Next Article in Journal
SO42−/Sn-MMT Solid Acid Catalyst for Xylose and Xylan Conversion into Furfural in the Biphasic System
Next Article in Special Issue
Photocatalytic Graphene-TiO2 Thin Films Fabricated by Low-Temperature Ultrasonic Vibration-Assisted Spin and Spray Coating in a Sol-Gel Process
Previous Article in Journal
Hydrothermal Fabrication of High Specific Surface Area Mesoporous MgO with Excellent CO2 Adsorption Potential at Intermediate Temperatures
Previous Article in Special Issue
Photocatalytic TiO2 Nanorod Spheres and Arrays Compatible with Flexible Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Synthesis of Plasmonic Ag/AgCl@TiO2 Continues Fibers with Enhanced Broadband Photocatalytic Performance

1
Shandong Key Laboratory of Water Pollution Control and Source Reuse, School of Environmental Science and Engineering, Shandong University, Jinan 250100, China
2
Institut National de la Recherche Scientifique (INRS), Centre Énergie Materiaux et Télécommunications, Université du Québec, 1650 Boulevard Lionel-Boulet, Varennes, Québec, QC J3X 1S2, Canada
*
Author to whom correspondence should be addressed.
Catalysts 2017, 7(4), 117; https://doi.org/10.3390/catal7040117
Submission received: 28 February 2017 / Revised: 2 April 2017 / Accepted: 13 April 2017 / Published: 17 April 2017
(This article belongs to the Special Issue Titanium Dioxide Photocatalysis)

Abstract

:
The plasmonic Ag/AgCl@TiO2 fiber (S-CTF) photocatalyst was synthesized by a two-step approach, including the sol-gel and force spinning method for the preparation of TiO2 fibers (TF), and the impregnation-precipitation-photoreduction strategy for the deposition of Ag/AgCl onto the fibers. NaOH aqueous solution was utilized to hydrolyze TiCl4, to synthesize TF and remove the byproduct HCl, and the produced NaCl was recycled for the formation and deposition of Ag/AgCl. The surface morphology, specific surface area, textural properties, crystal structure, elemental compositions and optical absorption of S-CTF were characterized by a series of instruments. These results revealed that the AgCl and Ag0 species were deposited onto TF successfully, and the obtained S-CTF showed improved visible light absorption due to the surface plasmon resonance of Ag0. In the photocatalytic degradation of X-3B, S-CTF exhibited significantly enhanced activities under separate visible or UV light irradiation, in comparison to TF.

1. Introduction

As a potential technology, photocatalytic oxidation is widely applied in water treatment [1,2] and energy conversion area [3,4]. TiO2 is considered to be one of the most promising semiconductor photocatalysts due to its low cost, chemical stability and non-toxicity [5]. However, its large band gap (3.2 eV) and high recombination rate of photo-generation carrier [6] limit its applications as an efficient and visible-light driven photocatalyst. Numerous strategies have been explored to enhance its activity, especially doping with nonmetal elements (N and S, etc.) [7,8], coupling with other semiconductors (GO and BiVO4, etc.) [9,10], and depositing with noble metals (Au and Ag, etc.) [11,12,13,14]. Ag0 nanoparticles-modified TiO2 exhibits excellent activity under visible light resulting from the surface plasmon resonance (SPR) [15,16], which is induced by the nano-size and morphology of Ag0 [17,18].
Kakuta et al. [19] observed that Ag0 species were formed on the surface of AgBr under ultraviolet (UV) irradiation. Afterward, Huang et al. [20] found that the modification of TiO2 with Ag/AgCl obviously improved its photocatalytic activity. However, previous research has mainly focused on the fabrication of nano-sized Ag/AgCl/TiO2 catalysts, such as nanoparticles [21,22], nanospheres [23], nanoarrays [20,24] and nanofibers [25]. These nano-sized catalysts are difficult to separate directly from post-slurry for recycling. Therefore, it is crucial to develop an easily recycled catalyst with a macroscopic shape and self-supporting structure. TiO2 continuous fibers, fabricated be our group using sol-gel and force spinning method, is a promising strategy.
In addition, for most studies on Ag/AgCl-deposited TiO2, HCl [26] or cetyltrimethylammonium chloride (CTAC) [27] were employed as chlorine sources, which were costly, toxic and harmful to environment. Hence, it is necessary to explore low-cost and non-toxic alternative sources. The cheap and extensive TiCl4 can be employed as a Ti source for TiO2 fabrication. However, its hydrolytic process is accompanied with the formation of hazardous HCl [28,29]. For this purpose, Xu et al. [30] reported a method in which NaOH was used to neutralize the produced HCl and form NaCl during the hydrolysis of TiCl4. This significant improvement allows us to recycle the produced NaCl, and also invites the investigation of the feasibility of using the NaCl as Cl source for Ag/AgCl deposition.
In this study, TiO2 fibers (TF) with a macroscopic shape were fabricated by a sol-gel method according to the previously developed TiO2 continuous fibers [31], and Ag/AgCl nanoparticles were deposited on TF by an impregnation-precipitation-photoreduction method [20]. TiCl4 served as the Ti source for TF, and NaOH aqueous solution was utilized to hydrolyze TiCl4, neutralize the produced HCl and form NaCl. The recycled NaCl was employed as a chlorine source for Ag/AgCl deposition. The morphology, textural properties, chemical component and optical property of Ag/AgCl@TiO2 fibers (S-CTF) were analyzed by scanning electron microscopy (SEM), N2 absorption and desorption isotherms, X-ray diffractometer (XRD), X-ray photoelectron spectroscopy (XPS) and UV-vis spectrophotometer (UV-vis DRS). The photocatalytic activity and stability were evaluated by brilliant red X-3B degradation under visible and UV light.

2. Results and Discussion

2.1. Morphology and Textural Properties

2.1.1. SEM

Figure 1 presents typical SEM images of TF, AgCl@TiO2 fibers (CTF) and S-CTF. The diameters of the fibers are about 30 μm. In Figure 1a,c,e, the surface of TF is homogeneous and smooth, while the surface of CTF and S-CTF are rough. Local enlarged images of TF, CTF and S-CTF are shown in Figure 1b,d,f, in which the porous structure that formed by the agglomeration of TiO2 nanoparticles can be observed. AgCl particles with diameters less than 300 nm (Figure 1d) were successfully deposited on TF. However, there is no significant difference between the SEM images of CTF and S-CTF. The existence of Ag0 cannot be observed through SEM, but will be confirmed by XRD, XPS and UV-vis DRS.

2.1.2. Textural Properties

The textural parameters of TF and S-CTF are presented in Table 1. After heat treatment, the Brunauer-Emmett-Teller (BET) surface area of TF increases from 16.0 of precursor fibers (PF) to 27.9 m2 g−1 (TF), due to the removal of organic matters and the formation of porous structure. With Ag/AgCl particles deposited on TF, the surface area of S-CTF increases to 30.6 m2 g−1, which is due to its rough surface [32], as shown in Figure 1.
N2 absorption and desorption isotherms of TF and S-CTF are shown in Figure 2a. Isotherms of TF and S-CTF belong to typical IV adsorption behavior and Type H1 hysteresis loop, according to International Union of Pure and Applied Chemistry (IUPAC) classification [33], which is consistent with pores with homogeneous and spindle cylindrical shapes. This reveals the formation of a mesoporous structure by the heat treatment. According to the pore-size distribution isotherms (Figure 2b), S-CTF shows narrower pore-size distribution (3–22 nm) and smaller average pore size (10.8 nm) than TF (5–40 nm and 16.7 nm). Hence, S-CTF (0.111 cm3 g−1) possesses smaller pore volume compared to TF (0.074 cm3 g−1). The decrease of average pore size and volume results from the deposition of nano-sized Ag/AgCl particles into macro-sized pores. In addition, S-CTF shows bimodal porous structure with two pore-size distribution peaks at 3.80 and 10.8 nm, which has also resulted from the pore size narrowing by AgCl deposition. The enlarged surface area and narrowed bimodal pore-size distribution obtained by S-CTF are both benefited to dye adsorption [34,35].

2.2. XRD

Figure 3a shows XRD patterns of NaCl produced in the neutral reaction of NaOH and HCl. According to Joint Committee on Powder Diffraction Standards (JCPDS) standards, its crystal form is confirmed to be halite NaCl by diffraction peaks at 27.4°, 31.7°, 45.4°, 53.7°, 56.4°, 66.2° and 57.3°. XRD patterns of AgCl, Ag/AgCl, TF, CTF and S-CTF are shown in Figure 3b. Diffraction peaks of TF, Ag@TiO2(CTF) and S-CTF at about 25.3°, 37.8°, 48.1°, 53.9°, 55.1°, 62.7°, 68.7°, 70.3° and 75.1° belong to (101), (004), (200), (105), (211), (204), (116), (220) and (215) reflections [36] of anatase TiO2 (JCPDS No. 21-1272), while diffraction peaks of AgCl, Ag/AgCl, CTF and S-CTF at about 27.9°, 32.3°, 46.3°, 54.9°, 57.5°, 67.5° and 76.7° are in accordance with the (111), (200), (220), (311), (222), (400) and (420) planes [37] of cubic phase AgCl (JCPDS No. 31-1238). A weak diffraction peak at 2 θ   =   38 . 3 °   of cubic phase Ag (JCPDS No. 65-2871) in Ag/AgCl confirmed the successful photoreduction of AgCl and the formation of Ag0. However, no obvious diffraction peak of Ag0 occurred in S-CTF, which is due to the peak overlap with anatase (at about 2 θ   =   37 . 8 ° ) and its high dispersion. Otherwise, anatase peaks of CTF and S-CTF show about 0.45 and 0.2 shifts compared with TF (inset in Figure 3b), which is mainly caused by the combination of AgCl and the formation of Ag0. The shift of anatase peaks towards higher angles indicate a lattice expansion [38] due to the AgCl and Ag0 nanoparticles in the interlayer space of anatase gains agglomerated fibers.

2.3. XPS

The XPS survey spectrum of S-CTF (Figure 4a) shows all peaks of Ti, O, Ag, Cl and C elements. The peak for C 1s at 284.6 eV is mainly from the adventitious hydrocarbon, due to the XPS instrument itself. The result of the survey of the XPS spectrum is consistent with the XRD results.
The valence state of Ti, O, Cl and Ag in S-CTF are investigated by the high-resolution XPS spectrum (Figure 4b–e). As shown in Figure 4b, the binding energy at 458.6 and 464.6 eV of Ti 2p1/2 and Ti 2p3/2 are lower than those of pure TiO2 (459.4 and 464.7 eV) [39]. The curve resolution of the Ti 2p signal of S-CTF indicates four different peaks. The peaks at binding energies of 458.2 and 463.9 eV belong to Ti 2p1/2 and Ti 2p3/2 of Ti3+, while the other two peaks at 458.7 and 464.4 eV belong to Ti 2p1/2 and Ti 2p3/2 of Ti4+ [40]. Ti3+ (and oxygen vacancy) in S-CTF was produced during the heat treatment by the reduction of TF (the removal of partial lattice oxygen) with organic matters in steam as a reducing atmosphere. The O 1s region is decomposed into two peaks at 530 and 531.9 eV (Figure 4b). The two kinds of oxygen contributions are ascribed to Ti–O in TiO2 and the OH in Ti–OH (adsorbed water on the surface of S-CTF) [35]. In Figure 4c, two signal peaks of Cl 2p1/2 and Cl 2p3/2 at binding energy of 197.1 and 198.7 eV with a separation of 1.6 eV confirmed the Cl from AgCl [41]. The Ag 3d region is decomposed into four peaks (Figure 4e). Binding energy peaks at 367.6 and 373.6 eV are ascribed to Ag0, while 367.1 and 373.1 eV correspond to Ag+ in AgCl particles [26,42]. The difference of binding energy (0.1 eV) between CTF and S-CTF also indicates the formation of Ag0. According to the results of XPS, the detail atomic percentages of Ag, AgCl, and TiO2 on the surface of S-CTF were about 8.5%, 40.7% and 51.8%, respectively. Moreover, the interaction among Ag, AgCl and TiO2 [43] is further confirmed by the binding energy shifts peaks for all elements compared with standard values.
The crystal structure and chemical component of S-CTF are investigated by XRD and XPS analysis. The formation of Ag0 and Ti3+ would enhance the optical property of S-CTF.

2.4. UV-Vis DRS

Compared with TF, S-CTF exhibits a slight enhancement in the UV light region (<400 nm) and considerable enhancement in the visible light region (>400 nm) (Figure 5). CTF shows the highest absorption in the UV light region among the samples, even higher than S-CTF. With the photoreduction of AgCl and the formation of Ag0, the absorption of UV light decreases but is still higher than TF. Therefore, the enhancement of absorption in the UV light region can be ascribed to the combination of TF and AgCl with excellent UV light absorption. Because of the formation of Ti3+ and oxygen vacancy, TF possesses localized states below the conduct bind(CB) minimum and shows expected light absorption in the visible light region [39,44,45]. TF shows an absorption band edge at about 400 nm, and the absorption band edge of CTF and S-CTF (at about 420 nm) shows an expected red-shift, in comparison to TF. Compared with Ag/AgCl, S-CTF almost retains the same absorption intensity in the visible light region, illustrating that the composition of TF and Ag/AgCl has no influence on the light absorption of Ag/AgCl. Obviously, an additional absorption for S-CTF which was observed in the region of 400–600 nm, which is attributed to the SPR of Ag0 nanoparticles [41]. The broad and enhanced absorption in the visible light region is also confirmed the formation of Ag0 [46,47]. These evidences of Ag0 formation are consistent with the results of XPS.

2.5. Photocatalytic Activity

To evaluate the effect of the interaction between Ag/AgCl and the TF surface, Ag/AgCl/precursor fibers (S-CPF) were prepared. To investigate the influence of photosensitization and to confirm the photoactivity of TF, pure TiO2 fibers (pure-TF) were also prepared by PF through heat treatment (the heating program was the same as that used for TF preparation) in O2 atmosphere with the flow rate at 0.5 L/min. The photocatalytic activities of pure-TF, TF, CTF, S-CPF and S-CTF toward X-3B degradation under visible and UV light are shown in Figure 6.
After dark treatment for 30 min (Figure 6), S-CTF and CTF (about 12%) have higher X-3B adsorption than pure-TF and TF (about 4%). Nearly no adsorption is observed for S-CPF. The enhancement for X-3B adsorption is a result of the enlarged surface area and narrowed bimodal mesoporous structure after AgCl deposition.
The degradation efficiencies within 120 min under visible light (Figure 6a) follow an order of pure TF   ( 25 . 6 % )   <   TF   ( 37 . 1 % )   < CT F   ( 73 . 4 % )   < S CPF   ( 91 . 8 % )   < S CTF   ( 98 . 1 % ) . After 120 min, the degradation efficiency by pure-TF was 25.6%, which was lower than TF (37.1%). The degradation of X-3B by pure-TF can be attributed to the photosensitization of dye, and the degradation elevation of TF can be attributed to the existence of Ti3+ and the optical absorption of visible light. CTF shows high photoactivity due to the reduction of AgCl and the formation Ag0 as the photoreaction progressed [32]. Because of the strong SPR of Ag0, S-CPF and S-CTF exhibit much higher degradation efficiencies than TF and CTF. The highest efficiency by S-CTF is ascribed to the better separation capacity of photo-generated carriers than S-CPF.
Under UV light, the degradation efficiencies of X-3B (Figure 6b) for 120 min are displayed as following sequence: pure TF   ( 38 . 6 % )   <   TF   ( 44 . 8 % )   < CT F   ( 84 . 5 % )   < S CPF   ( 85 . 7 % )   < S CTF   ( 95 . 1 % ) . The higher efficiency of TF than pure-TF can be attributed to the more efficient carrier separation with oxygen vacancy in TF. CTF and S-CPF exhibit much higher activity than TF. Compared with X-3B degradation efficiency under visible light, the efficiency elevation by CTF results from the excitement of AgCl and the higher activity of TF. Since Ag0 possesses the capacity to trap electrons [48], S-CTF shows the highest degradation efficiency.
The stability of S-CTF is further measured by a recycling experiment. After five runs, the degradation efficiencies of S-CTF under visible and UV light are still as high as 90.8% and 87.5%. The catalyst is stable enough, even considering the inevitable loss of weight during each cycle.

2.6. Photocatalytic Mechanism

2.6.1. Photoluminescence (PL)

The recombination of the photoinduced electrons and holes is a crucial factor in a photocatalytic reaction for the decrease of quantum yield. The efficiencies of photo-generated carriers [12,49,50] are explained by PL analysis. A strong emission peak of TF is observed at about 400 nm with the excitation wavelength at 325 nm (Figure 7), which indicates the direct recombination of electrons and holes. No obvious emission peak is observed in PL spectra of Ag/AgCl and S-CTF, while S-CTF shows lower PL intensity than Ag/AgCl, indicating its better separation capacity caused by the interaction among Ag, AgCl and TiO2. The results of PL analysis are in accordance with the results of XRD and XPS. Otherwise, an additional peak of TF (at about 470 nm), belonging to oxygen vacancy [51], further confirms the formation of oxygen vacancy and Ti3+.

2.6.2. Photocatalytic Mechanism

To better understand the photocatalytic mechanisms of S-CTF and to investigate the main oxidative species during the photodegradation of X-3B, tert-butyl alcohol (TBA), AgNO3, disodium ethylenediaminetetraacetate (EDTA-Na2) and benzoquinone (BQ) are used as •OH, electrons, holes and •O2 quenchers. Under visible light irradiation, the degradation efficiency of X-3B decreases from 98.0% to 87.2% (TBA), 74.6% (AgNO3), 43.9% (EDTA-Na2) and 36.7% (BQ). Under UV light irradiation, the degradation efficiency of X-3B reduces from 95.1% to 85.5% (TBA), 82.2% (AgNO3), 65.5% (EDTA-Na2) and 40.3% (BQ). These results illustrate that the critical species under visible light are holes and •O2, while the most critical specie under UV light is •O2.
The photocatalytic mechanism of S-CTF under visible light is shown in Figure 8a. Ag0 exhibits strong SPR and generates electron-hole pairs under visible irradiation [52]. TF with the existence of Ti3+ is also excited under visible light. With the efficient interaction of Ag0, AgCl and TF, the plasmon-excited electrons are injected into the conduction band (CB) of TF [20,45] easily, and subsequently produce •O2 with O2. Meanwhile, the photo-excited holes are transferred to the surface of AgCl, X-3B molecules [20,32] and the value band (VB) of TF. The efficient separation of carriers among Ag, AgCl and TF is confirmed by PL spectra. Holes participate in the degradation of organic dyestuff by formation of Cl0 and •OH. Firstly, the holes transfer to AgCl and convert Cl ions into Cl0 atoms [53]. The Cl0 species oxidize the X-3B molecules and form AgCl. These reactions ensure that Ag/AgCl and S-CTF remain stable during the photodegradation process [16,20,37]. Secondly, the holes generate by TF yield •OH radicals or degrade organic pollutants directly. The photosensitization of X-3B [54] (generated electrons can inject into the CB of TF from excited dyes) also can be regarded as an advantage in its degradation.
Figure 8b shows the photocatalytic mechanism of S-CTF under UV light. The transformation of electron-hole pairs is different from visible irradiation. According to the CB and VB edge potentials of TiO2 (−0.29 and 2.91 eV [40]) and AgCl (0.09 and 3.16 [55]), both TiO2 and AgCl are excited under UV irradiation. The generated holes of AgCl can easily inject into the VB of TF [41], and then yield •OH with adsorbed water molecules, or directly degrade organic pollutants such as the holes generated by TF. The holes also yield Cl0 and reduce this to Cl via the same route under visible irradiation. The electrons generated from the CB of TF and AgCl transfer to Ag0, which serves as an electron trap under UV irradiation [48,56]. These electrons participate in the formation of •O2 radicals. Moreover, the oxygen vacancy in TF is also beneficial to trap electrons and promote their separation under visible or UV light irradiation [40].

3. Materials and Methods

3.1. Chemicals

TiCl4, acetylacetone (AcAc), triethylamine (TEA), methyl alcohol (MeOH), tetrahydrofuran (THF), NaOH and AgNO3 were purchased from Sinopharm International Co. Ltd. (Shanghai, China). Brilliant red X-3B was purchased from Shanghai Dyestuff Chemical Plant. All of the experiments in this research used deionized water. All of the chemical reagents were analytical grade, and were used without further purification.

3.2. Preparation of Catalysts

TiO2 fibers (TF) were synthesized by a sol-gel method combined with centrifugal spinning technique, followed by stage-temperature-programmed heat treatment in a steam atmosphere. The hydrolysis of TiCl4 (10 mL in MeOH) was carried out by mixing 4 mL NaOH aqueous solution (0.5 g mL−1) and 25 mL MeOH under ice-cold conditions. The reaction can be expressed as: TiCl 4 + H 2 O TiO 2 + HCl . During the reaction, NaOH neutralized the produced HCl to form NaCl, and the formed NaCl was supersaturated and precipitated due to its relatively small saturation in MeOH. Then, NaCl was filtered and collected for Ag/AgCl deposition. The TiO2 was chelated by AcAc to form a spinning solution. After the centrifugal spinning, precursor fibers (PF) were obtained and converted to TF by the heat treatment in steam as a reducing atmosphere. The detailed heat treatment program is shown in Table 2.
Ag/AgCl@TiO2 fibers (S-CTF) were prepared by the deposition of Ag/AgCl on TF. The recycled NaCl and AgNO3 were prepared into impregnation liquids. TF was soaked in 1 M NaCl and 0.1 M AgNO3 solution for 10 min prior for AgCl deposition, and washed by deionized water to rinse the excess elements after each soak. This process was repeated for five trials. After being dried at 353 K, the prepared AgCl@TiO2 fibers (CTF) were irradiated under a 1000 W xenon lamp (50 cm away from the light source) for 10 min to reduce partial Ag+ to Ag0 on the surface of the AgCl particles.

3.3. Characterization of Catalysts

To observe the structure and morphology of the prepared samples, scanning electron microscopy (SEM, JSM 6700F, JEOL, Tokyo, Japan) was employed. A Quadrasorb SI-MP system (Quantachrome, Boynton Beach, FL, USA) was used to acquire the N2 adsorption-desorption isotherms. The Brunauer-Emmett-Teller (BET) specific surface area was detected by a multipoint BET method, and the adsorption data was acquired in the relative pressure (P/P0) range of 0.05–0.3. The pore size distribution was calculated by employing the Density Functional Theory (DFT). The average pore size and pore volume data were obtained by the N2 adsorption volume with the relative pressure at 0.991. An X-ray diffractometer (XRD, D8 Advance, Bruker AXS GmbH, Karlsruhe, Germany) was employed to determine the crystal structure of the prepared photocatalysts. An ESCALAB 250 spectrometer (Thermo Electron Corp., Cramlington, UK) with an Al KR source and a charge neutralizer was used to perform the X-ray photoelectron spectroscopy (XPS). All the spectra were calibrated to the C 1s peak at 284.6 eV. A UV-vis spectrophotometer (UV-3100, Shimadzu, Kyoto, Japan) was employed to investigate the light absorption property of the photocatalysts samples. To measure the solid-state UV-vis diffuse reflectance spectra, an integrating sphere attachment was equipped with BaSO4 as a background in room temperature and air conditions. A fluorescence spectrometer (F-4600, Hitachi, Tokyo, Japan) was used to measure the photoluminescence (PL) spectra of all the samples.

3.4. Photoactivity Studies

The photocatalytic performance of S-CTF was evaluated by the decolorization of X-3B under visible and UV light. A 1000 W Xe lamp inserted by a glass optical filter to cut off short wavelength components (λ ˂ 420 nm) and a 250 W high pressure mercury lamp with a maximum wavelength at 365 nm served as visible and UV light sources, respectively. Before the photocatalytic reaction, 50 mg of each sample was added to 200 mL X-3B aqueous solution (20 mg/L) in an open fixed-bed photo-reactor [57] which was cooled through the circulation of water at 293 K. To establish the adsorption/desorption equilibrium, the photo-reactor was stirred in dark conditions for 30 min. Then, the reactor was placed under visible and UV irradiation with the irradiation intensity at the surface of the dye solution being 250 W m−2 and 15 W m−2, respectively, and cycled by a peristaltic pump with a flow rate of 20 mL min−1. During the photocatalytic reaction, 3 mL reaction solution was taken at each interval and filtrated through a 0.45 μm syringe filter. A UV-vis 1601 spectrophotometer (Shimadzu, Japan) was used to measure the concentration of X-3B by the variation of the maximum absorption wavelength (536 nm) of filtrate which was determined. The degradation efficiency was calculated by the formula R = ( 1 C / C 0 ) × 100 % , in which C and C 0 stood for the initial concentration and the concentration of dye at each moment in the reaction, respectively.
Species capture experiments were performed to investigate the mechanism of TF. 10.0 mM of TBA), BQ, AgNO3 and EDTA-Na2 were added into X-3B solution acting as •OH, •O2, electrons and holes scavengers. The reaction processes are same as that of the degradation reaction by S-CTF.

4. Conclusions

In summary, the plasmonic photocatalyst S-CTF was prepared by the gel-sol method to synthesize TF, and the impregnation-precipitation-photoreduction method to deposit Ag/AgCl particles on the produced TF. AgCl particles were deposited on TF with recycled NaCl as Cl sources, which were produced in the hydrolysis of TiCl4 by NaOH aqueous solution. The improved dye adsorption compared to TF was due to the enlarged specific surface area and narrowed pore-size distribution after AgCl deposition. The increased photocatalytic efficiency resulted from the SPR of Ag0, enhanced light absorption, narrowed band gap and efficient charge separation. With its outstanding photocatalytic performance and high stability, S-CTF is a promising application in organic pollutant treatment and wastewater purification. Furthermore, this mild, less-toxic and lower-cost synthesis process is accordance with environmental protection concepts.

Acknowledgments

This work was supported by the Natural Science Foundation of Shandong Province, China (No. ZR2011BM005) and the Fundamental Research Funds of Shandong University (No. 2015JC022).

Author Contributions

Xinhan Miao and Xiuyan Jie prepared the catalysts; Xinhan Miao and Xinde Hu performed the experiment; Xinhan Miao and Xiyue Zheng prepared the manuscript; Qingzhe Zhang directed the manuscript drafting; Nan Bao directed the project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Legrini, O.; Oliveros, E.; Braun, A.M. Photochemical processes for water treatment. Chem. Rev. 1993, 93, 671–698. [Google Scholar] [CrossRef]
  2. Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, G.; Yin, L.-C.; Wang, J.; Niu, P.; Zhen, C.; Xie, Y.; Cheng, H.-M. A red anatase TiO2 photocatalyst for solar energy conversion. Energy Environ. Sci. 2012, 5, 9603–9610. [Google Scholar] [CrossRef]
  4. Kudo, A. Development of photocatalyst materials for water splitting with the aim at photon energy conversion. J. Ceram. Soc. Jpn. 2001, 109, S81–S88. [Google Scholar] [CrossRef]
  5. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef]
  6. Choi, W.; Termin, A.; Hoffmann, M.R. The role of metal ion dopants in quantum-sized TiO2: Correlation between photoreactivity and charge carrier recombination dynamics. J. Phys. Chem. 1994, 98, 13669–13679. [Google Scholar] [CrossRef]
  7. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef] [PubMed]
  8. Liu, Y.; Liu, J.; Lin, Y.; Zhang, Y.; Wei, Y. Simple fabrication and photocatalytic activity of S-doped TiO2 under low power led visible light irradiation. Ceram. Int. 2009, 35, 3061–3065. [Google Scholar] [CrossRef]
  9. Chen, C.; Cai, W.; Long, M.; Zhou, B.; Wu, Y.; Wu, D.; Feng, Y. Synthesis of visible-light responsive graphene oxide/TiO2 composites with p/n heterojunction. ACS Nano 2010, 4, 6425–6432. [Google Scholar] [CrossRef] [PubMed]
  10. Bao, N.; Yin, Z.; Zhang, Q.; He, S.; Hu, X.; Miao, X. Synthesis of flower-like monoclinic BiVO4/surface rough TiO2 ceramic fiber with heterostructures and its photocatalytic property. Ceram. Int. 2016, 42, 1791–1800. [Google Scholar] [CrossRef]
  11. Yang, Y.; Wen, J.; Wei, J.; Xiong, R.; Shi, J.; Pan, C. Polypyrrole-decorated Ag-TiO2 nanofibers exhibiting enhanced photocatalytic activity under visible-light illumination. ACS Appl. Mater. Interfaces 2013, 5, 6201–6207. [Google Scholar] [CrossRef] [PubMed]
  12. Li, F.B.; Li, X.Z. Photocatalytic properties of gold/gold ion-modified titanium dioxide for wastewater treatment. Appl. Catal. A Gen. 2002, 228, 15–27. [Google Scholar] [CrossRef]
  13. Zhang, Q.; Thrithamarassery Gangadharan, D.; Liu, Y.; Xu, Z.; Chaker, M.; Ma, D. Recent advancements in plasmon-enhanced visible light-driven water splitting. J. Materiomics 2017, 3, 33–50. [Google Scholar] [CrossRef]
  14. Kumar, R.; El-Shishtawy, M.R.; Barakat, A.M. Synthesis and characterization of Ag-Ag2O/TiO2@polypyrrole heterojunction for enhanced photocatalytic degradation of methylene blue. Catalysts 2016, 6, 76. [Google Scholar] [CrossRef]
  15. Awazu, K.; Fujimaki, M.; Rockstuhl, C.; Tominaga, J.; Murakami, H.; Ohki, Y.; Yoshida, N.; Watanabe, T. A plasmonic photocatalyst consisting of silver nanoparticles embedded in titanium dioxide. J. Am. Chem. Soc. 2008, 130, 1676–1680. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, P.; Huang, B.; Qin, X.; Zhang, X.; Dai, Y.; Wei, J.; Whangbo, M.-H. Ag@AgCl: A highly efficient and stable photocatalyst active under visible light. Angew. Chem. Int. Ed. 2008, 47, 7931–7933. [Google Scholar] [CrossRef] [PubMed]
  17. Jin, R.; Cao, Y.; Mirkin, C.A.; Kelly, K.L.; Schatz, G.C.; Zheng, J.G. Photoinduced conversion of silver nanospheres to nanoprisms. Science 2001, 294, 1901–1903. [Google Scholar] [CrossRef] [PubMed]
  18. Jain, P.K.; Lee, K.S.; El-Sayed, I.H.; El-Sayed, M.A. Calculated absorption and scattering properties of gold nanoparticles of different size, shape, and composition: Applications in biological imaging and biomedicine. J. Phys. Chem. B 2006, 110, 7238–7248. [Google Scholar] [CrossRef] [PubMed]
  19. Kakuta, N.; Goto, N.; Ohkita, H.; Mizushima, T. Silver bromide as a photocatalyst for hydrogen generation from CH3OH/H2O solution. J. Phys. Chem. B 1999, 103, 5917–5919. [Google Scholar] [CrossRef]
  20. Yu, J.; Dai, G.; Huang, B. Fabrication and characterization of visible-light-driven plasmonic photocatalyst Ag/AgCl/TiO2 nanotube arrays. J. Phys. Chem. C 2009, 113, 16394–16401. [Google Scholar] [CrossRef]
  21. Guo, J.-F.; Ma, B.; Yin, A.; Fan, K.; Dai, W.-L. Highly stable and efficient Ag/AgCl@TiO2 photocatalyst: Preparation, characterization, and application in the treatment of aqueous hazardous pollutants. J. Hazard. Mater. 2012, 211–212, 77–82. [Google Scholar] [CrossRef] [PubMed]
  22. Hu, C.; Lan, Y.; Qu, J.; Hu, X.; Wang, A. Ag/AgBr/TiO2 visible light photocatalyst for destruction of azodyes and bacteria. J. Phys. Chem. B 2006, 110, 4066–4072. [Google Scholar] [CrossRef] [PubMed]
  23. Nunes, D.; Pimentel, A.; Santos, L.; Barquinha, P.; Fortunato, E.; Martins, R. Photocatalytic TiO2 nanorod spheres and arrays compatible with flexible applications. Catalysts 2017, 7, 60. [Google Scholar] [CrossRef]
  24. Liao, W.; Zhang, Y.; Zhang, M.; Murugananthan, M.; Yoshihara, S. Photoelectrocatalytic degradation of microcystin-LR using Ag/AgCl/TiO2 nanotube arrays electrode under visible light irradiation. Chem. Eng. J. 2013, 231, 455–463. [Google Scholar] [CrossRef]
  25. Wang, D.; Li, Y.; Li Puma, G.; Wang, C.; Wang, P.; Zhang, W.; Wang, Q. Ag/AgCl@helical chiral TiO2 nanofibers as a visible-light driven plasmon photocatalyst. Chem. Commun. 2013, 49, 10367–10369. [Google Scholar] [CrossRef] [PubMed]
  26. Wang, P.; Huang, B.; Lou, Z.; Zhang, X.; Qin, X.; Dai, Y.; Zheng, Z.; Wang, X. Synthesis of highly efficient Ag@AgCl plasmonic photocatalysts with various structures. Chem. Eur. J. 2010, 16, 538–544. [Google Scholar] [CrossRef] [PubMed]
  27. Ma, B.; Guo, J.; Dai, W.-L.; Fan, K. Ag-AgCl/WO3 hollow sphere with flower-like structure and superior visible photocatalytic activity. Appl. Catal. B Environ. 2012, 123–124, 193–199. [Google Scholar] [CrossRef]
  28. Di Paola, A.; Bellardita, M.; Ceccato, R.; Palmisano, L.; Parrino, F. Highly active photocatalytic TiO2 powders obtained by thermohydrolysis of TiCl4 in water. J. Phys. Chem. C 2009, 113, 15166–15174. [Google Scholar] [CrossRef]
  29. Zhu, Y.; Zhang, L.; Gao, C.; Cao, L. The synthesis of nanosized TiO2 powder using a sol-gel method with TiCl4 as a precursor. J. Mater. Sci. 2000, 35, 4049–4054. [Google Scholar] [CrossRef]
  30. Xu, H.; Gao, L.; Guo, J. Hydrothermal synthesis of tetragonal barium titanate from barium chloride and titanium tetrachloride under moderate conditions. J. Am. Ceram. Soc. 2002, 85, 727–729. [Google Scholar] [CrossRef]
  31. Bao, N.; Yin, G.-B.; Wei, Z.-T.; Li, Y.; Ma, Z.-H. Preparation of TiO2 continuous fibers with oxygen vacancies and photocatalytic activity. Integr. Ferroelectr. 2011, 127, 97–105. [Google Scholar] [CrossRef]
  32. Zhou, J.; Cheng, Y.; Yu, J. Preparation and characterization of visible-light-driven plasmonic photocatalyst Ag/AgCl/TiO2 nanocomposite thin films. J. Photochem. Photobiol. A Chem. 2011, 223, 82–87. [Google Scholar] [CrossRef]
  33. Dai, Y.; Lu, X.; McKiernan, M.; Lee, E.P.; Sun, Y.; Xia, Y. Hierarchical nanostructures of K-birnessite nanoplates on anatase nanofibers and their application for decoloration of dye solution. J. Mater. Chem. 2010, 20, 3157–3162. [Google Scholar] [CrossRef]
  34. Zhuang, X.; Wan, Y.; Feng, C.; Shen, Y.; Zhao, D. Highly efficient adsorption of bulky dye molecules in wastewater on ordered mesoporous carbons. Chem. Mater. 2009, 21, 706–716. [Google Scholar] [CrossRef]
  35. Bao, N.; Li, Y.; Wei, Z.; Yin, G.; Niu, J. Adsorption of dyes on hierarchical mesoporous TiO2 fibers and its enhanced photocatalytic properties. J. Phys. Chem. C 2011, 115, 5708–5719. [Google Scholar] [CrossRef]
  36. Xie, C.; Yang, S.; Shi, J.; Niu, C. Highly crystallized C-doped mesoporous anatase TiO2 with visible light photocatalytic activity. Catalysts 2016, 6, 117. [Google Scholar] [CrossRef]
  37. Xu, H.; Li, H.; Xia, J.; Yin, S.; Luo, Z.; Liu, L.; Xu, L. One-pot synthesis of visible-light-driven plasmonic photocatalyst Ag/AgCl in ionic liquid. ACS Appl. Mater. Interfaces 2011, 3, 22–29. [Google Scholar] [CrossRef] [PubMed]
  38. Ghicov, A.; Yamamoto, M.; Schmuki, P. Lattice Widening in Niobium-Doped TiO2 Nanotubes: Efficient Ion Intercalation and Swift Electrochromic Contrast. Angew. Chem. Int. Ed. 2008, 47, 7934–7937. [Google Scholar] [CrossRef] [PubMed]
  39. Qiu, B.; Zhou, Y.; Ma, Y.; Yang, X.; Sheng, W.; Xing, M.; Zhang, J. Facile synthesis of the Ti3+ self-doped TiO2-graphene nanosheet composites with enhanced photocatalysis. Sci. Rep. 2015, 5, 8591. [Google Scholar] [CrossRef] [PubMed]
  40. Li, K.; Gao, S.; Wang, Q.; Xu, H.; Wang, Z.; Huang, B.; Dai, Y.; Lu, J. In-situ-reduced synthesis of Ti3+ self-doped TiO2/g-C3N4 heterojunctions with high photocatalytic performance under led light irradiation. ACS Appl. Mater. Interfaces 2015, 7, 9023–9030. [Google Scholar] [CrossRef] [PubMed]
  41. Bu, Y.; Chen, Z.; Feng, C.; Li, W. Study of the promotion mechanism of the photocatalytic performance and stability of the Ag@AgCl/g-C3N4 composite under visible light. RSC Adv. 2014, 4, 38124–38132. [Google Scholar] [CrossRef]
  42. Wang, P.; Huang, B.; Qin, X.; Zhang, X.; Dai, Y.; Whangbo, M.-H. Ag/AgBr/WO3·H2O: Visible-light photocatalyst for bacteria destruction. Inorg. Chem. 2009, 48, 10697–10702. [Google Scholar] [CrossRef] [PubMed]
  43. Xu, Y.; Xu, H.; Li, H.; Xia, J.; Liu, C.; Liu, L. Enhanced photocatalytic activity of new photocatalyst Ag/AgCl/ZnO. J. Alloys Compd. 2011, 509, 3286–3292. [Google Scholar] [CrossRef]
  44. Wang, Z.; Wen, B.; Hao, Q.; Liu, L.-M.; Zhou, C.; Mao, X.; Lang, X.; Yin, W.-J.; Dai, D.; Selloni, A.; et al. Localized excitation of Ti3+ ions in the photoabsorption and photocatalytic activity of reduced rutile TiO2. J. Am. Chem. Soc. 2015, 137, 9146–9152. [Google Scholar] [CrossRef] [PubMed]
  45. Yin, H.; Wang, X.; Wang, L.; Nie, Q.; Zhang, Y.; Yuan, Q.; Wu, W. Ag/AgCl modified self-doped TiO2 hollow sphere with enhanced visible light photocatalytic activity. J. Alloys Compd. 2016, 657, 44–52. [Google Scholar] [CrossRef]
  46. Han, L.; Wang, P.; Zhu, C.; Zhai, Y.; Dong, S. Facile solvothermal synthesis of cube-like Ag@AgCl: A highly efficient visible light photocatalyst. Nanoscale 2011, 3, 2931–2935. [Google Scholar] [CrossRef] [PubMed]
  47. Ma, B.; Guo, J.; Dai, W.-L.; Fan, K. Highly stable and efficient Ag/AgCl core-shell sphere: Controllable synthesis, characterization, and photocatalytic application. Appl. Catal. B Environ. 2013, 130, 257–263. [Google Scholar] [CrossRef]
  48. Sung-Suh, H.M.; Choi, J.R.; Hah, H.J.; Koo, S.M.; Bae, Y.C. Comparison of ag deposition effects on the photocatalytic activity of nanoparticulate TiO2 under visible and UV light irradiation. J. Photochem. Photobiol. A Chem. 2004, 163, 37–44. [Google Scholar] [CrossRef]
  49. Yu, J.-G.; Yu, H.-G.; Cheng, B.; Zhao, X.-J.; Yu, J.C.; Ho, W.-K. The effect of calcination temperature on the surface microstructure and photocatalytic activity of TiO2 thin films prepared by liquid phase deposition. J. Phys. Chem. B 2003, 107, 13871–13879. [Google Scholar] [CrossRef]
  50. Cheng, H.; Huang, B.; Dai, Y.; Qin, X.; Zhang, X. One-step synthesis of the nanostructured AgI/BiOI composites with highly enhanced visible-light photocatalytic performances. Langmuir 2010, 26, 6618–6624. [Google Scholar] [CrossRef] [PubMed]
  51. Klaysri, R.; Tubchareon, T.; Praserthdam, P. One-step synthesis of amine-functionalized TiO2 surface for photocatalytic decolorization under visible light irradiation. J. Ind. Eng. Chem. 2017, 45, 229–236. [Google Scholar] [CrossRef]
  52. Zhu, M.; Chen, P.; Liu, M. Ag/Agbr/graphene oxide nanocomposite synthesized via oil/water and water/oil microemulsions: A comparison of sunlight energized plasmonic photocatalytic activity. Langmuir 2012, 28, 3385–3390. [Google Scholar] [CrossRef] [PubMed]
  53. Sun, L.; Zhang, R.; Wang, Y.; Chen, W. Plasmonic Ag@AgCl nanotubes fabricated from copper nanowires as high-performance visible light photocatalyst. ACS Appl. Mater. Interfaces 2014, 6, 14819–14826. [Google Scholar] [CrossRef] [PubMed]
  54. Bao, N.; Li, Y.; Yu, X.-H.; Niu, J.-J.; Wu, G.-L.; Xu, X.-H. Removal of anionic azo dye from aqueous solution via an adsorption–photosensitized regeneration process on a TiO2 surface. Environ. Sci. Pollut. Res. 2013, 20, 897–906. [Google Scholar] [CrossRef] [PubMed]
  55. Ye, L.; Liu, J.; Gong, C.; Tian, L.; Peng, T.; Zan, L. Two different roles of metallic Ag on Ag/AgX/BiOX (X = Cl, Br) visible light photocatalysts: Surface plasmon resonance and Z-scheme bridge. ACS Catal. 2012, 2, 1677–1683. [Google Scholar] [CrossRef]
  56. Hirakawa, T.; Kamat, P.V. Charge separation and catalytic activity of Ag@TiO2 core-shell composite clusters under UV-irradiation. J. Am. Chem. Soc. 2005, 127, 3928–3934. [Google Scholar] [CrossRef] [PubMed]
  57. Zhang, Q.; Bao, N.; Wang, X.; Hu, X.; Miao, X.; Chaker, M.; Ma, D. Advanced fabrication of chemically bonded graphene/TiO2 continuous fibers with enhanced broadband photocatalytic properties and involved mechanisms exploration. Sci. Rep. 2016, 6, 38066. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEM images of TF (a,b), CTF (c,d) and S-CTF (e,f).
Figure 1. SEM images of TF (a,b), CTF (c,d) and S-CTF (e,f).
Catalysts 07 00117 g001
Figure 2. N2 adsorption-desorption isotherms (a) and pore-size distribution (b) of TF and S-CTF.
Figure 2. N2 adsorption-desorption isotherms (a) and pore-size distribution (b) of TF and S-CTF.
Catalysts 07 00117 g002
Figure 3. XRD patterns of NaCl separated from reaction solution (a), AgCl, Ag/AgCl, TF, CTF and S-CTF (b).
Figure 3. XRD patterns of NaCl separated from reaction solution (a), AgCl, Ag/AgCl, TF, CTF and S-CTF (b).
Catalysts 07 00117 g003
Figure 4. XPS spectra of all elements of S-CTF (a) and the band energy of Ti 2p (b), O 1s (c), Cl 2p (d) and Ag 3d (e).
Figure 4. XPS spectra of all elements of S-CTF (a) and the band energy of Ti 2p (b), O 1s (c), Cl 2p (d) and Ag 3d (e).
Catalysts 07 00117 g004aCatalysts 07 00117 g004b
Figure 5. UV-vis diffuse reflectance spectrum of Ag/AgCl, TF, CTF and S-CTF.
Figure 5. UV-vis diffuse reflectance spectrum of Ag/AgCl, TF, CTF and S-CTF.
Catalysts 07 00117 g005
Figure 6. Photocatalytic degradation of 100 mL X-3B (20 mg L−1) over pure-TF TF, CTF, S-CPF and S-CTF (0.05 g) under visible light (a) and ultrasonic light (b) with blank control groups.
Figure 6. Photocatalytic degradation of 100 mL X-3B (20 mg L−1) over pure-TF TF, CTF, S-CPF and S-CTF (0.05 g) under visible light (a) and ultrasonic light (b) with blank control groups.
Catalysts 07 00117 g006
Figure 7. Photoluminescence spectra of TF, Ag/AgCl and S-CTF.
Figure 7. Photoluminescence spectra of TF, Ag/AgCl and S-CTF.
Catalysts 07 00117 g007
Figure 8. Photocatalytic mechanisms of S-CTF under visible (a) and UV (b) light irradiation.
Figure 8. Photocatalytic mechanisms of S-CTF under visible (a) and UV (b) light irradiation.
Catalysts 07 00117 g008
Table 1. Textural Parameters of TF (TiO2 fibers) and S-CTF (Ag/AgCl@TiO2 fibers).
Table 1. Textural Parameters of TF (TiO2 fibers) and S-CTF (Ag/AgCl@TiO2 fibers).
SamplesBET Surface Area (m2 g−1)Pore Volume (cm3 g−1)Pore Diameter (nm)
TF27.90.11116.7
S-CTF30.60.07410.8
Table 2. Program of Segmented Heating.
Table 2. Program of Segmented Heating.
Temperature (°C)Heating Rate (°C min−1)Time (min)
25–951.6742
95–2501.29120
250–3500.83120
350–5505.0040
550–6000.8360
600–600-120

Share and Cite

MDPI and ACS Style

Bao, N.; Miao, X.; Hu, X.; Zhang, Q.; Jie, X.; Zheng, X. Novel Synthesis of Plasmonic Ag/AgCl@TiO2 Continues Fibers with Enhanced Broadband Photocatalytic Performance. Catalysts 2017, 7, 117. https://doi.org/10.3390/catal7040117

AMA Style

Bao N, Miao X, Hu X, Zhang Q, Jie X, Zheng X. Novel Synthesis of Plasmonic Ag/AgCl@TiO2 Continues Fibers with Enhanced Broadband Photocatalytic Performance. Catalysts. 2017; 7(4):117. https://doi.org/10.3390/catal7040117

Chicago/Turabian Style

Bao, Nan, Xinhan Miao, Xinde Hu, Qingzhe Zhang, Xiuyan Jie, and Xiyue Zheng. 2017. "Novel Synthesis of Plasmonic Ag/AgCl@TiO2 Continues Fibers with Enhanced Broadband Photocatalytic Performance" Catalysts 7, no. 4: 117. https://doi.org/10.3390/catal7040117

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop