocatalytic degradation of dye pollutant in water by graphite oxide grafted titanate nanotubes
Xiaoman Lei, Xiaolong Li, Zhiqiang Ruan, Tingmei Zhang, Fei Pan, Qiang Li, Dongsheng Xia, Jie Fu PII: DOI: Reference:
S0167-7322(18)32297-9 doi:10.1016/j.molliq.2018.06.053 MOLLIQ 9251
To appear in:
Journal of Molecular Liquids
Received date: Revised date: Accepted date:
2 May 2018 7 June 2018 13 June 2018
Please cite this article as: Xiaoman Lei, Xiaolong Li, Zhiqiang Ruan, Tingmei Zhang, Fei Pan, Qiang Li, Dongsheng Xia, Jie Fu , Adsorption-photocatalytic degradation of dye pollutant in water by graphite oxide grafted titanate nanotubes. Molliq (2018), doi:10.1016/j.molliq.2018.06.053
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT Adsorption-photocatalytic degradation of dye pollutant in water by graphite oxide grafted titanate nanotubes
Xiaoman Leia, Xiaolong Lia, Zhiqiang Ruana, Tingmei Zhanga, Fei Pana,b,*, Qiang Lia,
School of Environmental Engineering, Wuhan Textile University, Wuhan, 430073,
SC
a
RI
PT
Dongsheng Xiaa,b, Jie Fua,c,*
P.R. China
Engineering Research Centre for Clean Production of Textile Dyeing and Printing,
NU
b
c
MA
Ministry of Education, Wuhan, 430200, P.R. China
Department of Environmental Science & Engineering, Fudan University, Shanghai
*
PT E
D
200438, P.R. China
Corresponding authors
CE
Jie Fu, Tel. & Fax: 86-21-65647707. E-mail:
[email protected]
AC
Fei Pan, Tel. & Fax: 86-27-59367338. E-mail:
[email protected]
1
ACCEPTED MANUSCRIPT Abstract Aiming at synthesizing titanate-based photocatalysts with high adsorption capacity and enhanced photocatalytic activity for dyes treatment, a new type of graphite oxide grafted titanate nanotubes (TNTs@GO) was prepared through a
PT
one-step hydrothermal method and characterized by a series of techniques including
RI
HRTEM, XRD, FTIR, XPS, etc. At the optimal GO loading (1 wt.%), calcination
SC
temperature (300 °C) and time (0.5 h), TNTs@GO showed the highest photocatalytic activity for degrading Methylene Blue (MB, a model dye) under UV irradiation.
NU
Approximately 97.5% of MB was removed by TNTs@GO with 30-min
MA
pre-adsorption and 90-min UV irradiation. Radical scavengers quenching tests and electron paramagnetic resonance (EPR) measurements demonstrated several reactive
D
species, i.e., h+, •OH and O2•-, played the critical roles in the photocatalytic process,
PT E
and the presence of oxygen vacancy in TNTs@GO reduced the recombination rate of electron-hole pairs, thereby improving the photocatalytic activity. The effects of
CE
operational parameters were investigated, including TNTs@GO dosage, initial MB
AC
concentration, temperature, initial pH, humic acid and inorganic anions. Moreover, TNTs@GO showed a good reusability and applicability for multi-types of dyes. Therefore, TNTs@GO is a promising photocatalyst for dye wastewater treatment. Keywords: TNTs@GO; Photocatalysis; Methylene Blue; Dye wastewater treatment; Reusability
2
ACCEPTED MANUSCRIPT 1. Introduction Dyes are widely used in many industries, such as textile, leather, cosmetic, printing, drug, and food processing [1]. More than 100,000 dyes of various structures including acidic, basic, azo, diazo, anthroquinone based, and metal complex dyes are
PT
commercially available with a production amount of approximately 700,000 tons per
RI
year [2], and the most amount of dyes is manufactured in the developing countries,
SC
e.g., China, India, Thailand, Turkey, etc. [3]. During the application, large quantities of dyes were discharged into the water environment, and it is estimated that
NU
approximate 10% of total spent dyes are eventually discharged into the effluent in the
MA
textile process [4]. Most of dyes contain aromatic ring, which makes them carcinogenic, resistant to biodegradation, highly toxic, and mutagenic to both human
D
beings and aquatic life [5]. For example, some researchers have detected 50-500 ng/L
PT E
of Disperse Red 1 dye in a river influenced by the textile industry discharges; the dye concentration level exceeds the long-term Predicted No-Effect Concentration (PNEC)
CE
for Daphnia similis (60 ng/L) and thereby will pose a potential risk to freshwater biota
AC
[6]. Moreover, the dyes cause undesirable esthetic problem, prevent the solar light penetration and retard the photosynthetic reaction [4]. Many of developing countries do not have strict legislation, or the discharge standard is not mandatorily executed, which makes the dye pollution worse [4]. Therefore, effectively treating such dye pollutants prior to discharge is very important and has attracted increasing attention. Recently, the development of nanotechnology offers leapfrogging opportunities for developing innovative technologies for the more efficient removal of various
3
ACCEPTED MANUSCRIPT pollutants during water and wastewater treatment [7]. The extraordinary properties of nanomaterials, including large surface area, small size effect, quantum effect, photosensitivity, catalytic activity and electrochemical and magnetic properties, greatly benefit their application potentials in dye wastewater treatment [4]. Up to now,
PT
many nanomaterials, such as carbon nanotubes (CNTs) [8], graphene [9], titanium
RI
dioxide (TiO2) nanoparticles [10], and molybdenum disulfide (MoS2) nanosheets [11],
SC
have been used to develop novel adsorbents or photocatalysts for dye removal and showed superior performance than bulk counterparts.
NU
Titanate nanotubes (TNTs), which were firstly synthesized by TiO2 and sodium
MA
hydroxide (NaOH) in 1998 [12], are promising multifunctional nanomaterials in water/wastewater treatments and have drawn great attentions in recent years due to
D
their special physicochemical properties, like uniform microstructures, small tube
PT E
diameter, large surface area, good stability, ion-exchange properties, photo-electric function and quantum size effects [13-15]. Great efforts have been made to apply
CE
TNTs in the removal of various dye pollutants as adsorbents [16-18], and
AC
photocatalysts [19-21]. The adsorption of dyes by TNTs is a chemical sorption process and the electrostatic attraction in the initial bulk diffusion is the critical step [22], resulting in the limited adsorption capacity for basic dyes due to the negatively charged surface of TNTs. Alternatively, the photocatalysis is a more complete, widely applicable, and cost-effective technique, because it can achieve non-selective degradation and mineralization, and has high reusability and prolonged service life. However, the photocatalytic activity of TNTs is relatively limited [23], and thereby
4
ACCEPTED MANUSCRIPT different attempts have been made to improve the photoactivity of TNTs, including post-thermal treatment [24], hydrogen peroxide (H2O2) modification [25], decorating with cerium oxide (CeO2) [19], zinc (Zn) surface-doping [26], nitrogen-doping [27], and hybridizing with stannic oxide (SnO2) [28].
PT
In recent years, graphene, a flat monolayer of carbon atoms tightly packed into a
RI
two-dimensional honeycomb lattice, has attracted a great deal of attentions for its
SC
potential applications in many fields, such as nano-electronics, fuel-cell technology, and catalysts [29, 30]. Graphite oxide (GO) is one of the most important precursors of
NU
graphene, and shares similar sheet structures and properties with graphene, such as
MA
high adsorption capacity and stability, and semiconducting characteristics [31]. GO can enhance the light absorption and photocatalytic activity via suppressing the charge
D
recombination by serving as a photo-generated electron transmitter, when coupled
PT E
with a semiconductor like TiO2 [32]. Yet, few studies have combined GO with TNTs to enhance the photocatalytic activity. Sang et al. [33] used TNTs to decorate GO and
CE
the resultant nanocomposites can simultaneously improve the flame retardancy and
AC
photodegradability of a polymer, flexible polyvinyl chloride (PVC). In the follow-up work [34], they doped platinum into the TNTs/GO nanocomposites and enhanced the visible light photodegradation performance toward PVC. As far as we know, the application of TNTs/GO nanocomposites in the photocatalytic degradation of dyes has not been explored. The information on the effectiveness, feasibility and robustness of this nanomaterial on treating dye wastewater have kept unknown. To this end, we have synthesized a new type of GO grafted TNTs (TNTs@GO)
5
ACCEPTED MANUSCRIPT through hydrothermal method, and optimized the synthesis conditions based on the photocatalytic activity under UV irradiation toward a model dye, Methylene Blue (MB). Besides the material characterization and photocatalytic mechanism explanation, the effects of operational parameters on the adsorption-photocatalytic
PT
performance were comprehensively investigated, including TNTs@GO dosage, initial
RI
MB concentration, temperature, initial pH, humic acid and inorganic anions.
SC
Moreover, the reusability and applicability to different dyes of TNTs@GO were addressed. The overall goal of this work is to develop a promising photocatalyst with
NU
high adsorption capacity and excellent photocatalytic activity for dye wastewater
MA
treatment.
PT E
2.1. Materials and chemicals
D
2. Experimental
The chemicals of this experiment were of analytical grade or higher. Natural
CE
graphite, and 5,5-dimethyl-1-pyrolin-N-oxide (DMPO, >97% purity) were supplied
AC
by Aladdin Industrial Corporation (Shanghai, China). TiO2 (P25, 80% anatase and 20% rutile), the precursor for TNTs synthesis, was purchased from Dedussa of Germany. Methylene Blue, Disperse Blue 56 (DB56), Reactive Brilliant Red X-3B (X-3B), tert-butyl alcohol (TBA), 1,4-benzoquinone (BQ), ethylenediaminetetraacetic acid (EDTA), and all other reagents were obtained from Sinopharm Chemical Reagent Company (Shanghai, China), and used without further purification. The sample solutions throughout the experiments were prepared using deionized (DI) water.
6
ACCEPTED MANUSCRIPT
2.2. Synthesis of TNTs@GO nanocomposites Graphite oxide was prepared using a modified Hummers’ method [35]. Text S1 provides detailed procedures for the preparation of GO. TNTs@GO was prepared
PT
through a one-step hydrothermal method. Specifically, predetermined amount of GO
RI
and 0.6 g of TiO2 was added to 33.35 mL of a 10 M NaOH solution with stirring for
SC
12 h. After mixing completely, the resultant solution was transferred to a 50 mL Teflon-lined flask, and then maintained at 130 °C for 72 h. Next, the sediment was
NU
washed with DI water several times until the pH reached neutral, and then dried at
MA
80 °C in an oven for 4 h. Finally, the product was annealed at 300 °C and lasting 0.5 h. In selected experiments, the GO of predetermined mass was added to achieve
D
different loadings (0.5, 1, 3 and 5 wt.%). At the optimal GO loading, the calcination at
PT E
various temperature (200, 300, 400 and 500 C) were conducted to reveal the optimal calcination temperature. The calcination time (0.5, 1 and 2 h) was also optimized at
CE
the optimal GO loading and calcination temperature. For comparison, TNTs was also
AC
prepared by the similar procedure but without GO loading.
2.3. Material characterization and measurement X-ray powder diffraction (XRD) profiles were collected with a Rigaku Ultima IV diffractometer (Tokyo, Japan) using Cu Kα radiation and operated at 40 kV and 100 mA. Fourier transform infrared spectra (FTIR) characterizations of the samples were obtained with an Avatar 360 FT-IR spectrophotometer (Nicolet, Waltham, MA, USA).
7
ACCEPTED MANUSCRIPT X-ray photoelectron spectroscopy (XPS) measurements were performed ex situ using a Perkin-Elmer PHI system (Waltham, MA, USA). The dried samples were pressed into a clean and high purity indium foil for XPS analysis. The background corrected XPS spectra were fitted by constructing the parameter-controlled Gaussian/Lorentzian
2000).
High
resolution
RI
(http://www.phy.cuhk.edu.hk/~surface/XPSPEAK,
PT
peaks for different chemical states of the element with the XPSPeak software
SC
Transmission electron microscope (HRTEM) images and energy dispersive X-ray spectrometry (EDS) were acquired using a JEOL JEM-2100 TEM (Tokyo, Japan).
NU
The specific surface area and pore volumes of the samples were determined by
MA
nitrogen (N2) adsorption-desorption isotherms on a Quantachrome Autosorb-1 system (Boynton Beach, FL, USA) at 77K. The photo-generated free radicals from the
D
photodegradation system captured by a spin trapping agent DMPO were determined
PT E
by the electron paramagnetic resonance (EPR, Bruker A300, Rheinstetten, Germany), working at a microwave frequency of 9.86 GHz with a power of 1 mW, center field of
AC
CE
3510 G, sweep width of 100 G, and scan number of 3.
2.4. Adsorption-photocatalytic degradation experiment Adsorption-photocatalytic degradation experiments were performed in a glass reactor equipped with magnetic stirring at 25 °C. After 100 mL of dye (MB, RhB or X-3B) solution with the initial concentration of 20 mg/L was introduced into the reactor, the pH was adjusted by diluted NaOH and HCl until to the desired values. Then 20 mg of the catalyst (TNTs, GO or TNTs@GO) was dispersed in the solution
8
ACCEPTED MANUSCRIPT and stirred for 30 min in darkness to reach adsorption-desorption equilibrium. Later, the solution was radiation by a 175 W mercury lamp with the main output at 365 nm and a light intensity of 1.40-1.45 mW/cm2. The samples were withdrawn at pre-determined time intervals, and centrifuged to remove the catalyst. The of
dye
was
then
immediately
determined
by
a
UV-vis
PT
concentration
SC
RI
spectrophotometer (Agilent Technologies, Santa Clara, CA, USA).
2.5. Reusability evaluation
NU
To evaluate the recyclability of TNTs@GO, five successive repeatability
MA
experiments were performed through the following method. After each photocatalytic process, the mixture solution was still for 1 h, allowing the spent TNTs@GO to settle
D
by gravity. Then, about 90% of the supernatant was discarded and the remaining
PT E
liquid-solid mixture was transferred to another glass photoreactor. The reactor was irradiated with UV light for 1 h and washed with methanol, then dried at 80 °C. The
AC
CE
product was subjected to the photocatalysis again.
3. Results and discussion 3.1. Adsorption-photocatalytic performance of TNTs@GO prepared under different conditions The adsorption-photocatalytic performance of TNTs@GO prepared with different GO loadings (0.5-5 wt.%, with a fixed calcination temperature of 300 C and calcination time of 0.5 h), calcination temperatures (200-500 C, with a fixed GO
9
ACCEPTED MANUSCRIPT loading of 1 wt.% and calcination time of 0.5 h), and calcination times (0.5-2 h, with a fixed GO loading of 1 wt.% and calcination temperature of 300 C) were evaluated using MB as a model dye pollutant. GO exhibited a remarkable adsorption capacity for MB, and the removal
PT
efficiency reached 93.2% after 30-min adsorption; after 90-min UV irradiation,
RI
another more 4.4% of MB in the solution were removed by photocatalytic degradation
SC
(Fig. 1(a)). TNTs showed considerable adsorption capacity for MB, and 77.6% of MB can be removed by adsorption; TNTs also had limited photocatalytic activity and 9.3%
NU
of MB was removed by TNTs facilitated photodegradation (Fig. 1(b)). Increasing the
MA
GO loading in TNTs@GO from 0.5 to 1 wt.% gradually decreased the adsorption percentage of MB to 63.6%, however, the photocatalytic removal of MB was
D
significantly increased to 33.9%. Further increasing GO loading to 5 wt.% decreased
PT E
the photocatalytic removal of MB to 5.5%, though the adsorption removal of MB was enhanced. Consequently, the optimum GO loading for TNTs@GO was set at 1 wt.%,
CE
when both photocatalytic removal and total removal of MB were the largest. The
AC
results suggest that the GO loading is of crucial importance for the photocatalytic activity of TNTs@GO. The GO can promote the charge separation in the TNTs@GO, and enhance the photocatalytic activity [36]. However, Excessive GO can cover the active surface of TNTs and hinder its light absorbance, thereby reducing the photocatalytic activity. Increasing the calcination temperature from 200 to 300 C decreased the adsorption capacity of TNTs@GO but enormously promoted the photocatalytic
10
ACCEPTED MANUSCRIPT activity (Fig. 1(c)). The maximum total removal of MB (97.5%) was achieved with a calcination temperature of 300 C. However, continuously increasing the calcination temperature to 400 and 500 C would impact the total removal of MB. Thus, the optimal calcination temperature was determined to be 300 C. Extending the
PT
calcination time from 0.5 to 1 and 2 h showed a modest effect to promote the
RI
adsorption capacity TNTs@GO but an evident effect to decline the photocatalytic
SC
activity (Fig. 1(d)).
Taken together, the TNTs@GO with GO loading of 1%, calcination temperature
NU
of 300 C and calcination time of 0.5 h showed the optimum performance on the
MA
removal of MB, and was further characterized and tested in the following sections.
D
3.2. Characterization of as-synthesized photocatalysts
PT E
Fig. 2 displays the XRD patterns of GO, TNTs and TNTs@GO. For GO, the peak at 10.1° (2θ) corresponds to the (002) interlayer distance of 0.950 nm. Because of the
CE
presence of oxygen functional groups attached on both sides of the graphite flakes, it
AC
creates atomic defects (sp3 bonding) in the graphite structure and tends to exfoliate to a few layers or individual layers of GO in an aqueous medium [37]. The as-synthesized TNTs exhibits five characteristic peaks at 2θ = 9.8°, 24.3°, 28.4°, 48.4° and 62.5°, which can be attributed to sodium tri-titanate and assigned as the (200), (110), (310), (020) and (202) planes of crystalline H2Ti2O5·H2O (JCPDS 47-0124) [38, 39]. The peaks of TNTs@GO are the same as TNTs, and the sample shows no diffraction peaks of GO layered structure. This is due to the layer-stacking or the little
11
ACCEPTED MANUSCRIPT amount of GO-doping or the reduction of GO to graphene [40, 41]. The FTIR spectrum of GO shows several strong peaks (Fig. 3). We can observe the peaks of oxygen functional groups at 1067 cm-1, 1366 cm-1 and 1724 cm-1, which are attributed to C-O, C-O-C and C=O, respectively. The peak at 1620 cm-1 can be
PT
assigned to in-plane vibrations of C=C group in graphene [29]. Furthermore, the wide
RI
absorption ranging from 3038 cm-1 to 3571 cm-1 derives from the O-H stretching
SC
vibration. Compared to GO, the C=O peak at 1724 cm-1 and other sharp peaks at 600-1100 cm-1 are disappeared in the TNTs@GO, which confirms the reduction of
NU
GO.
MA
Based on N2 adsorption-desorption isotherms, the BET surface area was calculated to be 94.4 m2/g and 139.9 m2/g for TNTs and TNTs@GO, and the pore
D
volume was 0.18 cm3/g and 0.244 cm3/g, respectively. The increase of BET surface
PT E
area and pore volume may be owed to that the introduction of GO has resulted in a progressive change on the structure of support. The high BET surface and porous
CE
structure of photocatalysts are beneficial for adsorption and photodegradation of
AC
targeted pollutants. The detailed descriptions for N2 adsorption-desorption isotherms are shown in Text S2 and Fig. S1. The HRTEM was used to investigate the morphology of the synthesized GO and TNTs@GO. The as-prepared GO exhibited a typical layered, flaky degree, and crumpled structure owing to its thin and larger sheetlike morphology (Fig. 4(a-b)). The lateral dimensions of GO nanosheets are several micrometers [42]. For TNTs@GO, we can see that TNTs occupy most of the available surface area of GO,
12
ACCEPTED MANUSCRIPT and an interconnected network of TNTs floated on the surface of the GO (Fig. 4(c-d)), which shows that GO serves as a matrix for the densely packed TNTs. To confirm that the TNTs was successfully distributed on the GO, the EDS was carried out to quantitative analysis for the presence of elements in the materials. The EDS spectrum
PT
shows peaks corresponding to C, O and Ti (Fig. 4(e)), which ensures that the TNTs
RI
are grown on the GO surface.
SC
XPS was investigated to examine the elemental composition and chemical state of the elements that exist within TNTs@GO. It can be seen from the Ti 2p spectrum
NU
that the significant peaks were located in 458.52 eV and 464.17 eV (Fig. 5(a)), which
MA
are the characteristics of Ti 2p3/2 and Ti 2p1/2, respectively [15], and both are derived from Ti4+-O bonds [43]. The O 1s XPS spectrum of TNTs@GO (Fig. 5(b)) contains
D
three dominated peaks at 529.97 eV, 531.22 eV and 535.22 eV, which corresponds to
PT E
O-Ti4+, O-Ti3+, and C-O bonds, respectively. Fig. 5(c) shows the XPS spectrum of C 1s in the TNTs@GO, and the strong peak at 284.72 eV is ascribed to C-C bond.
CE
Additionally, the oxygen functionalities attached to the carbons show deconvoluted
AC
peaks for C-O (285.57 eV) and C=O (288.72 eV) at higher binding energies [44]. The quantitative results and parameters of the XPS fit are listed in Table 1. To further identify the presence of oxygen vacancy (Vo) and Ti3+ in TNTs or TNTs@GO under UV irradiation, EPR was employed. We can see the TNTs has no characteristic EPR resonances (Fig. 6). However, TNTs@GO yields a perfect data fit to the EPR resonance at g = 2.0079, indicating the existence of Ti3+, and the low field signal corresponds to the electrons trapped at the Vo [45]. The reduction of Ti4+ to Ti3+ 13
ACCEPTED MANUSCRIPT owes to the efficient photo-induced interfacial charge transfer. Hence, the recombination rate of electron-hole pairs in TNTs@GO could be reduced, thereby improving the photocatalytic activity for the decomposition of dye pollutants.
PT
3.3. Effects of operational parameters on removal of Methylene Blue
RI
We studied the effect of TNTs@GO dosage on the removal of MB and the
SC
results are shown in Fig. 7. When the dosage of TNTs@GO was raised from 50 to 400 mg/L, the removal efficiency of MB was enhanced significantly; nearly complete
NU
removal of MB was achieved when the TNTs@GO dosage was higher than 200 mg/L.
MA
Higher catalyst dosage probably leaded to more active sites for adsorption and oxidation, and produced more reactive radical species to attack dye pollutants.
D
However, the enchantment effect will be limited when the system is saturated due to
PT E
the shielding of partial UV illumination and self-consumption of radicals by excess catalysts [39, 46]. Since highest MB removal efficiency was got at 200 mg/L, we
CE
chose this material dosage in the following tests for economy and cost consideration.
AC
The MB removal experiments were carried out with different initial MB concentrations (10-40 mg/L), and the results are shown in Fig. 8. The adsorption process was dominant and the photodegradation process can be ignorable with an initial MB concentration of 10 mg/L. As the initial MB concentration was increased from 20 to 40 mg/L, the total MB removal efficiency was gradually decreased from 97.5% to 69.7% after 90-min UV irradiation. The possible reason was that adsorption sites and reactive species approached saturation with an increase in initial MB 14
ACCEPTED MANUSCRIPT concentration [5]. Additionally, higher initial MB concentration introduced more chloride anions (Cl-), which competed with organic substances for reaction with reactive species [47]. In order to ensure a sufficient reaction rate, we selected 20 mg/L as the initial MB concentration in the adsorption-photocatalytic degradation
PT
experiments.
RI
The effect of reaction temperature on photocatalytic degradation of MB by
SC
TNTs@GO was investigated, and the results are displayed in Fig. 9. The calculated pseudo-first-order rate constant (k1) of MB photodegradation was 0.02064 min-1 (R2 =
NU
0.9749) at 15 °C,0.02845 min-1 (R2 = 0.9844) at 25 °C, and 0.02465 min-1 (R2
MA
=0.9861) at 35 °C, respectively. Thus, increasing temperature had a limited promoting effect on the photodegrdation rate of MB, which can be explained by the classical
D
collision theory where the higher temperature accelerates the molecular collision rate,
PT E
and the change in interatomic interactions weakens the chemical bonds, resulting in faster reaction rates [29]. Due to the limited effect and economy consideration, the
CE
photocatalytic experiments can be conducted in room temperature (25 °C).
AC
Chloride (Cl-), sulfate (SO42-), bicarbonate (HCO3-) and phosphate (HPO42-) anions are commonly existed in wastewater. As illustrated in Fig. 10, in the presence of these anions of 200 mg/L, a significant inhibitory effect on MB adsorption and photodegradation was observed, resulting in approximately 31-37% decrease in MB removal after 90-min UV irradiation. When these inorganic anions were introduced into the reaction solution, they were adsorbed on the TNTs@GO surface and competed with MB for limited adsorption sites. What’s more, inorganic anions are 15
ACCEPTED MANUSCRIPT generally regarded as scavengers of photo-generated holes (h+) and/or hydroxyl radicals (•OH), and can pose inhibitory effects through following reactions [39, 48]: Cl- + h+ → •Cl (1)
PT
Cl- + •OH → •ClOH-
RI
(2)
SC
SO42- + •OH → •SO42- + OH-
HCO3- + •OH → •CO3- + H2O
MA
(4)
NU
(3)
HPO42- + •OH → •PO43- + H2O
D
(5)
PT E
The effects of initial solution pH and humic acid (HA) on the removal of MB were also studied. The removal efficiency of MB in the acidic and neutral conditions
CE
(pH 3-7) was slightly higher than that in the alkaline conditions (pH 8-10), and the
AC
neutral condition (pH = 7) provided the highest removal efficiency of MB (Text S3; Fig. S2). The addition of HA (2.5-10 mg/L) promoted the adsorption of MB on TNTs@GO, but inhibited the photodegradation efficiency, which may be due to competition of HA for the light energy and reactive species, and shielding MB molecules from irradiation and reactive species (Text S3; Fig. S3) [29].
3.4. Reusability of TNTs@GO 16
ACCEPTED MANUSCRIPT To evaluate the recyclability of TNTs@GO, five successive repeatability experiments were performed. As shown in Fig. 11, in the first run, 97.5% of MB removal efficiency was achieved within 2 h, and in the second run, the removal of MB was 91.3% within a longer time of 5 h. After five cycles, the photocatalytic efficiency
SC
RI
performance of the material in water treatment application.
PT
of TNTs@GO was still maintained about 86%, indicating a promising reusing
3.5. Mechanistic investigation of photocatalytic degradation of Methylene Blue
NU
In order to clarify the mechanism of photocatalytic degradation of MB by
MA
TNTs@GO under UV irradiation, TBA, BQ and EDTA were added into the reaction system as radical scavengers to capture •OH (reaction rate constant k2 = 6.0×108
D
M-1·s-1 [49]), superoxide radical (O2•-) (k2 = 0.0-1.0×1010 M-1·s-1 [50]) and h+,
PT E
respectively. With the addition of 20 mg/L EDTA, the final removal efficiency of MB was reduced by 20% (Fig. 12(a)), indicating that h+ plays an important role for the
CE
photocatalytic activity of TNTs@GO. Combining the discussion of Fig. 6, the
AC
presence of Vo-Ti3+ can be further confirmed. The photocatalytic activity was also greatly inhibited by the addition of the other two radical scavengers, suggesting that both •OH and O2•- play a significant role in the photocatalytic degradation of MB. Typically, •OH and O2•- are important ROS in the photocatalytic degradation process. •OH is usually generated by the reaction of photo-generated h+ with adsorbed water, whereas O2•- is formed through photo-generated electrons (e-) activating dissolved oxygen [7]. DMPO is commonly used to trap •OH and O2•- in EPR 17
ACCEPTED MANUSCRIPT measurements. As illustrated in Fig. 12(b), a typical four signals of DMPO-•OH adduct with a peak intensity ratio of 1:2:2:1 was observed for TNTs and TNTs@GO. The signals of •OH for TNTs@GO was stronger than that for TNTs, suggesting that TNTs@GO can improve the photocatalytic activity for the degradation of MB. Fig.
PT
12(c) shows the EPR spectra of DMPO-O2•- adduct, and we can see the characteristic
RI
signal peaks of O2•-, indicating that O2•- also plays a very important role in the
SC
photocatalytic degradation of MB.
In summary, the enhanced photocatalytic degradation of MB by TNTs@GO is a
NU
complicated multifunctional process and many kinds of reactive species are involved
MA
such as h+, •OH and O2•-. Fig. 13 presents the mechanism of enhanced photocatalytic ability for TNTs@GO. GO can enhance the photocatalytic activity of TNTs by e-
D
transportation [36]. The following equations present the proposed photocatalytic
PT E
process:
TNTs + hv → TNTs(e- + h+)
CE
(6)
(7)
AC
TNTs(e-) + GO → TNTs + GO(e-)
h+ + H2O → •OH + H+ (8) e- + O2 → O2•(9)
18
ACCEPTED MANUSCRIPT h+ + Pollutants → Degradation products (10) •OH + Pollutants → Degradation products (11)
PT
O2•- + Pollutants → Degradation products
SC
RI
(12)
3.6. Adsorption-photocatalytic removal of different types of dye pollutants
NU
Besides MB (cationic dye), we also tested the removal efficiency of DB56
MA
(neutral dye) and X-3B (anionic dye) by TNTs@GO. The adsorption capacity of TNTs@GO for various dyes is in the order of MB > X-3B > DB56 (Fig. 14). The high
D
adsorption capacity of MB onto TNTs@GO is due to the electric attraction between
PT E
MB cations and negatively charged surface of TNTs@GO. As a consequence, the total dye removal after 90-min UV irradiation is in the same order, and the removal
CE
efficiency reached to 97.5%, 87% and 72%, respectively. Therefore, TNTs@GO has
AC
an excellent adsorption-photocatalytic performance on various types of dyes, and is a promising photocatalyst for treatment of dye wastewater.
4. Conclusions A series of TNTs@GO nanocomposites with different GO loadings, calcination temperatures
and
calcination
times
were
prepared
and
evaluated
their
adsorption-photocatalytic performance on the removal of the model dye MB. The 19
ACCEPTED MANUSCRIPT optimization results indicated the TNTs@GO with GO loading of 1%, calcination temperature of 300 C and calcination time of 0.5 h showed the optimum performance. Thus, the as-prepared TNTs@GO at this condition was further characterized their morphology, crystal phase, functional groups, elemental composition and chemical
PT
state by HRTEM, EDS, XRD, FTIR, XPS, EPR as well as N2 adsorption-desorption
RI
isotherms. For the purpose of practical use, effects of various operational parameters
SC
including TNTs@GO dosage, initial MB concentration, temperature, initial pH, HA, inorganic anions were comprehensively studied. By using the radical scavengers and
NU
EPR measurements, several reactive species, i.e., h+, •OH and O2•-, were demonstrated
MA
to play the critical roles in the photocatalytic process. Overall, TNTs@GO showed an excellent performance on the removal of different types of dyes, a robustness for wide
D
range of pH and temperature, and a good reusability, and therefore has the application
PT E
potential in dye wastewater treatment. However, before the broader engineered
CE
application, the economy and safety of TNTs@GO should be carefully addressed.
AC
Acknowledgements
The authors thank the Collaborative Innovation Plan of Hubei Province for Key Technologies in the Eco-Ramie Industry. This work was financially supported by the Natural Science Foundation of Hubei Province, China (No. 2018CFB515). Dr. Jie Fu thanks the financial support from the Shanghai Pujiang Program (17PJ1400900).
Supplementary materials 20
ACCEPTED MANUSCRIPT Supplementary material associated with this article can be found, in the online version, at doi:.
References
PT
[1] Q.F. Zeng, J. Fu, Y.T. Shi, H.L. Zhu, Degradation of C.I. Disperse Blue 56 by
RI
ultraviolet radiation/sodium hypochlorite, Ozone Sci. Eng., 31 (2009) 37-44.
Bioresour. Technol., 97 (2006) 1061-1085.
SC
[2] G. Crini, Non-conventional low-cost adsorbents for dye removal: a review,
NU
[3] G.T. Güyer, K. Nadeem, N. Dizge, Recycling of pad-batch washing textile
MA
wastewater through advanced oxidation processes and its reusability assessment for Turkish textile industry, J. Clean. Prod., 139 (2016) 488-494.
D
[4] Z. Cai, Y. Sun, W. Liu, F. Pan, P. Sun, J. Fu, An overview of nanomaterials applied
PT E
for removing dyes from wastewater, Environ. Sci. Pollut. Res., 24 (2017) 15882-15904.
CE
[5] Q.F. Zeng, J. Fu, Y. Zhou, Y.T. Shi, H.L. Zhu, Photooxidation degradation of
AC
Reactive Brilliant Red K‐2BP in aqueous solution by ultraviolet radiation/sodium hypochlorite, CLEAN-Soil Air Water, 37 (2009) 574-580. [6] F.I. Vacchi, P.C. Von der Ohe, A.F.d. Albuquerque, J.A.d.S. Vendemiatti, C.C.J. Azevedo, J.G. Honório, B.F.d. Silva, M.V.B. Zanoni, T.B. Henry, A.J. Nogueira, G.d.A. Umbuzeiro, Occurrence and risk assessment of an azo dye – The case of Disperse Red 1, Chemosphere, 156 (2016) 95-100. [7] Z. Cai, A.D. Dwivedi, W.-N. Lee, X. Zhao, W. Liu, M. Sillanpaa, D. Zhao, C.-H. 21
ACCEPTED MANUSCRIPT Huang, J. Fu, Application of nanotechnologies for removing pharmaceutically active compounds from water: development and future trends, Environ. Sci. Nano, 5 (2018) 27-47. [8] J. Ma, F. Yu, L. Zhou, L. Jin, M. Yang, J. Luan, Y. Tang, H. Fan, Z. Yuan, J. Chen,
PT
Enhanced adsorptive removal of methyl orange and methylene blue from aqueous
RI
solution by alkali-activated multiwalled carbon nanotubes, ACS Appl. Mater. Inter., 4
SC
(2012) 5749-5760.
[9] Z.-Y. Sui, Y. Cui, J.-H. Zhu, B.-H. Han, Preparation of three-dimensional graphene
NU
oxide–polyethylenimine porous materials as dye and gas adsorbents, ACS Appl. Mater.
MA
Inter., 5 (2013) 9172-9179.
[10] Q. Wang, R. Jin, C. Yin, M. Wang, J. Wang, S. Gao, Photoelectrocatalytic
D
removal of dye and Cr(VI) pollutants with Ag2S and Bi2S3 co-sensitized TiO2
PT E
nanotube arrays under solar irradiation, Sep. Purif. Technol., 172 (2017) 303-309. [11] A.T. Massey, R. Gusain, S. Kumari, O.P. Khatri, Hierarchical microspheres of
CE
MoS2 nanosheets: Efficient and regenerative adsorbent for removal of water-soluble
AC
dyes, Ind. Eng. Chem. Res., 55 (2016) 7124-7131. [12] T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, K. Niihara, Formation of Titanium Oxide Nanotube, Langmuir, 14 (1998) 3160-3163. [13] D.V. Bavykin, J.M. Friedrich, F.C. Walsh, Protonated titanates and TiO2 nanostructured materials: Synthesis, properties, and applications, Adv. Mater., 18 (2006) 2807-2824. [14] Q. Chen, W. Zhou, G.H. Du, L.M. Peng, Trititanate nanotubes made via a single 22
ACCEPTED MANUSCRIPT alkali treatment, Adv. Mater., 14 (2002) 1208-1211. [15] W. Liu, X. Zhao, T. Wang, J. Fu, J. Ni, Selective and irreversible adsorption of mercury(II) from aqueous solution by a flower-like titanate nanomaterial, J. Mater. Chem. A, 3 (2015) 17676-17684.
PT
[16] C.-K. Lee, S.-S. Liu, L.-C. Juang, C.-C. Wang, M.-D. Lyu, S.-H. Hung,
RI
Application of titanate nanotubes for dyes adsorptive removal from aqueous solution,
SC
J. Hazard. Mater., 148 (2007) 756-760.
[17] L.-C. Juang, C.-K. Lee, C.-C. Wang, S.-H. Hung, M.-D. Lyu, Adsorptive removal
MA
Environ. Eng. Sci., 25 (2008) 519-528.
NU
of Acid Red 1 from aqueous solution with surfactant modified titanate nanotubes,
[18] C.K. Lee, K.S. Lin, C.F. Wu, M.D. Lyu, C.C. Lo, Effects of synthesis
D
temperature on the microstructures and basic dyes adsorption of titanate nanotubes, J.
PT E
Hazard. Mater., 150 (2008) 494-503.
[19] B.C. Viana, O.P. Ferreira, A.G. Souza Filho, C.M. Rodrigues, S.G. Moraes, J.
CE
Mendes Filho, O.L. Alves, Decorating titanate nanotubes with CeO2 nanoparticles, J.
AC
Phys. Chem. C, 113 (2009) 20234-20239. [20] A.Y. Zhang, M.H. Zhou, L. Liu, W. Wang, Y.L. Jiao, Q.X. Zhou, A novel photoelectrocatalytic system for organic contaminant degradation on a TiO2 nanotube (TNT)/Ti electrode, Electrochim. Acta, 55 (2010) 5091-5099. [21] A.Y. Zhang, M.H. Zhou, L. Han, Q.X. Zhou, Combined potential of three catalysis types on TiO2 nanotube (TNT)/Ti and nanoparticle (TNP)/Ti photoelectrodes: A comparative study, Appl. Catal. A-Gen., 385 (2010) 114-122. 23
ACCEPTED MANUSCRIPT [22] L. Xiong, Y. Yang, J.X. Mai, W.L. Sun, C.Y. Zhang, D.P. Wei, Q. Chen, J.R. Ni, Adsorption behavior of methylene blue onto titanate nanotubes, Chem. Eng. J., 156 (2010) 313-320. [23] S. Mozia, E. Borowiak-Palen, J. Przepiorski, B. Grzmil, T. Tsumura, M. Toyoda,
PT
J. Grzechulska-Damszel, A.W. Morawski, Physico-chemical properties and possible
RI
photocatalytic applications of titanate nanotubes synthesized via hydrothermal method,
SC
J. Phys. Chem. Solids, 71 (2010) 263-272.
[24] S. Mozia, Application of temperature modified titanate nanotubes for removal of
NU
an azo dye from water in a hybrid photocatalysis-MD process, Catal. Today, 156
MA
(2010) 198-207.
[25] M.A. Khan, H.-T. Jung, O.B. Yang, Synthesis and characterization of ultrahigh
D
crystalline TiO2 nanotubes, J. Phys. Chem. B, 110 (2006) 6626-6630.
PT E
[26] J.C. Xu, M. Lu, X.Y. Guo, H.L. Li, Zinc ions surface-doped titanium dioxide nanotubes and its photocatalysis activity for degradation of methyl orange in water, J.
CE
Mol. Catal. A: Chem., 226 (2005) 123-127.
AC
[27] X.X. Lin, F. Rong, X. Ji, D.G. Fu, C.W. Yuan, Preparation and enhanced visible light photocatalytic activity of N-doped titanate nanotubes by loaded with Ag for the degradation of X-3B, Solid State Sci., 13 (2011) 1424-1428. [28] C.Y. Tsai, C.W. Liu, C. Fan, H.C. Hsi, T.Y. Chang, Synthesis of a SnO2/TNT heterojunction nanocomposite as a high-performance photocatalyst, J. Phys. Chem. C, 121 (2017) 6050-6059. [29] J. Fu, G.Z. Kyzas, Z. Cai, E.A. Deliyanni, W. Liu, D. Zhao, Photocatalytic 24
ACCEPTED MANUSCRIPT degradation of phenanthrene by graphite oxide-TiO2-Sr(OH)2/SrCO3 nanocomposite under solar irradiation: Effects of water quality parameters and predictive modeling, Chem. Eng. J., 335 (2018) 290-300. [30] Y. Liang, Y. Li, H. Wang, J. Zhou, J. Wang, T. Regier, H. Dai, Co₃O₄
PT
nanocrystals on graphene as a synergistic catalyst for oxygen reduction reaction, Nat.
RI
Mater., 10 (2011) 780-786.
SC
[31] J. Guo, S. Zhu, Z. Chen, Y. Li, Z. Yu, Q. Liu, J. Li, C. Feng, D. Zhang, Sonochemical synthesis of TiO2 nanoparticles on graphene for use as photocatalyst,
NU
Ultrason. Sonochem., 18 (2011) 1082-1090.
MA
[32] P. Zeng, Q. Zhang, X. Zhang, T. Peng, Graphite oxide–TiO2 nanocomposite and its efficient visible-light-driven photocatalytic hydrogen production, J. Alloys Compd.,
D
516 (2012) 85-90.
Graphene
Oxide
PT E
[33] B. Sang, Z.W. Li, X.H. Li, L.G. Yu, Z.J. Zhang, Titanate Nanotubes Decorated Nanocomposites:
Preparation,
Flame
Retardancy,
and
CE
Photodegradation, Nanoscale Res. Lett., 12 (2017) 12.
AC
[34] G.Y. Liu, B. Sang, Z.Q. Zhou, Z.W. Li, Platinum-doped titanate nanotubes/reduced graphene oxide: photocatalytic activity and flame retardancy, Mater. Res. Express, 5 (2018) 9. [35] W.S. Hummers, R.E. Offeman, Preparation of Graphitic Oxide, J. Am. Chem. Soc., 80 (1958) 1339-1339. [36] D. Zhao, G. Sheng, C. Chen, X. Wang, Enhanced photocatalytic degradation of methylene blue under visible irradiation on graphene@TiO2 dyade structure, Appl. 25
ACCEPTED MANUSCRIPT Catal. B, 111-112 (2012) 303-308. [37] J. Shen, B. Yan, M. Shi, H. Ma, N. Li, M. Ye, One step hydrothermal synthesis of TiO2-reduced graphene oxide sheets, J. Mater. Chem., 21 (2011) 3415-3421. [38] X. Sun, Y. Li, Synthesis and characterization of ion-exchangeable titanate
PT
nanotubes, Chemistry, 9 (2003) 2229-2238.
RI
[39] X. Zhao, P. Du, Z. Cai, T. Wang, J. Fu, W. Liu, Photocatalysis of bisphenol A by
SC
an easy-settling titania/titanate composite: Effects of water chemistry factors, degradation pathway and theoretical calculation, Environ. Pollut., 232 (2017).
NU
[40] Y.L. Chen, Z.A. Hu, Y.Q. Chang, H.W. Wang, Z.Y. Zhang, Y.Y. Yang, H.Y. Wu,
MA
Zinc oxide/reduced graphene oxide composites and electrochemical capacitance enhanced by homogeneous incorporation of reduced graphene oxide sheets in zinc
D
oxide matrix, J. Phys. Chem. C, 115 (2011) 2563-2571.
PT E
[41] X.Y. Zhang, H.P. Li, X.L. Cui, Y. Lin, Graphene/TiO2 nanocomposites: synthesis, characterization and application in hydrogen evolution from water photocatalytic
CE
splitting, J. Mater. Chem., 20 (2010) 2801-2806.
AC
[42] D. Zhao, Q. Zhang, H. Xuan, Y. Chen, K. Zhang, S. Feng, A. Alsaedi, T. Hayat, C. Chen, EDTA functionalized Fe3O4/graphene oxide for efficient removal of U(VI) from aqueous solutions, J. Colloid Interface Sci., 506 (2017) 300-307. [43] H. Li, J. Chen, Z. Xia, J. Xing, Microwave-assisted preparation of self-doped TiO2 nanotube arrays for enhanced photoelectrochemical water splitting, J. Mater. Chem. A, 3 (2014) 699-705. [44] J.P. Rourke, P.A. Pandey, J.J. Moore, M. Bates, I.A. Kinloch, R.J. Young, N.R. 26
ACCEPTED MANUSCRIPT Wilson, The real graphene oxide revealed: Stripping the oxidative debris from the graphene‐like sheets, Angew. Chem. Int. Ed., 123 (2011) 3173-3177. [45] T. Berger, M. Sterrer, O. Diwald, E. Knözinger, D. Panayotov, T.L. Thompson, Y.J. Jr, Light-induced charge separation in anatase TiO2 particles, J. Phys. Chem. B,
PT
109 (2005) 6061-6068.
RI
[46] T. Wentien, L. Meikuei, T.Y. Su, Y.M. Chang, Photodegradation of bisphenol-A
SC
in a batch TiO2 suspension reactor, J. Hazard. Mater., 168 (2009) 269-275. [47] C.H. Liao, S.F. Kang, F.A. Wu, Hydroxyl radical scavenging role of chloride and
NU
bicarbonate ions in the H2O2/UV process, Chemosphere, 44 (2001) 1193-1200.
MA
[48] D.S. Bhatkhande, V.G. Pangarkar, A.A.C.M. Beenackers, Photocatalytic degradation for environmental applications – a review, J. Chem. Technol. Biotechnol.,
D
77 (2002) 102-116.
PT E
[49] Z. He, C. Wang, H. Wang, F. Hong, X. Xu, J. Chen, S. Song, Increasing the catalytic activities of iodine doped titanium dioxide by modifying with tin dioxide for
CE
the photodegradation of 2-chlorophenol under visible light irradiation, J. Hazard.
AC
Mater., 189 (2011) 595-602. [50] E.M. Rodríguez, G. Márquez, M. Tena, P.M. Álvarez, F.J. Beltrán, Determination of main species involved in the first steps of TiO2 photocatalytic degradation of organics with the use of scavengers: The case of ofloxacin, Appl. Catal. B, 178 (2015) 44-53.
27
ACCEPTED MANUSCRIPT
Fig. 1. Adsorption-photocatalytic performance of TNTs@GO prepared under different conditions for removal of MB: (a-b) Effect of GO loading, (c) effect of calcination temperature, and (d) effect of calcination time. Experimental conditions: [MB]0 = 20
PT
mg/L, catalysts dosage = 200 mg/L, pH0 = 7.0±0.1, and T = 25 °C. Fig. 2. XRD patterns of GO, TNTs and TNTs@GO.
RI
Fig. 3. FTIR spectra of GO and TNTs@GO.
SC
Fig. 4. (a-b) TEM pictures of GO sheets, (c-d) HRTEM pictures of TNTs@GO, and (e)
NU
EDS analysis of TNTs@GO.
Fig. 5. XPS spectra of the synthesized TNTs@GO: (a) Ti 2p, (b) O 1s, and (c) C 1s.
MA
Fig. 6. EPR spectra of TNTs and TNTs@GO.
Fig. 7. Effect of TNTs@GO dosage on removal of MB. Experimental conditions:
PT E
D
[MB]0 = 20 mg/L, catalysts dosage = 50-400 mg/L, pH0 = 7.0±0.1, and T = 25 °C. Fig. 8. Effect of initial MB concentration on removal of MB. Experimental conditions:
CE
[MB]0 = 10-40 mg/L, catalysts dosage = 200 mg/L, pH0 = 7.0±0.1, and T = 25 °C. Fig. 9. Effect of temperature on photocatalytic degradation of MB. Experimental
AC
conditions: [MB]0 = 20 mg/L, catalysts dosage = 200 mg/L, pH0 = 7.0±0.1, and T = 15-35 °C. The inset is the pseudo-first-order kinetic model fitting. Fig. 10. Effect of inorganic anions on removal of MB. Experimental conditions: [MB]0 = 20 mg/L, catalysts dosage = 200 mg/L, pH0 = 7.0±0.1, T = 25 °C, and [anion]0 = 200 mg/L.
28
ACCEPTED MANUSCRIPT Fig. 11. Recycling experiments of TNTs@GO for photocatalytic degradation of MB. Experimental conditions: [MB]0 = 20 mg/L, catalysts dosage = 200 mg/L, pH0 = 7.0±0.1, and T = 25 °C. Fig. 12. (a) Effects of TBA, BQ and EDTA on photocatalytic degradation of MB. (b)
PT
EPR spectra of DMPO-•OH adduct with water dispersion. (c) EPR spectra of
RI
DMPO-O2•- adduct with methanol dispersion. Experimental conditions: [MB]0 = 20
SC
mg/L, catalysts dosage = 200 mg/L, pH0 = 7.0±0.1, T = 25 °C, [scavenger]0 = 20 mg/L, and DMPO = 5 mM.
NU
Fig. 13. Proposed mechanisms for the enhanced photocatalytic activity of TNTs@GO
MA
under UV irradiation.
Fig. 14. Adsorption-photocatalytic performance of TNTs@GO for removal of
D
different dyes. Experimental conditions: [dye]0 = 20 mg/L, catalysts dosage = 200
AC
CE
PT E
mg/L, pH0 = 7.0±0.1, and T = 25 °C.
29
ACCEPTED MANUSCRIPT
Table 1. XPS curve fitting results and parameters for Ti 2p, C 1s and O 1s spectra of TNTs@GO Ti 2p
C 1s
O 1s
Ti 2p1/2 C-C/C-H
C-O
C=O
O-Ti4+
O-Ti3+ C-O bonds
Position
458.52
464.17
284.72
285.57
288.72
529.97
531.22
FWHM
1.216
2.151
1.132
2.467
1.811
1.296
2.818
2.29
Area
164502
81102
28820
16392
4444
193092
63291
17038
535.22
AC
CE
PT E
D
MA
NU
SC
RI
PT
Ti 2p3/2
30
ACCEPTED MANUSCRIPT Highlights
Graphite oxide grafted titanate nanotubes was prepared through hydrothermal method. Synthesis condition was optimized to achieve the excellent photocatalytic activity.
Oxygen vacancy in TNTs@GO reduced the recombination rate of electron-hole
PT
RI
pairs.
The effects of operational parameters were comprehensively investigated.
TNTs@GO showed a good reusability and applicability for multi-types of dyes.
AC
CE
PT E
D
MA
NU
SC
31
Graphics Abstract
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13
Figure 14