Facet-Mediated Photodegradation of Organic Dye over Hematite Architectures by Visible Light
Xuemei Zhou,Jinyao Lan,Gang Liu,Ke Deng,Yanlian Yang,Guangjun Nie,Jiaguo Yu,Linjie Zhi
DOI: https://doi.org/10.1002/anie.201105028
2011-01-01
Angewandte Chemie International Edition
Abstract:Photodegradation of rhodamine B in the presence of H2O2 by visible light over α-Fe2O3 architectures has been investigated (see picture; left to right: 1D nanorods, 2D nanoplates, 3D nanocubes). A link between the exposed facets of α-Fe2O3 architectures and their photoreactivity is established, following {110}>{012}≫{001}. Structure–reactivity correlations are a central theme in heterogeneous catalysis.1–4 In general, the crystallographic surface structure of a catalyst is determined by its exposed facets, and the enclosed facets of a particle-like catalyst in turn determine its geometric shape as well as catalytic properties. Tuning the shape of catalysts, therefore, is essential in developing new catalysts and modifying existing ones with desirable reactivity, selectivity, and stability. Indeed, great advances have been achieved on model catalysts, and insights into the structure–reactivity correlations are crucial not only for our understanding of catalytic processes, but also for generating new concepts to guide the rational design of practical catalysts.5–8 In recent years, significant attention has been directed towards using solar-driven photocatalysts to degrade aqueous organic pollutants (for example, azo dyes).9–13 In view of being naturally abundant and environmentally benign, iron oxides show great promise. Among iron oxides, hematite (α-Fe2O3) is the most thermodynamically stable semiconductor (Eg=2.1–2.2 eV) that can absorb visible light, that is, a substantial fraction of the solar spectrum. α-Fe2O3 has a wide range of applications, such as light-induced water splitting,14 solar cells,15 lithium ion batteries,16 and biotechnology.17 Most studies to date have been carried out on powder substrates in which particle shapes are inherently not well-defined, making it difficult to explore the structure–reactivity correlations. While the vast majority of studies on the impact of particle shape on photocatalytic reactivity are limited to titanium dioxide (TiO2),18–27 far less information is available regarding the shape effects on other photocatalysts, including iron oxides. In this regard, systematic studies on heterogeneous photo-Fenton catalysis, a technologically promising process in wastewater treatment, by iron-bearing nanocatalysts with particular shapes are still lacking.28, 29 Over the past decade, size, shape, and architecture control of low-dimensional nanomaterials (for example, nanodots, nanorods, and nanosheets) with unusual properties has seen rapid growth.30–36 For example, in one-dimensional (1D) anisotropic nanostructures, it is possible to enhance the photoreactivity by tuning the direction and path of photogenerated charge carriers through quantum confinement while minimizing the e−–h+ recombination.37, 38 Herein, we investigate visible-light-induced photodegradation of model dye rhodamine B (RhB) in the presence of hydrogen peroxide (H2O2) over hematite architectures, namely 1D nanorods, 2D nanoplates, and 3D nanocubes. Herein we use “architectures” to describe hematite nanostructures that can be assembled by nano-building units by oriented attachment.35, 36, 39 To the best of our knowledge, this is the first study to investigate heterogeneous photo-Fenton catalysis by nanocatalysts with well-defined architectures. The detailed synthesis of α-Fe2O3 architectures39–42 is described in the Experimental Section and the Supporting Information. The structure, stoichiometry, and oxidation state of the as-prepared α-Fe2O3 architectures were characterized using X-ray diffraction (XRD; Supporting Information, Figure S1a), micro-Raman (Supporting Information, Figure S1b), and high-resolution X-ray photoelectron spectroscopy (XPS; Supporting Information, Figure S1c,d), and results demonstrate that all samples are pure α-Fe2O3 with a rhombohedral hexagonal phase (space group Rc). The morphology and crystallinity of as-prepared α-Fe2O3 architectures were analyzed using transmission electron microscopy (TEM), high-resolution TEM (HRTEM), scanning electron microscopy (SEM), and atomic force microscopy (AFM). A TEM image for 2D nanoplates is shown in Figure 1 a. All of the nanoplates display a well-defined hexagonal shape. Based on a SEM (Supporting Information, Figure S2a) and TEM analysis (Figure 1 a; Supporting Information, Figure S2b), the width and thickness of the plates is determined to be (208.5±27.3) and (14.6±2.6) nm, respectively. A representative HRTEM image (Figure 1 b) and fast Fourier transforms (FFT; inset in Figure 1 b) show the lattice fringe to be 0.25 nm, which is consistent with (110), (−120), and (−210) planes, respectively. Thus, the resulting basal plane is (001). Vertically aligned plates that are frequently observed (Supporting Information, Figure S2b,c) are wedge-shaped and the lateral facets are ascribed to {102}.42 A TEM image (Figure 1 c) shows that α-Fe2O3 3D nanocubes appear to be square in shape with an average size of (22.0±3.2) nm. A HRTEM image (Figure 1 d) and corresponding FFT (inset in Figure 1 d) indicate that the lattice fringe is 0.36 nm. A close look at this cube reveals that the shape is pseudocubic and one of dihedral angles between adjacent lateral facets is 86°. The above data demonstrate that the cubes are enclosed by {012}, {102}, and {112} facets.43 The TEM image in Figure 1 e reveals that the majority of 1D nanorods exhibit nearly faceted ends with a needle shape. Statistical analysis of dimensional distributions of rods based on TEM and AFM (Supporting Information, Figure S2d) indicates that the average length, width, and height is (40.3±8.7), (10.9±5.0), and (5.0±1.2) nm, respectively. A HRTEM image (Figure 1 f) and FFT pattern (inset in Figure 1 f) indicate that three sets of lattice fringes (0.25 nm) fit well to the α-Fe2O3 (110), (−120), and (−210) planes, respectively. Therefore, the basal plane for α-Fe2O3 nanorods is (001). The physical parameters of the as-prepared samples, including dimensions, specific surface areas, and dominant facets are summarized in Table 1. On the basis of the above analysis, 1D nanorods, 2D nanoplates, and 3D nanocubes offer crystallographically distinct facets that are expected to be photocatalytically different. Representative morphologies and structures of α-Fe2O3 architectures. a) TEM image and b) HRTEM image of 2D α-Fe2O3 nanoplates. Insets: FFT pattern and drawing of a plate. c) TEM image and d) HRTEM image of 3D α-Fe2O3 nanocubes. Insets: FFT pattern and drawing of a cube. e) TEM image and f) HRTEM image of 1D α-Fe2O3 nanorods. Insets: FFT pattern and drawing of a rod. Sample Length [nm] Width [nm] Height [nm] SSA[a] [m2 g−1] Dominant facets Reaction rate constants k[b] [×10−3 min−1] Normalized rate constants ks[c] [×10−4 min−1 L m−2] nanoplates – 208.5±27.3 14.6±2.6 30.1±4.1 {001} 4.38±0.35 7.28±1.14 nanocubes 22.0±3.2 22.0±3.2 22.0±3.2 51.8±8.0 {012} 28.20±1.70 27.2±4.5 nanorods 40.3±8.7 10.9±5.0 5.0±1.2 119.8±23.3 {001}{110} 73.26±4.96 30.6±6.3 Photodegradation of RhB on α-Fe2O3 with different architectures under visible-light illumination in the presence of H2O2 additive. (Blank: photolysis in the presence of H2O2 only.) Reaction conditions: RhB concentration 2×10−5 M, catalyst concentration 0.2 g L−1, H2O2 molar concentration 50 mM, initial pH 6.8, 300 W Xe-lamp (λ>420 nm) with an average light intensity of 150 mW cm−2. where h is Planck’s constant, c is the velocity of light, NA is Avogadro’s constant, ΔnRhB is the difference of RhB concentration before and after photoreaction, Φ is light intensity, A is illuminated area, and λ is the wavelength of light. ξ at λ=420 nm is 0.013 %, 0.0083 %, and 0.0072 % for nanorods, nanocubes, and nanoplates, respectively. The trend of ξ values essentially reflects that of the reaction rate constant k from above. Herein, ξ is estimated as the lower limit of photonic efficiency, taking into account the following factors:50 First, the incident photons can be scattered by the suspended particles in an aqueous solution; second, the recombination rate of photogenerated electrons and holes is fast, and the annihilated charge carriers are not involved in relevant reaction processes. With regard to reaction products concerning RhB degradation, such as total organic carbon (TOC), intermediates, and inorganic mineralization species, detailed analyses are given in the Supporting Information (Tables S1, S2 and Figure S7–S9). It is generally accepted that the reactivity of a photocatalyst is determined by its surface atomic and electronic structure.10 The electronic structures of the as-prepared samples were investigated by diffuse-reflectance UV/Vis (DRUV/Vis) spectroscopy and high-resolution XPS. The absorption spectra converted to the Kubelka–Munk function (Supporting Information, Figure S6a) shows that all α-Fe2O3 samples share a comparable absorption edge around 550–600 nm. The band gap is determined to be 2.22, 2.23, 2.20 eV for nanoplates, nanocubes, and nanorods, respectively. The valence band spectra (Supporting Information, Figure S6b) illustrate that the line shape and the width of the band structure are almost identical. Regardless of the role of electronic structures in the heterogeneous photo-Fenton process, the as-prepared hematite architectures show comparable electronic structures, and the differences in the observed photocatalytic performance could be rationalized by the inherent variations in crystal facets exposed by α-Fe2O3 architectures.24 Table 1 lists the dominant facets with large-percentage enclosed α-Fe2O3 architectures. The nanoplates mainly display the {001} facet, which presents either a single iron or single oxygen termination layer on the surface. For {001} facets with an iron termination layer, the density of low-coordinate surface iron cations is 4.6 atoms nm−2. The above two coexisting but crystallographically distinct terminations of α-Fe2O3 {001} were observed under ultrahigh vacuum51 and in aqueous solution.52 Figure 3 a depicts the atomic model of {001} facet with an oxygen and iron termination layer, respectively. Given coexisting oxygen and iron of termination layers on the {001} facet, the fraction of surface low-coordinate iron cations on the {001} facet is further limited. On the other hand, both {012} (Figure 3 b) and {110} facets (Figure 3 c) display ridge-and-valley topography. This particular topography can expose a greater number of low-coordinate surface iron cations than the {001} facet with mixed oxygen and iron terminations. For example, the density of iron bonded to five oxygen anions (denoted as Fe5c) on {110} and {012} is 10.1 and 7.3 atoms nm−2, respectively. Unlike bulk iron cations, which are octahedrally coordinated by six oxygen anions, the coordinatively unsaturated iron cations can offer catalytically active sites for dye adsorbates, and the observed photoreactivity differences are a direct consequence of the availability of surface iron sites, with the reactivity following {110}>{012}≫{001}. The reaction mechanism for heterogeneous photo-Fenton catalysis involving photosensitization and catalytically active sites is discussed in detail in the Supporting Information. In contrast to previous studies focusing on amorphous and nonstoichiometric FeOx catalysts in heterogeneous photo-Fenton catalysis,28, 29, 54, 55 our findings demonstrate that the construction of photocatalysts on the nanoscale with particular architectures and associated facets is crucial when considering the design of high-performance nanocatalysts. Side views of surface terminations of α-Fe2O3.53 a) {001}, b) {012}, and c) {110}. Large black spheres are oxygen and small gray spheres are iron. The coordinatively unsaturated iron atoms on the {012} and {110} surfaces are shown by arrows. In conclusion, visible-light-induced RhB degradation with H2O2 at near neutral pH over hematite was strongly architecture-dependent and the reactivity trend can be rationalized as exposed facets in the order {110}>{012}≫{001}. This study suggests a promising new strategy for engineering practical photocatalysts for wastewater treatment based on heterogeneous photo-Fenton oxidation. Furthermore, this study has broad implications in hematite-based water-splitting and solar cells, among others. All chemicals (ACS grade) were purchased from Alfa Aesar and used without further purification. Milli-Q water (18 MΩ cm in resistivity, Millipore Corporation) was used throughout the experiment. The synthesis for hematite nanorods has been reported.39, 40, 47 For the synthesis of rodlike goethite (α-FeOOH) precursors, dialyzed suspensions of ferrihydrite nanoparticles were quickly adjusted to pH 12 using a 5 M sodium hydroxide, and the resulting suspensions were aged at a specific temperature for 30–90 h. The aged suspensions were then dialyzed using a membrane (MWCO=2000, SpectraPor) in Milli-Q water for three days. The dimensions of α-FeOOH nanorods can be tuned by changing the pH, aging temperature, and aging time. α-Fe2O3 nanorods were obtained by heating corresponding α-FeOOH precursors at 300 °C in air for 1 h. Nanocubes41 and nanoplates42 of hematite were synthesized following procedures reported previously and described in the Supporting Information. Powder XRD data were collected using a Rigaku Corporation X-ray diffractometer (XRD-6000) with CuKα radiation (λ=0.154178 nm, 50 kV, 300 mA). Particle morphology was characterized on a Hitachi S4800 field-emission scanning electron microscope (SEM). TEM and HRTEM images were obtained using a Tecnai G2 F20 U-TWIN microscope operating at 200 kV. Tapping-mode AFM measurements were carried out under ambient conditions using a Nanoscope IIIa (Veeco). Micro-Raman spectroscopy measurement was conducted using a Renishaw Micro-Raman Spectroscopy System (Renishaw in Via plus). A Renishaw red laser at 785 nm was employed as the excitation source. Diffuse-reflectance ultraviolet and visible (DRUV/Vis) spectra were obtained using a Perkin–Elmer Lambda 950 UV/Vis spectrometer. Fine BaSO4 powder was used as a standard. X-ray photoelectron spectroscopy (XPS) data were obtained using an ESCALab 250 electron spectrometer from Thermo Scientific Corporation. Monochromatic 150 W AlKα radiation was utilized and low-energy electrons were used for charge compensation to neutralize the samples. The binding energies were referenced to the adventitious C 1s line at 284.8 eV. A Shirley-type background was subtracted from each spectrum and Avantage 4.15 software was used for curve-fitting. The atomic model of hematite facets was depicted using Materials Studio 5.5. Materials Studio is registered software of Accelrys Software Inc. The photocatalytic activity of the as-prepared samples for the degradation of RhB in aqueous solutions was evaluated by measuring the absorbance of the irradiated solution. Prior to irradiation, 10 mg photocatalyst was mixed with RhB (50 mL, with a concentration of 0.02 mM) in a 100 mL round bottom flask. Afterwards, the suspension was magnetically stirred in the dark to reach a complete adsorption–desorption equilibrium, followed by the addition of 0.255 mL of hydrogen peroxide solution (H2O2, 30 wt %). The suspension (initial pH 6.8) was then illuminated by a 300 W xenon lamp. A cutoff filter of 420 nm was utilized to allow visible light to transmit, and a water filter was placed between the sample and the light source to eliminate infrared irradiation. The light intensity in the position of the center of the flask was measured to be about 150 mW cm−2 using a Newport optical power/energy meter (842-PE). During the irradiation, the reaction suspension was magnetically stirred and kept at RT. At certain time intervals, about 4 mL aliquots were sampled, centrifuged and filtered through a membrane (0.22 μm in diameter, Agela Technologies). The dye concentration in the filtrate was measured the absorption intensity of RhB at 554 nm using a PerkinElmer Lambda 950 UV/Vis spectrometer. The reaction products were evaluated using a series of analytical instruments described in the Supporting Information. Detailed facts of importance to specialist readers are published as ”Supporting Information”. Such documents are peer-reviewed, but not copy-edited or typeset. They are made available as submitted by the authors. Please note: The publisher is not responsible for the content or functionality of any supporting information supplied by the authors. Any queries (other than missing content) should be directed to the corresponding author for the article.