Highly Efficient Photocatalytic Hydrogen Evolution ... - ACS Publications

0 downloads 0 Views 2MB Size Report
Jul 19, 2016 - TiO2/CuO/Cu Thoroughly Mesoporous Nanofibers. Huilin Hou, Minghui Shang, Fengmei Gao, Lin Wang, Qiao Liu, Jinju Zheng, Zuobao Yang, ...
Research Article www.acsami.org

Highly Efficient Photocatalytic Hydrogen Evolution in Ternary Hybrid TiO2/CuO/Cu Thoroughly Mesoporous Nanofibers Huilin Hou, Minghui Shang, Fengmei Gao, Lin Wang, Qiao Liu, Jinju Zheng, Zuobao Yang, and Weiyou Yang* Institute of Materials, Ningbo University of Technology, Ningbo City 315016, P. R. China S Supporting Information *

ABSTRACT: Development of novel hybrid photocatalysts with high efficiency and durability for photocatalytic hydrogen generation is highly desired but still remains a grand challenge currently. In the present work, we reported the exploration of ternary hybrid TiO2/CuO/Cu thoroughly mesoporous nanofibers via a foaming-assisted electrospinning technique. It is found that by adjusting the Cu contents in the solutions, the unitary (TiO2), binary (TiO2/CuO, TiO2/Cu), and ternary (TiO2/CuO/Cu) mesoporous products can be obtained, enabling the growth of TiO2/CuO/Cu ternary hybrids in a tailored manner. The photocatalytic behavior of the as-synthesized products as well as P25 was evaluated in terms of their hydrogen evolution efficiency for the photodecomposition water under Xe lamp irradiation. The results showed that the ternary TiO2/CuO/Cu thoroughly mesoporous nanofibers exhibit a robust stability and the most efficient photocatalytic H2 evolution with the highest release rate of ∼851.3 μmol g−1 h−1, which was profoundly enhanced for more than 3.5 times with respect to those of the pristine TiO2 counterparts and commercial P25, suggesting their promising applications in clean energy production. KEYWORDS: TiO2/CuO/Cu, electrospinning, mesoporous, nanofibers, photocatalytic hydrogen evolution

1. INTRODUCTION Photocatalytic water splitting for hydrogen production by utilizing solar energy is considered as one of the promising strategies for resolving global energy and environmental problems, owing to its renewable energy production with no reliance on fossil fuels and no carbon dioxide emission.1−4 Titanium dioxide (TiO2), as one of the most representative photocatalysts, is used widely because of its excellent chemical resistance, low cost, nontoxicity, availability, and long-term stability against the photochemical corrosion in aggressive aqueous environments.5−8 However, the conventional TiO2 material has the intrinsic shortcomings: (i) low specific surface areas; (ii) easy aggregation; (iii) the usually fast electron−hole recombination; (iv) large band gap barrier resulting in low photocatalytic efficiency.9,10 To overcome these hindrances, many effective techniques are applied to pursue the satisfied structures and/or modify the TiO2 by incorporating other cocatalytic materials.11−14 Typically, one-dimensional (1D) mesoporous nanofibers could be an ideal candidate, since they possess low aggregation tendency, high porosity, and large specific surface area, which can promote the charge and mass transfer for enhanced photocatalytic activities,15−18 especially for the heterostructured counterparts.19−21 Introduction of CuO into the TiO2 matrix is recognized as a cost-effective and emerging strategy, which could not only help the full utilization of the solar light energy but also render a built-in electric field at the interface © 2016 American Chemical Society

within the hybrid, which profoundly facilitates the separation of photogenerated electron and hole pairs.22,23 Additionally, metallic Cu-modified TiO2 can promote the electron transfer and e−h separation, leading to enhanced photocatalytic activities.24,25 That is, the exploration of CuO and/or Cu-modified TiO2 mesoporous hybrid nanofibers could be a promising strategy to offer the photocatalysts with innovative physicochemical properties. Electrospinning is a versatile, compatible, low cost, and productive technique for generating 1D nanofibers in various material systems with controllable morphologies.26−29 By virtue of this technique assisted by the subsequent air calcination, TiO2/CuO nanofibers have been successfully fabricated to meet the requirements of outstanding photocatalysts,30 which could endow the synergy of favorable photocatalytic reactions such as good contact of the heterojunctions, long lifetimes of photogenerated charge carriers, and 1D structure for efficient charge transfer.31−33 However, to date, there is little work devoted to the investigation on ternary TiO2/CuO/Cu mesoporous nanofibers. Herein, we report the exploration of TiO2/CuO/Cu thoroughly mesoporous nanofibers via the electrospinning technique. The diisopropyl azodiformate (DIPA) used as the foaming agents Received: June 3, 2016 Accepted: July 19, 2016 Published: July 19, 2016 20128

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces

2. EXPERIMENTAL PROCEDURE

was added into the initial spinning solutions and homogeneously boxed throughout the precursor (tetrabutyl titanate (TBOT), polyvinylpyrrolidone (PVP)) and C4H6CuO4·H2O to create the porous structures. The obtained 1D ternary TiO2/CuO/Cu mesoporous nanostructures exhibited significantly enhanced photocatalytic activities, as compared to the binary hybrids (i.e., TiO2/CuO and TiO2/Cu), unitary photocatalysts (i.e., TiO2), and the commercial P25 products.

Materials. Polyvinylpyrrolidone (PVP, Mw ≈ 1 300 000), butyl titanate (TBOT), diisopropyl azodiformate (DIPA), absolute ethyl alcohol, acetic acid, copper(II) acetate monohydrate (C4H6CuO4·H2O), methanol, and deionized water were all purchased from Aladdin. All chemicals were used directly without further purification. Preparation of Bare TiO2 and CuO and/or Cu-Modified TiO2 Thoroughly Mesoporous Nanofibers. Both pristine TiO2 and CuO and/or Cu-modified mesoporous nanofibers were synthesized according to the foaming agent assisted electrospinning method as reported in our previous work.34 In a typical procedure, 6.5 wt % of PVP, 30 wt % of TBOT, and 10 wt % of foamer (DIPA) were added into a solution comprising ethanol and acetic acid (3:1 in volume) with vigorous stirring for 12 h to obtain a homogeneous precursor solution for the fabrication of pristine TiO2. The precursor solution for the fabrication of CuO and/or Cu-modified TiO2 thoroughly mesoporous nanofibers was performed by adding 2 wt % of C4H6CuO4·H2O into the above mixture with further magnetically stirring for 2 h. For comparison, another blank experiment without the introduction of DIPA in the initial solutions.

Table 1. Details of Five Solutions Used for Electrospinning Polymer Precursor Fibers sample

PVP (wt %)

TBOT (wt %)

Cu(Ac)2 (wt %)

former (wt %)

A B C D E

6.5 6.5 6.5 6.5 6.5

30 30 30 30 30

0 2 2 1 3

10 0 10 10 10

Figure 1. (a, b) Typical SEM images of the calcined products of sample C under low magnifications. (c, d) Representative SEM images under high magnifications of the calcined products of sample C. (e) A representative XRD pattern of the calcined products of sample C. (f) A corresponding close-up XRD pattern (30°−45°) of (e). 20129

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces

Figure 2. TEM characterization of the hybrid nanofibers of sample C. (a) A typical TEM image of a single mesoporous nanofiber under low magnification. (b) A typical EDX pattern recorded from the marked area of A in (a). (c) The element mapping of Cu within a single fiber. (d) The corresponding SAED pattern recorded from the marked area of B in (a). (e) A typical HRTEM image recorded from the marked area of C in (a). (g, f) Enlarged images recorded from the marked areas of D and E in (e), respectively. To disclose the effect of incorporated Cu on the growth of the hybrid fibers, two solution samples are also prepared with the Cu sources decreased to 1 wt % and otherwise increased to 3 wt %. The details for these five initial solutions are shown in Table 1, and the resultant products are referred to as samples A−E. Then, the as-prepared solutions were transformed into a plastic syringe with a stainless steel nozzle, which was sized in ∼0.2 mm in diameter and used as the anode for electrospinning. The tip of the stainless steel nozzle was placed in the front of a metal cathode (collector) with a fixed distance of 20 cm. An electrical potential of 20 kV was applied for electrospinning of the precursor fibers. Subsequently, the precursor fibers were heated in a conventional tube furnace to the desired temperature of 550 °C with a heating rate of 1 °C min−1 and maintained there for 2 h in air, followed by furnace cool to ambient temperature. Characterization. The obtained products were characterized with X-ray powder diffraction (XRD, D8 Advance, Bruker, Germany) with Cu Kα radiation (λ = 1.5406 Å), field emission scanning electron microscopy (FESEM, S-4800, Hitachi, Japan), and high-resolution transmission electron microscopy (HRTEM, JEM-2010F, JEOL, Japan) equipped with energy dispersive X-ray spectroscopy (EDX, Quantax-STEM,

Bruker, Germany). The porous properties of the as-prepared mesoporous nanofibers were characterized using N2 adsorption at 77 K on a specific surface area and porosity analyzer (Micromeritics, ASAP 2020HD88, USA). The elemental compositions and chemical states of the hybrid fibers were studied by X-ray photoelectron spectroscopy (Shimadzu, AXIS ULTRA DLD, Japan). The diffuse reflectance absorption spectra of the products were recorded on a UV−vis spectrophotometer (UV-3900, Hitachi, Japan) equipped with an integrated sphere attachment. Photocatalytic Activity Measurements. The photocatalytic reaction is performed in an inner-irradiation quartz annular reactor with a 300 W xenon lamp (CEL, HUL300), a vacuum pump, a gas collection, a recirculation pump, and a water-cooled condenser. The as-synthesized samples (0.1 g) were suspended in deionized water and methanol mixed solutions (40 mL, 3:1 in volume) by an ultrasonic oscillator. Then the mixture was transferred into the reactor and deaerated by the vacuum pump. The xenon lamp was utilized as a light source, and the cooling water was circulated through a cylindrical Pyrex jacket located around the light source to maintain the reaction temperature. The reactor was sealed with ambient air during irradiation, 20130

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces

The diameter of the products is slightly increased to ∼400 nm, which could be due to the gas release from the foaming agents during the calcination process, causing the fiber in somewhat expanding. Closer observations of the fiber bodies represent that they have rough surfaces with numerous irregular-shaped pores within the fibers, which are typically sized in ∼25 nm. Their representative cross-section view of the mesoporous nanofibers (Figure 1d) discloses that the fibers possess a thoroughly mesoporous structure throughout the entire bodies. Figure 1e shows the typical XRD pattern of the as-fabricated mesoporous nanofibers, which well matches the phase of anatase TiO2 (JCPDS: No. 21-1272).35 The close-up XRD pattern (Figure 1f) in the region of 2θ = 30°−45° suggests that besides the major phase of anatase TiO2, some low intensity diffraction peaks are also detected (i.e., 2θ = 35.6°, 38.8°, and 43.2°), which correspond to the diffractions of (002) and (111) crystal faces of CuO (JCPDS,

and the hydrogen evolution was monitored by an online gas chromatograph (GC, 7900) equipped with a Porapak-Q column, high-purity nitrogen carrier, and a thermal conductivity detector (TCD). In order to investigate the stability and recyclability, the products were reused for three cycles. For comparison, Degussa P25 was commercially available and used directly for the photocatalytic H2 generation.

3. RESULTS AND DISCUSSION The electrospun hybrid precursor nanofibers of sample C are first observed by SEM, which are shown in Figure S1 (Supporting Information). They appear smooth without any obvious pores with a mean diameter of ∼300 nm. Figure 1a−d displays the typical SEM images of the corresponding calcined fibers under different magnifications. It seems that all the precursor fibers of sample C have been completely converted into mesoporous nanofibers, disclosing their high purity in morphology (Figure 1a,b).

Figure 3. Typical XPS spectra of TiO2/CuO/Cu mesoporous nanofibers of sample C: (a) survey spectrum, (b) Ti 2p, (c) O 1s, and (d) Cu 2p. (e) A schematic illustration of the atomic structure of constructed TiO2−CuO−Cu heterojunction. 20131

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces

Figure 4. Typical SEM images, TEM images, and XRD patterns of the calcined products of sample A (a1−c1), sample B (b1−b3), sample D (c1−c3), and sample E (d1−d3), respectively.

(Figure 2d) taken from the marked area of B in (a) exhibits the typical diffraction rings, which is suggestive of their polycrystalline nature. Figure 2e and Figure S3 show the representative HRTEM images recorded form the marked area of C in (a), showing three sets of lattice fringes. The left side area with the d-space of 0.357 nm matches that of the (101) plane of anatase TiO2. Figures 2f and 2g are the enlarged images corresponding to the marked areas of D and E in Figure 2d, respectively. The measured two spacings in Figure 2f with ∼0.232 and 0.208 nm correspond to the d-distances of the (111) plane of CuO and the (111) plane of Cu, respectively. Figure 2e also means that the Cu is typically inlaid in the CuO matrix. More details concerning the elemental compositions and chemical states of the mesoporous hybrid nanofibers were further

No. 48-1548) and (111) of Cu (JCPDS, No. 04-0836), respectively.24,36 These results verify that the as-fabricated thoroughly mesoporous fibers are ternary phases, which consist anatase TiO2, CuO, and Cu. Figure 2a presents a typical TEM image of the obtained fiber under a low magnification, suggesting that the fiber possesses a thoroughly mesoporous structure based on the different contrast throughout the fiber. The corresponding EDX spectrum (Figure 2b) recorded from the marked area of A in Figure 2a (also see the element line scanning spectra in Figure S2, Supporting Information) implies the coexistence of Ti, Cu, and Cu in the hybrids. It is notable that the Cu elements have a uniform spatial distribution within the TiO2 mesoporous matrix (Figure 2c). The corresponding selective area electron diffraction (SAED) pattern 20132

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces investigated by X-ray photoelectron spectroscopy (XPS). Figure 3a shows the representative XPS survey spectrum of sample C, revealing that the existence of Ti, O, Cu, and C elements within the fibers, which is in consistent with the EDX spectrum. The presence of C 1s can be ascribed to the adventitious carbon-based contaminant from the XPS instrument itself. The high resolution XPS spectra of Ti 2p, O 1s, and Cu 2p are given in Figure 3b−d, respectively. The peaks in Ti 2p spectrum (Figure 3b) show the peaks centered at ca. 458.1 and 463.7 eV, which are ascribed to the Ti 2p3/2 and Ti 2p1/2, respectively, suggesting the existence of Ti4+ oxidation state.37 Figure 3c provides the high resolution XPS spectra of O 1s. The peak at the binding energy of ca. 529.3 eV is assigned to Ti−O and Cu−O species, whereas the one at a higher binding energy of ca. 531.2 eV indicates the surface contamination by hydroxides and carbonate from the atmosphere.30 Therefore, the O is considered as coming from metal oxides (TiO2 and CuO) and adsorbed oxygen such as the surface hydroxyl species and carbonate species. As shown in Figure 3d, the main peaks at ca. 933.4 and 953.2 eV, along with the presence of their characteristic shakeup satellite peaks at ca. 941.3 and 961.5 eV, respectively, verify the existence of CuO.38 Besides the peaks belong to CuO, the detected two minor peaks centered at ca. 931.2 and 951.1 eV are possibly attributed to other copper species (i.e., Cu2O or metallic Cu). They should be derived from the metallic Cu,24,39 since no Cu2O has been observed according to the characterizations of XRD and HRTEM. Accordingly, it means that the as-fabricated thoroughly mesoporous fibers should be ternary TiO2/CuO/Cu heterojunctions, which is schematically illustrated in Figure 3e. In order to account for the growth mechanism and achieve the tailored fabrication of the TiO2/CuO/Cu thoroughly mesoporous hybrid nanofibers, another four comparative experiments are carried out by adjusting the compositions of the initial solutions with other similar experimental conditions (see Table 1 for details). Figure 4 shows the SEM, TEM, and XRD characterizations of the obtained samples. It seems like that only pristine anatase TiO2 mesoporous nanofibers (Figure 4(a1−a3)) of sample A can be obtained. Meanwhile, the as-fabricated products of sample B (i.e., without the foaming agent introduce) present the ordinary smooth TiO2/Cu mesoporous nanofibers (Figure 4(b1−b3)). As compared between sample B and C, it implies that the introduced foaming agents can not only create the mesopores throughout the fiber body but also play a profound effect on the evolution of the phase compositions. To further verify this point, the Cu sources in the initial solutions are reduced to 1 wt % (sample D, Figure 4(c1−c3)) and otherwise increased to 3 wt % (sample E, Figure 4(d1−d3)). The experimental results demonstrate that both obtained samples are mesoporous fibers. However, the XRD patterns (Figure 4(c3,d3)) disclose that they possess different phases with TiO2/CuO and TiO2/CuO/Cu for sample D and E, respectively. According to the experimental results of samples B−E, it discloses that when the foaming agents is absent (i.e., sample B) and the introduced Cu sources in the initial solutions is low enough (i.e., sample D), only binary hybrid products can be obtained, suggesting that a higher content of Cu sources (i.e., samples C and E) favors the formation of ternary TiO2/CuO/Cu hybrids. Figure 5 shows the nitrogen adsorption−desorption isotherms (a) and pore size distribution curves (b) of samples A−E. Except for sample B, the other four samples exhibit the typical type IV adsorption isotherms with hysteresis loops according to BDDT classification, indicating the formation of mesoporous fibers.40,41 Their structural details are summarized in Table 2. It shows that

Figure 5. (a) Typical N2 adsorption and desorption isotherms of the calcined products of samples A−E. (b) The corresponding pore size distribution of samples A−E. (c) UV−vis diffuse reflectance absorption spectra of samples A−E.

Table 2. Structural Parameters of Samples A−E Derived from the Nitrogen Adsorption−Desorption Isotherms samples

SBETa (m2/g)

pore volb (cm3/g)

av pore sizeb (nm)

A B C D E

45.3 11.3 19.6 24.8 13.6

0.10 0.04 0.11 0.13 0.08

10.6 6.1 34.8 24.4 35.7

a

The BET specific surface area was determined by multipoint BET method using the adsorption data. bThe pore volume and average pore size were determined by nitrogen adsorption volume.

the pristine TiO2 products of sample A possess the highest BET surface area (45.3 m2/g) than the other products, which can be attributed to the increscent pore diameter by the foamer complexing of Cu2+ in the Cu and/or CuO incorporated TiO2 hybrids. However, the hybrid mesoporous products of samples C and D possess higher pore volumes than the pristine counterpart, implying 20133

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces a more complex mesoporosity structure of the hybrid fibers. The UV−vis absorption spectra were employed to track the change of light absorbance characteristics of the as-prepared products. As shown in Figure 5c, there is only a steep absorption edge at the UV region and scarcely absorption in the visible-light region for the pristine TiO2 of sample A. However, after Cu species incorporated to the TiO2 matrix, the hybrids exhibit an additional broad absorption band at 400−800 nm, whose intensities are enhanced with the increase of Cu sources introduced within the solutions, suggesting their narrowed bandgap and improved photoinduced electrons−holes generation. The smaller bandgap can be attributed to the transition from the valence band of CuO to the conduction band of TiO2, which are both in good contact as witnessed by the HRTEM image in Figure 2e. The difference among the UV−vis absorption spectra of the samples could be resulted from the various surface areas and porosity volumes, as

evidenced by the BET specific surface area and N2 adsorption− desorption isotherm characterizations (Figure 5a,b and Table 2). The photocatalytic H2 production activities over the as-prepared five products as well as P25 were evaluated by using methanol as sacrificial agents and irradiation under a 300 W xenon arc lamp. Figure 6a plots the amounts of hydrogen evolved from the aqueous suspensions over the six photocatalysts according to the observed GC analysis patterns (Figure S5), and the corresponding average hydrogen production rates are presented in Figure 6b. The results suggest that the hydrogen evolution rate of the pristine TiO2 mesoporous nanofibers of sample A (ca. 231.7 μmol g−1 h−1) is higher than that of P25 (ca. 200.5 μmol g−1 h−1) and lower than those of samples B−E. It is notable that the TiO2/CuO/Cu ternary thoroughly mesoporous products of sample C (ca. 851.3 μmol g−1 h−1) and D (ca. 550.3 μmol g−1 h−1) exhibit the significantly enhanced photocatalytic activities as compared to

Figure 6. (a) H2 evolution over the photocatalysts of samples A−E and P25 under different irradiation times. (b) The corresponding average H2 evolution rates of samples A−E. (c) The reusability for photocatalytic H2 generation over the photocatalysts of samples A−E and P25. 20134

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces those of the binary products of sample B (ca. 251.2 μmol g−1 h−1) and E (ca. 368.3 μmol g−1 h−1). That is, based our experiments, the as-prepared TiO2/CuO/Cu ternary thoroughly mesoporous hybrid photocatalysts present a profoundly enhanced photocatalytic H2 production, whose H2 release rate could be enhanced for more than ∼3.5 times with respect to those of the pristine TiO2 thoroughly mesoporous fibers and commercial P25, respectively. To show their stability and reusability, these six photocatalysts are recovered and reused for photocatalytic H2 production under the same condition. As shown in Figure 6c, it seems that there is nearly no loss of H2 evolution rate for samples A−D, whereas the photocatalytic ability of P25 has been evidently declined after three cycles, indicating their robust stability and reusability. To further investigate their reusability for long-term photocatalytic application, as an example of sample C, the structure and phases of the TiO2/CuO/Cu photocatalysts after the photocatalytic reactions for three cycles are investigated, as shown in Figure S6. It suggests that their structures and phases are almost identical before and after the photocatalytic reactions, verifying that the TiO2/CuO/Cu photocatalysts hold a satisfied stability. In order to understand the enhanced and stable photocatalytic ability of the ternary TiO2/CuO/Cu thoroughly mesoporous nanofibers, a proposed diagram and their band energy structures are schematically illustrated in Figures 7a and 7b, respectively.

water/photocatalyst interface.1 Thus, the obtained thoroughly mesoporous structure with satisfied specific surface area is beneficial for the first one, which possesses sufficient reactive sites to improve its photocatalytic efficiency.16,34,42 Additionally, the steady 1D geometry could limit the aggregation of the photocatalysts, which allows them to be long-term serviced.30 More importantly, in the present TiO2/CuO/Cu ternary system, TiO2 and CuO are both introduced to create electrons from their valence bands (VB) to conduction bands (CB) under the Xe lamp irradiation. Because of the fact that the position of the CuO conduction band is much more positive than that of TiO2, the photogenerated electrons should be transferred from TiO2 to the CuO conduction, while the holes would be accumulated at the VB of the TiO2 and CuO. In this way, the recombination of electrons and holes can be avoided. Unfortunately, the photoexcited electrons in CuO cannot directly contribute to the hydrogen production, since the CB edge position of CuO is more positive than the H+/H2 potential (NHE). However, as the continuous accumulation of the excess electrons in the CB of CuO, it would cause a negative shift in its Fermi level, behaving higher electron availability for the interfacial transfer to H+ in solution, which thus facilitate the generation of H2. This endows the CuO with required overvoltage for proton reduction to be worked as the hydrogen formation site. Hence, the CuO in this composite nanofiber acts dual roles to enhance the photocatalytic H2 production: (i) promote the electron transfer from TiO2 to CuO in the heterojunctions for the enhancement of the charge separation; (ii) the cocatalyst behaviors by offering the reduction sites for H2 production. Meanwhile, the physicochemical properties of the ternary hybrids would be further improved by the introduction of Cu0, which could act not only as the reducing centers by trapping excited electrons in the CB of TiO2,24 but also as the unique conductive electron transport “highway”. With respect to the Fermi level of Cu lower than that of p-type CuO, the electrons on the CB of CuO can further migrate to Cu until the systems is equilibration.43,44 That is to say, the presence of Cu prolongs the lifetime and accelerated the transfer speed of photogenerated carriers, conferring the ternary photocatalyst with enhanced photocatalytic efficiency. Overall, the constructed thoroughly mesoporous TiO2/CuO/Cu ternary hybrid photocatalysts can bring the synergistic effect such as suppressed charge recombination, improved interfacial charge transfer, increased photocatalytic reaction centers, and the favorable 1D thoroughly mesoporous nanostructures, which consequently make a significant enhancement on their photocatalytic performances with excellent efficiency and robust stabilities.

4. CONCLUSIONS In summary, we have reported the exploration of ternary hybrid TiO2/CuO/Cu thoroughly mesoporous nanofibers via a foaming-assisted electrospinning strategy. The introduced Cu contents within the initial solutions play a profound effect on the growth of the ternary hybrid fibers, including the phase compositions, porosities, specific surface areas, and light harvest. The ternary TiO2/CuO/Cu thoroughly mesoporous system exhibits a uncompromising stability and a prominent photocatalytic H2 evolution efficiency of ∼851.3 μmol g−1 h−1, which could be significantly enhanced for more than ∼3.5 times as compared to those of the pristine TiO2 counterparts and commercial P25. We mainly attribute the enhanced photocatalytic behaviors to the constructed heterojunctions among the TiO2, CuO, and Cu interfaces, which could favor the strengthened space separation and transfer of the photogenerated charge carriers.

Figure 7. (a) Schematically illustrated mechanism for the possible photocatalytic mechanism of the H2 generation over ternary TiO2/CuO/Cu thoroughly mesoporous nanofibers under xenon lamp irradiation. (b) Schematic illustration of the energy band structures of TiO2/CuO/Cu nanofibers.

According to the semiconductor photocatalytic theory, the total hydrogen production is mainly governed by the following issues. The first is the capacity of adsorption reactants and desorption product. The second is the amount of excited electrons in the 20135

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces

(12) Kumar, S. G.; Devi, L. G. Review on Modified TiO2 Photocatalysis Under UV/visible Light: Selected Results and Related Mechanisms on Interfacial Charge Carrier Transfer Dynamics. J. Phys. Chem. A 2011, 115 (46), 13211−13241. (13) Zhang, R.; Elzatahry, A. A.; Al-Deyab, S. S.; Zhao, D. Mesoporous Titania: From Synthesis to Application. Nano Today 2012, 7 (4), 344− 366. (14) Daghrir, R.; Drogui, P.; Robert, D. Modified TiO2 for Environmental Photocatalytic Applications: a Review. Ind. Eng. Chem. Res. 2013, 52 (10), 3581−3599. (15) Zhang, W.; Zhu, R.; Ke, L.; Liu, X.; Liu, B.; Ramakrishna, S. Anatase Mesoporous TiO2 Nanofibers with High Surface Area for SolidState Dye-Sensitized Solar Cells. Small 2010, 6 (19), 2176−2182. (16) Zhang, X.; Thavasi, V.; Mhaisalkar, S.; Ramakrishna, S. Novel Hollow Mesoporous 1D TiO2 Nanofibers as Photovoltaic and Photocatalytic Materials. Nanoscale 2012, 4 (5), 1707−1716. (17) Liu, H.; Hou, H.; Gao, F.; Yao, X.; Yang, W. Tailored Fabrication of BiVO4 Thoroughly Mesoporous Nanofibers and Their Visible Light Photocatalytic Activities. ACS Appl. Mater. Interfaces 2016, 8 (3), 1929− 1936. (18) Hou, H.; Gao, F.; Wang, L.; Shang, M.; Yang, Z.; Zheng, J.; Yang, W. Superior Thoroughly Mesoporous Ternary Hybrid Photocatalysts of TiO2/WO3/g-C3N4 Nanofibers for Visible-light-driven Hydrogen Evolution. J. Mater. Chem. A 2016, 4 (17), 6276−6281. (19) Huang, H.; He, Y.; Du, X.; Chu, P. K.; Zhang, Y. A General and Facile Approach to Heterostructured Core/Shell BiVO4/BiOI p−n Junction: Room-Temperature in Situ Assembly and Highly Boosted Visible-Light Photocatalysis. ACS Sustainable Chem. Eng. 2015, 3 (12), 3262−3273. (20) Huang, H.; Han, X.; Li, X.; Wang, S.; Chu, P. K.; Zhang, Y. Fabrication of Multiple Heterojunctions with Tunable Visible-lightactive Photocatalytic Reactivity in BiOBr−BiOI Full-range Composites Based on Microstructure Modulation and Band Structures. ACS Appl. Mater. Interfaces 2015, 7 (1), 482−492. (21) Huang, H.; Xiao, K.; He, Y.; Zhang, T.; Dong, F.; Du, X.; Zhang, Y. In situ Assembly of BiOI@ Bi12O17Cl2 pn Junction: Charge Induced Unique Front-lateral Surfaces Coupling Heterostructure with High Exposure of BiOI {001} Active Facets for Robust and Nonselective Photocatalysis. Appl. Catal., B 2016, 199, 75−76. (22) Gombac, V.; Sordelli, L.; Montini, T.; Delgado, J. J.; Adamski, A.; Adami, G.; Cargnello, M.; Bernal, S.; Fornasiero, P. CuOx−TiO2 Photocatalysts for H2 Production from Ethanol and Glycerol Solutions. J. Phys. Chem. A 2009, 114 (11), 3916−3925. (23) In, S. I.; Vaughn, D. D., II; Schaak, R. E. Hybrid CuO-TiO2‑xNx Hollow Nanocubes for Photocatalytic Conversion of CO2 into Methane under Solar Irradiation. Angew. Chem., Int. Ed. 2012, 51 (16), 3915− 3918. (24) Xing, J.; Chen, Z. P.; Xiao, F. Y.; Ma, X. Y.; Wen, C. Z.; Li, Z.; Yang, H. G. Cu-Cu2O-TiO2 Nanojunction Systems with an Unusual Electron-hole Transportation Pathway and Enhanced Photocatalytic Properties. Chem. - Asian J. 2013, 8 (6), 1265−1270. (25) Mondal, I.; Pal, U. Manifestation of MOF Templated Cu/CuO@ TiO2 Nanocomposite for Synergistic Hydrogen Production. Phys. Chem. Chem. Phys. 2016, 18 (6), 4780−4788. (26) Sarkar, S.; Zou, J.; Liu, J.; Xu, C.; An, L.; Zhai, L. Polymer-derived Ceramic Composite Fibers with Aligned Pristine Multiwalled Carbon Nanotubes. ACS Appl. Mater. Interfaces 2010, 2 (4), 1150−1156. (27) Luo, C.; Stoyanov, S. D.; Stride, E.; Pelan, E.; Edirisinghe, M. Electrospinning Versus Fibre Production Methods: from Specifics to Technological Convergence. Chem. Soc. Rev. 2012, 41 (13), 4708−4735. (28) Zhang, C.-L.; Yu, S.-H. Nanoparticles Meet Electrospinning: Recent Advances and Future Prospects. Chem. Soc. Rev. 2014, 43 (13), 4423−4448. (29) Valente, T. A.; Silva, D. M.; Gomes, P. S.; Fernandes, M. H.; Santos, J. D.; Sencadas, V. Effect of Sterilization Methods on Electrospun poly (lactic acid) (PLA) Fibre Alignment for Biomedical Applications. ACS Appl. Mater. Interfaces 2016, 8 (5), 3241−3249. (30) Lee, S. S.; Bai, H.; Liu, Z.; Sun, D. D. Novel-structured Electrospun TiO2/CuO Composite Nanofibers for High Efficient

It is expected that current work might provide a novel and facile strategy for exploring superior photocatalysts to be used in photocatalytic clean energy production.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.6b06644. Typical SEM images, element scanning patterns, TEM images, XRD patterns, and GC patterns of the obtained products (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail [email protected]; Tel +86-574-87080966; Fax +86-574-87081221 (W.Y.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by National Natural Science Foundation of China (NSFC, Grant No. 51372122, 51372123, and 51572133), Postdoctoral Science Foundation of China (No. 2015M581966), and Natural Science Foundation of Ningbo Municipal Government (Grant No. 2016A610102 and 2016A610108).



REFERENCES

(1) Chen, X.; Shen, S.; Guo, L. Semiconductor-based Photocatalytic Hydrogen Generation. Chem. Rev. 2010, 110 (11), 6503−6570. (2) Hisatomi, T.; Kubota, J.; Domen, K. Recent Advances in Semiconductors for Photocatalytic and Photoelectrochemical Water Splitting. Chem. Soc. Rev. 2014, 43 (22), 7520−7535. (3) Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S. T.; Zhong, J.; Kang, Z. Metal-free Efficient Photocatalyst for Stable Visible Water Splitting Via a Two-electron Pathway. Science 2015, 347 (6225), 970−974. (4) Wang, Q.; Hisatomi, T.; Jia, Q.; Tokudome, H.; Zhong, M.; Wang, C.; Pan, Z.; Takata, T.; Nakabayashi, M.; Shibata, N. Scalable Water Splitting on Particulate Photocatalyst Sheets with a Solar-to-hydrogen Energy Conversion Efficiency Exceeding 1%. Nat. Mater. 2016, 15 (6), 611−615. (5) Yang, D.; Liu, H.; Zheng, Z.; Yuan, Y.; Zhao, J. C.; Waclawik, E. R.; Ke, X.; Zhu, H. An Efficient Pphotocatalyst Structure: TiO2(B) Nnanofibers with a Shell of Anatase Nanocrystals. J. Am. Chem. Soc. 2009, 131 (49), 17885−17893. (6) Chen, X.; Liu, L.; Yu, P. Y.; Mao, S. S. Increasing Solar Absorption for Photocatalysis with Black Hydrogenated Titanium Dioxide Nanocrystals. Science 2011, 331 (6018), 746−750. (7) Zhang, L.; Jing, D.; She, X.; Liu, H.; Yang, D.; Lu, Y.; Li, J.; Zheng, Z.; Guo, L. Heterojunctions in g-C3N4/TiO2(B) Nanofibres with Exposed (001) Plane and Enhanced Visible-light Photoactivity. J. Mater. Chem. A 2014, 2 (7), 2071−2078. (8) Jo, W.-K.; Sivakumar Natarajan, T. Facile Synthesis of Novel Redox-Mediator-free Direct Z-Scheme CaIn2S4 Marigold-Flower-like/ TiO2 Photocatalysts with Superior Photocatalytic Efficiency. ACS Appl. Mater. Interfaces 2015, 7 (31), 17138−17154. (9) Li, H.; Bian, Z.; Zhu, J.; Huo, Y.; Li, H.; Lu, Y. Mesoporous Au/ TiO2 Nanocomposites with Enhanced Photocatalytic activity. J. Am. Chem. Soc. 2007, 129 (15), 4538−4539. (10) Li, H.; Wang, D.; Wang, P.; Fan, H.; Xie, T. Synthesis and Studies of the Visible-Light Photocatalytic Properties of Near-Monodisperse BiDoped TiO2 Nanospheres. Chem. - Eur. J. 2009, 15 (45), 12521−12527. (11) Wang, X.; Yu, J. C.; Ho, C.; Hou, Y.; Fu, X. Photocatalytic Activity of a Hierarchically Macro/mesoporous Titania. Langmuir 2005, 21 (6), 2552−2559. 20136

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137

Research Article

ACS Applied Materials & Interfaces Photocatalytic Cogeneration of Clean Water and Energy from Dye Wastewater. Water Res. 2013, 47 (12), 4059−4073. (31) Bandara, J.; Udawatta, C.; Rajapakse, C. Highly Stable CuO Incorporated TiO2 Catalyst for Photocatalytic Hydrogen Production from H2O. Photochem. Photobiol. Sci. 2005, 4 (11), 857−861. (32) Li, G.; Dimitrijevic, N. M.; Chen, L.; Rajh, T.; Gray, K. A. Role of Surface/Interfacial Cu2+ Sites in the Photocatalytic Activity of Coupled CuO-TiO2 Nanocomposites. J. Phys. Chem. C 2008, 112 (48), 19040− 19044. (33) Chen, W. T.; Jovic, V.; Sun-Waterhouse, D.; Idriss, H.; Waterhouse, G. I. The Role of CuO in Promoting Photocatalytic Hydrogen Production over TiO2. Int. J. Hydrogen Energy 2013, 38 (35), 15036−15048. (34) Hou, H.; Wang, L.; Gao, F.; Wei, G.; Tang, B.; Yang, W.; Wu, T. General Strategy for Fabricating Thoroughly Mesoporous Nanofibers. J. Am. Chem. Soc. 2014, 136 (48), 16716−16719. (35) Wu, H. B.; Hng, H. H.; Lou, X. W. D. Direct Synthesis of Anatase TiO2 Nanowires with Enhanced Photocatalytic Activity. Adv. Mater. 2012, 24 (19), 2567−2571. (36) Chang, Y.; Zeng, H. C. Controlled Synthesis and self-assembly of Single-crystalline CuO Nanorods and Nanoribbons. Cryst. Growth Des. 2004, 4 (2), 397−402. (37) Song, Z.; Hrbek, J.; Osgood, R. Formation of TiO2 Nanoparticles by Reactive-layer-assisted Deposition and Characterization by XPS and STM. Nano Lett. 2005, 5 (7), 1327−1332. (38) Xu, S.; Du, A. J.; Liu, J.; Ng, J.; Sun, D. D. Highly Efficient CuO Incorporated TiO2 Nanotube Photocatalyst for Hydrogen Production from Water. Int. J. Hydrogen Energy 2011, 36 (11), 6560−6568. (39) Xu, H.; Li, H.; Wu, C.; Chu, J.; Yan, Y.; Shu, H.; Gu, Z. Preparation, Characterization and Photocatalytic Properties of Culoaded BiVO4. J. Hazard. Mater. 2008, 153 (1), 877−884. (40) Sing, K. S. Physisorption of Nitrogen by Porous Materials. J. Porous Mater. 1995, 2 (1), 5−8. (41) Storck, S.; Bretinger, H.; Maier, W. F. Characterization of Microand Mesoporous Solids by Physisorption Methods and Pore-size analysis. Appl. Catal., A 1998, 174 (1), 137−146. (42) Hou, H.; Wang, L.; Gao, F.; Wei, G.; Zheng, J.; Tang, B.; Yang, W. Fabrication of Porous Titanium Dioxide Fibers and Their Photocatalytic Activity for Hydrogen Evolution. Int. J. Hydrogen Energy 2014, 39 (13), 6837−6844. (43) Dvoranova, D.; Brezova, V.; Mazúr, M.; Malati, M. A. Investigations of Metal-doped Titanium Dioxide Photocatalysts. Appl. Catal., B 2002, 37 (2), 91−105. (44) Abdulla-Al-Mamun, M.; Kusumoto, Y.; Ahmmad, B.; Islam, M. S. Photocatalytic Cancer (HeLa) Cell-killing Enhanced with Cu−TiO2 Nanocomposite. Top. Catal. 2010, 53 (7−10), 571−577.

20137

DOI: 10.1021/acsami.6b06644 ACS Appl. Mater. Interfaces 2016, 8, 20128−20137