Weiwei Xia†
a,
Haoyu Qian†a,
Xianghua Zeng*ab,
Jiawei Suna,
Pengdi Wanga,
Min Luoa and
Jing Donga
aCollege of Physics Science and Technology, Institute of Optoelectronic Technology, Yangzhou University, Yangzhou 225002, P. R. China. E-mail: xhzeng@yzu.edu.cn; wwxia@yzu.edu.cn
bCollege of Electrical, Energy and Power Engineering, Yangzhou 225127, P. R. China
First published on 29th July 2019
Semiconductor heterostructures are regarded as an efficient way to improve the photocurrent in photoelectrochemical cell-type (PEC) photodetectors. To better utilize solar energy, TiO2@Sn3O4 arrays vertically aligned on carbon fiber papers were synthesized via a hydrothermal route with a two-step method and used as photoanodes in a self-powered photoelectrochemical cell-type (PEC) photodetector under visible light. TiO2@Sn3O4 heterostructures exhibit a stable photocurrent of 180 μA, which is a 4-fold increase with respect to that of the Sn3O4 nanoflakes on carbon paper, and a two-order increase with respect to that of the TiO2 NRs arrays. The evolution of hydrogen according to the photo-catalytic water-splitting process showed that Sn3O4/TiO2 heterostructures have a good photocatalytic hydrogen evolution activity with the rate of 5.23 μmol h−1, which is significantly larger than that of Sn3O4 nanoflakes (0.40 μmol h−1) and TiO2 nanorods (1.13 μmol h−1). Furthermore, the mechanism behind this was discussed. The detector has reproducible and flexible properties, as well as an enhanced photosensitive performance.
Among their methods, heterogeneous photocatalysts were regarded as an effective way to improve photocatalytic activity, since heterojunctions can promote charge migration across both the semiconductor/electrolyte and the semiconductor/photoelectrode interface effectively separating photogenerated electrons and holes.8–10 P-SnO/n-TiO2 composite electrodes revealed a better PEC performance as reported by Naeem et al.11 And tree-like TiO2 architectures assembled with CdS and reduced graphene oxide can improve photoelectrochemical performance as reported by Pathak et al.12
On the other hand, a mixed-valence tin oxide (Sn2+ and Sn4+) has a strong resistance against acidic/alkaline solutions and exhibits a good photocatalytic activity without generating secondary pollutants under irradiation with visible light. It is a desirable material for PEC-type photodetectors and an efficient visible-light-driven photocatalyst.13–16 To improve the photoelectric properties, Sn3O4 nanomaterials combined with others have been used as electrodes and photocatalysts. For example, integrated Sn3O4 nanosheets and rGO nanosheets as hybrid nanostructures were used as high-performance photocatalysts.17 With the combination of visible light active Sn3O4 and ultraviolet (UV) light active TiO2, this composite can be a broad spectrum photocatalytic material from the UV to visible region as well as having enhanced separation of photogenerated electrons and holes,18 therefore people paid more attention to TiO2@Sn3O4 nanostructures. For example, 3D lupinus-like TiO2@Sn3O4 nanostructures on a transparent F-doped SnO2 (FTO) glass substrate were prepared and used as photoanode for PEC water splitting as reported by Zhu et al.19 TiO2@Sn3O4 hybrid nanobelts exhibited markedly enhanced photo-electrochemical (PEC) response, which caused higher photocatalytic hydrogen evolution even without the assistance of Pt as a co-catalyst, and enhanced the degradation ability of organic pollutants under both UV and visible light irradiation.20
Recently, self-powered photodetectors (SPPD) are becoming a promising candidate for application in high-sensitivity and high-speed SPPDs because they do not require batteries as an external power source.21,22 For example, TiO2 based self-powered photochemical UV detectors have been reported, but they are restricted to UV light.23–25 To better use solar light for energy conversion, it is very important to extend the spectrum to the visible region. Although there are some discussions on the TiO2@Sn3O4 electrodes, flexible SPPDs based on TiO2@Sn3O4 haven't been reported yet.
Herein, TiO2 nanorod arrays (NRs) vertically aligned on the carbon fiber papers were synthesized via a hydrothermal route, where TiO2 nanorods have a diameter of about 500–1000 nm; then Sn3O4 nanoflakes (NFs) with a thickness of less than 100 nm and width of 100–300 nm were grown on the surface of TiO2 nanorod. Furthermore, hybrid PEC-type photodetectors were designed using three prepared 3-D (three dimensional) hierarchical arrays, TiO2, Sn3O4 and TiO2@Sn3O4 as photoanodes, respectively. As a visible light SPPD device, TiO2@Sn3O4 heterostructures exhibit a stable photocurrent of 180 μA, which is a 4-fold increase with respect to that of Sn3O4 NFs on carbon paper, and a two-order increase with respect to that of TiO2 NRs on carbon paper. The detector exhibits reproducible and flexible properties, as well as an enhanced photosensitive performance. Finally, the photocatalytic properties of the pure Sn3O4, TiO2 NRs and their combinations have been studied, where the evolution of hydrogen was measured according to the photo-catalytic water-splitting process. The results show that Sn3O4/TiO2 heterostructures have hydrogen evolution activity with the rate of 5.23 μmol h−1, significantly larger than that of others. The mechanism behind has been discussed.
Fig. 1 (a) Schematic of fabrication process and (b) XRD patterns of TiO2 NRs, Sn3O4 NFs and TiO2@Sn3O4 heterostructure arrays grown on carbon fiber papers. |
To observe the morphology and microstructure of the samples, the FESEM images of TiO2 NRs and TiO2@Sn3O4 hierarchical structure are shown in Fig. 2. Fig. 2a–c show FESEM images of TiO2 NRs on carbon fiber paper substrate with different magnifications. With the hydrothermal synthesis, TiO2 NRs with a size of 500–1000 nm grow vertically along the surface of the carbon fiber paper substrate forming nanowire arrays with good uniformity, each TiO2 NR has a square shape. After the growth of Sn3O4 nanoflake branches by the followed hydrothermal method, it can be obviously noted that high-density secondary Sn3O4 nanoflake branches are successfully grown on the surface of the TiO2 nanowires with good uniformity, which lead to a thicker and rougher surface of TiO2 nanowires. The magnified image (Fig. 2f) reveals that Sn3O4 NFs have a thickness of less than 100 nm and width of 100–300 nm. Because of the existence of convenient diffusion pathways, these TiO2@Sn3O4 nanowire array hierarchical structures are extraordinarily accessible to electrolytes during the electrochemical measurement process. In addition, the compact nanoflakes interlaced with each other can also greatly increase the capture of light. Further morphological and structural characterizations of the TiO2@Sn3O4 hierarchical structure are performed using HRTEM. The low-magnification TEM image (Fig. 3a) reveals the typical heterostructured nanowire taken from hierarchical TiO2@Sn3O4 nanowire arrays, which can be evidently observed that the individual TiO2 nanowire with a length of several micrometers is covered by compact Sn3O4 nanoflake branches. A close examination of the exposed profile (inset of Fig. 3a) reveals that thickness of the outer Sn3O4 NFs is around several tens nm. The high resolution TEM examination shown in Fig. 3b reveals a distinct set of visible lattice fringes of 0.33 nm, corresponding to the (111) plane of triclinic Sn3O4 (JCPDS 16-0737). To determine the atomic distribution in TiO2@Sn3O4 hierarchical structure, EDS element mapping analysis of the individual hybrid nanowire was performed (Fig. 3c–f), Ti, Sn and O elements are found clearly in the hierarchical heterostructure of TiO2@Sn3O4 nanowire.
Fig. 2 FESEM images of (a–c) TiO2 NRs and (d–f) TiO2@Sn3O4 hierarchical structure grown on carbon fiber paper. |
Fig. 3 TEM image (a) and HRTEM image (b) of a representative TiO2@Sn3O4 heterostructured nanowire. (c–f) Elemental mapping images of an individual hybrid nanowire. |
Optical properties of the as-prepared TiO2 NRs, Sn3O4 NFs and TiO2@Sn3O4 hierarchical structure grown on carbon fiber paper were investigated by diffuse reflectance UV-vis spectra (ultraviolet-visible spectroscopy) (Fig. 4a). Compared with bare TiO2 NRs, TiO2@Sn3O4 hierarchical structure shows significantly increased light absorption in the visible-light region. The absorption edge of the TiO2 NRs (black line) is located at 408 nm, which indicates the energy gap is 3.04 eV. The absorption band edge of the Sn3O4 NFs (red) is closed to 435 nm (with the energy gap of 2.85 eV), which can absorb the visible light. For the 3D TiO2@Sn3O4 hierarchical structure (blue line), its absorption band edge is at ∼558 nm, showing a significant redshift compared with the bare TiO2 NRs. Therefore, coupling TiO2 NRs with Sn3O4NFs can significantly broaden the range of photo-response, which can result in an enhancement of utilization of solar light and improved PEC performance. XPS measurements were used to investigate the surface composition of the as-synthesized products. Fig. 4b shows the XPS survey spectra of Sn3O4 NFs and TiO2@Sn3O4 hierarchical structure. Chemical state binding energy of Sn 3d in the Sn3O4 NFs and TiO2@Sn3O4 hierarchical structure was carried by XPS measurements as shown in Fig. 4c. Consistent with our previous study, the binding energies of Sn 3d3/2 and Sn 3d5/2 peaks in the pure Sn3O4 NFs locate at 494.64 eV and 486.15 eV, respectively. After growing onto the TiO2 NRs, however, the Sn 3d3/2 and Sn 3d5/2 peaks of TiO2@Sn3O4 hierarchical structure obviously shifted to higher binding energies by 0.46 and 0.47 eV, respectively. Therefore, the observed peak shift indicates that the electrons are injected from Sn3O4 NFs to the TiO2 NRs in TiO2@Sn3O4 hierarchical structure. This result can be further supported by the shift by 0.4 eV of the Ti 2p3/2 peak in Fig. 4d. The TiO2@Sn3O4 hierarchical structure is possibly advantageous because of their electronic band structure.
UPS measurements were carried out to study the surface electron behavior of TiO2 NRs, as shown in Fig. 4e and f. Fig. 4e is a view of the secondary electron edge (SEE) energy corresponding to the left spectra in UPS data. The work function (ϕ) can be by observing the low energy secondary electron cutoff, which is ∼5.20 eV for TiO2 NRs. Fig. 4f is a view of the valence band maximum (VBM) region corresponding to the right spectra in UPS data and the VBM can be extracted from the Fig. 4f, which is about 2.2 eV for TiO2 NRs. The energy band of Sn3O4 NFs has been discussed previously.22 Then, the conduction band and valence band (vs. vacuum) of TiO2 NRs and Sn3O4 NF scan be obtained, as shown in Table 1.
Samples | Eg (eV) | EC (eV vs. vacuum) | EF (eV vs. vacuum) | Ev (eV vs. vacuum) |
---|---|---|---|---|
TiO2 | 3.04 | 4.43 | 5.20 | 7.47 |
Sn3O4 | 2.85 | 3.7 | 3.9 | 6.55 |
The PEC measurements were carried out in a three-electrode configuration in 0.1 M Na2SO4 using Pt wire as a counter electrode, Ag/AgCl in saturated KCl as a reference electrode, and carbon-paper-supported Sn3O4 NFs, TiO2 NRs and TiO2@Sn3O4 hierarchical structures as active photoanodes, respectively, as shown in the Fig. 5a. The incident radiation is switched with an on/off interval of 20 s. Twenty repeated cycles are displayed in Fig. 5b, in which TiO2 NRs display a quite weak photocurrent under visible-light irradiation. The as-obtained Sn3O4 NFs exhibit a much enhanced photocurrent density of ∼40 μA cm−2 in the visible-light spectral region. However, the photocurrent exhibited nearly 40% decrease after 20 cycles. The TiO2@Sn3O4 hierarchical electrode shows the highest photocurrent and a decrease of 30% compared with the initial photocurrent. From the magnified rising and decaying edges of photocurrent shown in Fig. 5c, a fast photo-response can be seen clearly. The rising time (τr, defined as the time to increase from 10% to 90% of the maximum photocurrent) and the decaying time (τd, defined as the time to recover from 90% to 10% of the maximum photocurrent) of TiO2@Sn3O4 hierarchical electrodes are about 0.27 s and 0.87 s, respectively, faster than 0.30 s and 0.92 s for Sn3O4 NFs. The mechanism of reason for enhanced PEC performance will be explained in detail later. PEC measurements that the 3D TiO2@Sn3O4 photoelectrode exhibited a stable photocurrent of 180 μA, which is 4-fold increase with respect to that of Sn3O4 NFs, and two-order increase respect to that of TiO2 NRs. In comparison to the report of ref. 19, the formation of the heterostructure on the photocurrent is much larger without external power source.
Furthermore, the photocatalytic property of the pure Sn3O4, TiO2 NRs and their combinations have been studied under visible light, where the evolution of hydrogen was measured according to the photo-catalytic water-splitting process, as shown in Fig. 5d. The results show that Sn3O4 NFs show negligible photocatalytic hydrogen evolution activity with the rate of 0.40 μmol h−1 cm−2 and the rate of the TiO2 NRs is 1.13 μmol h−1 TiO2 NRs, which is much lower than that of Sn3O4/TiO2 heterostructures of 5.23 μmol h−1, this value is comparable with the results reported by Kodiyath et al.27 Contrary to the photocurrents, the H2 generation of Sn3O4 is less than that of TiO2. Several aspects, such as the density of the photoanode catalysts, electron transportation and carrier's concentration etc., will influence the H2 generation. In the further work we will study them.
To further understand the reasons of the charge transfer process occurring at the interface of photoelectrode/electrolyte, electrochemical impedance spectroscopy (EIS)28,29 was carried out and presented in Fig. 6a and b. The equivalent RC circuit was used to interpret the EIS results (see inset of Fig. 6b). As shown in Fig. 6a and b, each Nyquist plots are composed of one semicircle and a slope line. For deeper analysis, all Nyquist plots display a semicircle at high frequencies whose diameter represents the charge-transfer resistance (Rct), which reflects the electron transfer kinetics of the redox probe at the interface. The slope line at low frequency is related to the diffusion process. The corresponding equivalent circuit is depicted in the inset of Fig. 6b, where Rs denotes the series resistance at the interface of the photoelectrode material and the carbon fiber paper substrate, Cd1 reflects the constant phase element that models capacitance of the double layer, and Zw stands for the Warburg impedance originated from the diffusion process at the electrode surface. After being decorated with Sn3O4 NFs, the value Rct of TiO2@Sn3O4 hierarchical photoelectrode is almost the same, which is smaller than the corresponding value of Sn3O4 NFs. The above result demonstrates that TiO2 NRs provide a conduction path and rapidly transfer the photoelectrons coming from Sn3O4 NFs to the carbon fiber paper substrate along the vertically oriented TiO2 NRs, which is also correlated to the enhanced PEC performance. To obtain deep insights into the enhanced PEC properties of the TiO2@Sn3O4 hierarchical electrodes, intensity modulated photocurrent spectroscopy (IMPS) and intensity modulated photovoltage spectroscopy (IMVS) have been employed to investigate further the electron transport and recombination process, as shown in Fig. 6c and d. The IMPS and IMVS plots display a semicircle in the complex plane under the incident light source was an LED with a wavelength of 564 ± 60 nm. From the IMPS measurement (Fig. 6c), the transport time (τd) of injected electrons from Sn3O4 NFs to TiO2 NRs can be calculated from the following equation:
τd = 1/(2πfd,min), |
τn = 1/(2πfn,min), |
Based on the above measurements and analyses, by combining both the band gap estimated from optical absorption and the conduction band and valence band (vs. vacuum) obtained by UPS, the energy band alignment of TiO2 NRs and Sn3O4 NFs is displayed in Fig. 7a. When Sn3O4 NFs contact with TiO2 NRs to form a heterojunction, the band structure of TiO2@Sn3O4 is reconfigured. The bandgap energy of Sn3O4 (2.85 eV) is smaller than TiO2 (3.04 eV) and both the potentials of VB and CB of Sn3O4 are higher than those of TiO2, so TiO2@Sn3O4 heterostructure belongs to a typical type-II heterojunction. Under visible-light irradiation, photogenerated holes and electrons appear in the VB and CB of Sn3O4 NFs. Due to type-II heterostructure, the photogenerated electrons in Sn3O4 CB were easily injected into the TiO2 CB. Effective separation of photoexcited electron–hole pairs could be accomplished by a longer electron lifetime. To explain the enhanced PEC performance of TiO2@Sn3O4 photoelectrode in the visible-light spectral region, we present the scheme of the electron–hole pair separation and the transfer process in TiO2@Sn3O4 hierarchical photoelectrode (Fig. 7b). Generally, enhanced PEC performance of TiO2@Sn3O4 photoelectrode may be attributed to the reasonable design of TiO2NRs @Sn3O4NFs heterostructure. On one hand, the aligned TiO2 NRs can provide a conduction path and rapidly transfer electrons to carbon fiber paper substrate along the vertically oriented TiO2 NRs. On the other hand, the type-II heterojunction of TiO2@Sn3O4 heterostructure can drive photoexcited electron transfer from Sn3O4 NFs to TiO2 NRs. In one word, for our TiO2@Sn3O4 photoelectrode under the synergistic effects between unique band structures and morphology, which is crucial for the enhancement of PEC performance.
Fig. 7 (a) Proposed energy band alignment and (b) scheme of the electron–hole pair separation and the transfer process in TiO2@Sn3O4 photoelectrode. |
Footnote |
† W. X. and H. Q. contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2019 |