Shuang Shuanga,
Ruitao Lvb,
Zheng Xieac,
Weipeng Wanga,
Xiaoyang Cuia,
Shuai Ninga and
Zhengjun Zhang*b
aSchool of Materials Science and Engineering, State Key Laboratory of New Ceramics and Fine Processing, Tsinghua University, Beijing 100084, People's Republic of China
bSchool of Materials Science and Engineering, Key Laboratory of Advanced Materials, Tsinghua University, Beijing 100084, People's Republic of China. E-mail: zjzhang@mail.tsinghua.edu.cn; zjzhang@tsinghua.edu.cn
cHigh-Tech Institute of Xi'an, Xi'an 710025, China
First published on 16th December 2015
Vertically aligned α-Fe2O3 nanopillar arrays (NPAs) were fabricated by thermally oxidizing Fe NPAs on Si, quartz and F-doped SnO2 (FTO) substrates prepared by glancing angle e-beam deposition (GLAD). The photocatalytic activity of these NPAs was evaluated by measuring the photodegradation of crystal violet (CV) and methyl orange (MO) in the presence of H2O2 under visible light irradiation. Moreover, the photoelectrochemical (PEC) performance was also studied. Typically the sample oxidized at 400 °C exhibits both the highest degradation efficiency and photocurrent density compared with those oxidized at other temperatures (e.g. 300 °C, 350 °C, 450 °C, 500 °C). This phenomenon might be attributed to a trade-off between two opposite effects. On the one hand, with the increase of the oxidation temperature, the improvement of NPAs' crystallinity will enhance the photocatalytic performance accordingly. On the other hand, increasing oxidation temperature may cause the reduction of oxygen vacancies on the NPAs' surface, which are regarded as the photoreaction active sites. This will thus degrade the photocatalytic performance.
Hematite (α-Fe2O3), an abundant environmentally friendly n-type semiconductor, has been widely used as catalysts,4 pigment,5 gas sensors.6 It has a band gap of ∼2.1 eV, and works in the visible light region with a high theoretical solar-to-hydrogen efficiency of ∼17% and good stability against photocorrosion.7 Therefore, α-Fe2O3 has attracted tremendous interests in recent years.8–15 However, a drawback of α-Fe2O3 catalyst is the short hole diffusion length (2–4 nm) due to its poor minority charge carrier mobility (0.2 cm2 V−1 s−1), which results in a high recombination rate of electrons and holes, a very short excited state lifetime (∼10 ps), and poor electrical conductivity.16,17 To address or to overcome this issue, a lot of efforts have been paid to synthesize nanostructures of hematite, such as nanotubes, nanoparticles, nanocubes, nanowires, nanofibers, nanorods, and hierarchical structures.8–15,18 Besides morphology control, other ways like doping, constructing heterojunctions, surface functionalization of nanostructures, etc., can also be used to optimize its photocatalytic performance.9,19–22
Photocatalysts in previous reports are usually in the form of powders, which is hard to handle and collect. Here we report the synthesis of vertically aligned hematite nanopillar arrays on specific substrates to form self-standing structures, which are much easier to recycle. Furthermore, they do prevent the occurrence of secondary pollution which means more than recycles. The α-Fe2O3 arrays are fabricated by thermally oxidizing Fe nanopillars using a glancing angle deposition (GLAD) technique. The PEC and degradation performance under visible light radiation are evaluated. The influence of the oxidization temperature on the growth of hematite nanopillars and their catalytic performance are also investigated.
The steady sate current density (j–V) and electrochemical impedance spectroscopy (EIS) measurements were carried out by an electrochemistry workstation (CHI 660D, Chenhua instrument). The nanostructured films were used as the working electrode, an Ag/AgCl electrode (saturated KCl) and Pt sheet were used as the reference and counter electrodes, respectively. The working electrode was illuminated with a 300 W Xe lamp. An ultraviolet filter was placed between the light source and the quartz cell to cut off the UV light in wavelength <420 nm. Photocurrent densities were measured in the light on–off process with a pulse of 30 s under visible light illumination (200 mW cm−2) at 0.4 V bias vs. Ag/AgCl electrode.
The photocatalytic activity of the hematite nanopillar arrays was evaluated by the photodegradation of crystal violet (CV) and methyl orange (MO) photodegradation with the light source of 300 W Xe lamp at ambient temperature. The sample on quartz substrate (15 mm × 15 mm) was placed in a quartz cell containing 5 mL of CV (5 μM) and MO (5 μM). Prior to light irradiation, the photocatalyst was immersed into solution in the dark room for 30 min to reach an adsorption/desorption equilibrium, then 12 μL of 30 wt% H2O2 was added to the solution, and the Xe lamp was turned on for different time spans. After that, the UV-vis absorbance spectra of the dye aqueous solution were obtained to measure the dye concentration after photodegradation.
Fig. 2 shows the XRD patterns of the as-prepared samples. The XRD pattern of the Fe sample is well matched with the standard spectrum of Fe (JCPDS 50-1275). When the oxidation temperature are 300 and 350 °C, the main phase is α-Fe2O3 (JCPDS 33-0664) with a little diffraction peak located at 30.1°, which can be ascribed to the (220) plane of γ-Fe2O3 (JCPDS 39-1346). This indicates that a few amount of γ-Fe2O3 co-existing in the sample. When the annealing temperature is higher than 350 °C, the phase will be totally α-Fe2O3.
Fig. 3 shows the Raman spectra of the as-deposited film and films oxidized at temperatures ranging from 300 to 500 °C, measured at room-temperature. It can be seen that, all samples oxidized at temperatures above 300 °C demonstrate the characteristic peaks corresponding to α-Fe2O3, i.e. peaks at 226 and 493 cm−1 of the A1g mode, and peaks at 295, 411 and 610 cm−1 of the Eg mode.23,24 The peak centered at 1318 cm−1 may be caused by the two-magnon scattering due to the interaction between two magnons created on antiparallel close spin sites. It is noticed that the films oxidized at low temperatures (e.g. 300 °C, 350 °C) exhibited broader and lower intensity Raman peaks, whilst the films oxidized at higher temperatures exhibited sharper and higher intense Raman peaks because of better crystallinity.25
Fig. 4 shows XPS spectra of the films oxidized at 300, 400 and 500 °C, where the binding energies were corrected by referencing the C 1s line to 284.6 eV. The peaks located at ∼710 and ∼724 eV are in good agreement with the Fe 2p3/2 and Fe 2p1/2 binding energy of Fe3+.26 The little shakeup satellite structure between two main peaks is the fingerprint of the electronic structure of Fe3+.27 Based on Raman and XPS analysis, it can be known that Fe nanopillars are oxidized into hematite nanopillars by the oxidization treatment at 300–500 °C.
The morphology and structure of individual nanopillars after oxidization were also investigated by TEM analysis. Fig. 5(a)–(d) shows typical low-magnification and high-magnification HRTEM images of the nanopillars oxidized at 300 and 400 °C, respectively. Insets of Fig. 5(b) and (d) show their corresponding selected-area electron diffraction (SAED) patterns. Patterns are of 300 and 400 °C oxidation temperature. We check α-Fe2O3 PDF card with some indices of crystal face to confirm its phase, marked in figures. The SAED rings of the sample oxidized at 400 °C correspond to the (113), (110), (104), (012), and (101) planes of α-Fe2O3, respectively. It can be seen that the nanopillars are of 1D cone-shaped rod-like morphology, ∼50 nm in diameter and ∼200 nm in length. The HRTEM image of 400 °C shows that the lattice fringes are about 0.20 nm, which is in good agreement with the interplanar spacing of α-Fe2O3 (113) planes. Similarly, the 0.27 nm lattice fringes of 300 °C sample can be attributed to the α-Fe2O3 (104) plane. It is worthwhile to mention that diffraction rings of α-Fe2O3 become more distinguishable at higher oxidization temperatures (e.g. 400 °C), which is in good agreement with the Raman analysis.
PEC experiments were performed to evaluate the photoelectrochemical property of these oxidized samples in a three-electrode cell, using the oxidized samples as the working electrode with an exposed area of 1.5 cm2 under visible light irradiation. Here 0.01 M KOH aqueous solution was used as the electrolyte. Fig. 6(a) shows the photocurrent density measured for these samples with light on and off, at a bias potential of +0.4 V, from which photo-generated electron–hole separation is clearly observed.28 It can be seen from Fig. 6(a) that for samples oxidized at 300, 350 and 400 °C, an increase in the photocurrent was observed for samples. For samples oxidized at >400 °C, i.e., 450 and 500 °C, a monotonic decrease in the photocurrent density was observed. The largest photocurrent density (i.e. 1.4 mA cm−2) was obtained with the sample oxidized at 400 °C, suggesting a higher efficiency in the separation of electrons and holes and a better photocatalytic activity.29 Fig. 6(b) shows the photocurrent density of the sample oxidized at 400 °C, as a function of time with the light on and off. It can be seen that the photocurrent density is quite stable and repeatable.
Fig. 6(c) shows the photocurrent density of these samples measured at various bias potentials ranging from −0.2 V to 0.4 V. The open circuit potential (Voc) relates closely to the difference in the apparent Fermi level between α-Fe2O3 and the reference electrode,30 is dependent on the oxidation temperature. From Fig. 6(c), it can be seen that the sample oxidized at 400 °C has a minimum of Voc of ∼−0.049 V, while the sample oxidized at 500 °C has a maximum value of ∼0.183 V. It is well known that a more negative Voc means better electron and hole separation in semiconductors.31–33 This can be responsible partially for the largest and smallest photocurrent density observed for the two samples.
EIS measurements were conducted for all samples at frequencies ranging from 0.1 Hz to 100 kHz in 0.01 M KOH electrolyte at open circuit voltage conditions under visible light irradiation. An equivalent circuit was used to fit the Nyquist plots, as shown in Fig. 6(d), where the Rs, C1, R1, Rct, and C2 are the series resistance, the capacitance of α-Fe2O3 and the electrolyte, the resistance of the holes trapped at surface states, the resistance of holes transferred to the electrolyte through surface states and the capacitance of surface states, respectively. All above values are obtained and listed in Table 1. The specific values of C1 and C2 for different oxidation temperatures are basically in the same order of magnitude with little change, respectively. One sees that all samples exhibited a similar Rs with a value around 0.5 Ω. While Rct is normally larger than R1 two magnitude which indicates the former plays a more important role in circuit. Though values of Rct on 400, 450 and 500 °C are close, R1 increases from 349.5 to 3092 Ω gradually, which also leads to reduction on photocurrent.34–36 Based on above analysis it is concluded that the oxidation temperature of 400 °C owns the maximum photocurrent, which is also confirmed by the PEC results.
| Oxidation temperature (°C) | 300 | 350 | 400 | 450 | 500 |
| Rs (Ω) | 0.52 | 0.54 | 0.51 | 0.51 | 0.52 |
| C1 (F) | 1.2 × 10−5 | 3.1 × 10−6 | 1.9 × 10−6 | 1.3 × 10−6 | 1.1 × 10−6 |
| R1 (Ω) | 31.3 | 139.7 | 349.5 | 431.1 | 3092 |
| C2 (F) | 8.9 × 10−5 | 4.6 × 10−5 | 9.3 × 10−5 | 8.4 × 10−5 | 6.7 × 10−5 |
| Rct (Ω) | 106 620 |
102 700 |
57 749 |
65 029 |
58 155 |
The photoactivities of the obtained α-Fe2O3 NPAs are investigated by the degradation of CV with H2O2 under visiblelight irradiation, and the degradation processes were tested by recording the characteristic absorbance peak at 584 nm. Fig. 7(a) shows the time course of the decrease in the dye concentration. In absence of α-Fe2O3, the CV is degraded by H2O2 79.4% until 60 min. While the enhancement is observed after the introduction of α-Fe2O3 film; for example, the degradation efficiency of sample oxidized at 400 °C reaches to 90.9% due to the constitution of a Fenton system.37 After data processing, all reaction follows first-order kinetics as shown in Fig. 7(b). The rate constant of 400 °C oxidation sample presented by slope is 1.34 times higher than that in absence of α-Fe2O3.
According to previous research, α-Fe2O3 can be photodecomposed because of the photoreduction of Fe3+ to Fe2+ under light illumination. Once Fe2+ is generated, either it reacts with H2O2 generating hydroxy radicals or it detaches from the oxide surface, leaving a vacancy on the semiconductor surface. Once effective collision of dye molecule and the holes happens, CV and/or CV± ions form. After the reaction between ions and hydroxyl radicals produced by the decomposition of H2O2, CV is degraded completely.38 The very good photostability of the Fe2O3 film strongly suggests that the reaction between surface Fe2+ and H2O2 is kinetically faster than the Fe2+ dissolution reaction in the Fe2O3 catalyst surface, thus preventing occurrence of catalyst photocorrosion. We also supplemented experiment choosing methyl orange (MO) as model dye solution (5 μM) in Fig. S1,† which cannot be degraded under visible lights. According to figure, α-Fe2O3 samples with same quantity of H2O2 show the similar regular in photocatalysis, in which sample oxidized at 400 °C also shows best degradation performance. And we compare the photocatalytic performance of Fe2O3 nanopillar arrays with Fe2O3 powders shown in same picture. The sectional view of powder Fe2O3 sample is shown in Fig. S2.†
The band gap energy (Eg) of semiconductor following the equation:
| αEphoton = K(Ephoton − Eg)1/2, |
| Oxidation temperature (°C) | 300 | 400 | 500 |
| Area ratio | 1.631 | 1.109 | 0.650 |
According to analysis mentioned above, the strong intensity of the Raman peak and small full width at half maximum (FWHM) form progressively with improving oxidation treatment temperature, which means better crystallinity and higher crystallinity can truly enhance the PEC and photodegradation performance of α-Fe2O3 nanostructures.31,42,43 As for why there is an inflexion on property, we look into more information of XPS about oxygen element. The deconvolution peaks display in Fig. 8, O 1s spectrum were observed from 529.0 to 536.0 eV. The low binding energy (BE) component observed at ∼529 eV corresponds to the main lattice oxygen (Olat) forming oxide with iron. The bands around ∼532 eV can be assigned to adsorbed oxygen (Oads) while the peak between them (∼521 eV) indicates the ionization of oxygen species that could allow compensation for some deficiencies (Odef) connected in part to the variations in the concentration of oxygen vacancies in iron oxides.3,44 Based on Table 2, it is obvious to find the area ratio of oxygen vacancy/oxygen lattice decreases step by step from 1.63 to 0.65 when raising the treatment temperature.
![]() | ||
| Fig. 8 XPS spectra of O 1s of α-Fe2O3 oxidized at (a) 300 °C, (b) 400 °C, (c) 500 °C. Red line and black dash lines are the fitted curve and original data, respectively. | ||
Combining former research analysis, oxygen vacancies can trap the photogenerated electrons thus reducing the electron–hole recombination and enhancing photocurrent density and photocatalytic character.45 When higher oxidation temperature was carried out, the photocatalytic activity of the α-Fe2O3 thin films gradually increases due to the improvement of crystallization. However it also reduces oxygen vacancies existing on the surface of oxide, which eventually cuts down active sites and weakens the catalytic performance.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra17894k |
| This journal is © The Royal Society of Chemistry 2016 |