Immobilization of cadmium ions to synthesis hierarchical flowerlike cadmium phosphates microspheres and their application in the degradation of organic pollutants under light irradiation

Tingjiang Yan*ab, Wenfei Guana, Liting Cuia, Yanqiu Xua and Jun Tiana
aThe Key Laboratory of Life-Organic Analysis, College of Chemistry and Chemical Engineering, Qufu Normal University, Qufu, Shandong 273165, P. R. China. E-mail: tingjiangn@163.com
bState Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, P. R. China

Received 21st April 2015 , Accepted 8th May 2015

First published on 8th May 2015


Abstract

The treatment of wastewater especially that contaminated by heavy metals and/or organic pollutants by a green, cost-efficient and robust route is highly desirable. Herein, by addition of phosphates to immobilize cadmium ions, hierarchical flowerlike Cd5H2(PO4)4·4H2O microspheres were successfully prepared by nanosheet formation and following self-assembly at room temperature without additive assistance. The influence of experimental parameters such as pH value and raw materials (cadmium ions and phosphate ions) on the morphology and crystal structure of the products was studied. Hierarchical Cd5(PO4)2P2O7 was readily obtained by calcination of the Cd5H2(PO4)4·4H2O precursor. The microstructures of the products were characterized by X-ray power diffraction (XRD), field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM), Fourier transform infrared (FTIR), N2 adsorption–desorption (BET), and UV-vis diffuse reflectance spectroscopy (DRS). The as-prepared Cd5H2(PO4)4·4H2O and Cd5(PO4)2P2O7 possessed a wide band gap energy of 5.56 eV and ca. 5.00 eV, respectively. Owing to the strong oxidation/reduction ability and the unique hierarchical structure, hierarchical Cd5(PO4)2P2O7 microspheres exhibited excellent photocatalytic activity and durability for the degradation of organic pollutants (rhodamine B, RhB) under UV light irradiation. The main reactive species and the photocatalytic mechanism of Cd5(PO4)2P2O7 towards RhB degradation were also discussed.


Introduction

Photocatalytic oxidation technology has been proven to be potentially advantageous for environmental remediation such as wastewater treatment and air purification with solar energy.1 Over the past several years, considerable efforts have been devoted to acquire high photocatalytic performance materials through exploring novel photocatalytic materials or modifying reported photocatalysts. More than 150 photocatalysts including oxides, hydroxides, sulfides, nitrides, and halides have been reported to be able to degrade pollutants under light irradiation.2 Nevertheless, some photocatalysts show low activity and poorly structural stability, while others possess weak photooxidation ability toward the elimination of contaminants.3 Thus, developing new efficient and stable photocatalysts is still a great challenge in the photocatalysis field. Additionally, it has been shown that photocatalytic materials with three-dimensional (3D) hierarchical micro/nanostructures usually exhibit high efficiency and have specificity (i.e., good recyclability and high stability) for degradation of pollutants because of their unique dimensional and structural characteristics different from mono-morphological structures.4–6 Up to now, remarkable progress in the controllable fabrication of hierarchical structures with different shapes has been achieved. Many hierarchical superstructures with different composition, including elements, metal oxides, and metal sulfides were successfully synthesized via various methods.7–10 Although template-directed synthesis has been demonstrated as a versatile approach to fabricate hierarchical structures, the wealth of templates including hard templates and soft templates complicated the synthetic procedures and usually brought in harmful substrates to the environment and human health.11 Thus, the development for template-free and environmentally benign methods to synthesize hierarchical structures is still an ongoing progress. More recently, self-assembly of low-dimensional building blocks into complex 3D uniform micro/nanostructures on the basis of some intrinsically physical phenomena, such as the Kirkendall effect, Ostwald ripening, and oriented attachment process, to fabricate hierarchical structures provides new alternative routes for template-free synthesis of hierarchical materials.12–15

Metal phosphates, as one of important inorganic materials, have attracted the widespread attention recent decades because of their potential applications in the fields of ion-exchange, proton conductivity, biocompatibility, adsorption and catalysis.16–18 Various hierarchical metal phosphates have been controllably prepared and shown interesting properties. For instance, hydroxyapatite (Ca(PO4)6(OH)2), a typically bioactive and biocompatible material, was fabricated with various hierarchical microstructures via simple hydrothermal method in the absence of any surfactants.19–21 The as-synthesized hydroxyapatite exhibited highly structural stability (up to 800 °C), high efficiency in protein/heavy metal adsorption, and enzymatic catalysis.21–24 Wang et al. synthesized flowerlike titanium phosphate (Ti(HPO4)2·H2O) crystals and investigated their properties in removal of lead ions.21 Hierarchical iron hydrogen phosphate (Fe3H9(PO4)6·6H2O) and strontium flourapatite (Sr5(PO4)3OH1−xFx) were also prepared and shown special application in wastewater treatment and H2O2 sensing.25,26 In particular, metal phosphates have been found to be good photocatalysts with strong photocatalytic ability to decompose organic contaminants and split water due to the inductive effect of PO43−, which favors the separation efficiency of electrons–holes pairs in photocatalysis.27 Numerous metal phosphate materials such as hydroxyapatite (Ca(PO4)6(OH)2, TiO2/Ca(PO4)6(OH)2, Ag3PO4/Ca(PO4)6(OH)2), BiCu2PO6, BiPO4, Ag3PO4, Cu2(OH)PO4, and Ti2O(PO4)2(H2O)2 have been explored as novel photocatalysts.27–34 Among them, cadmium phosphate was found to possess excellent adsorption capacities for Pb(II) ions and novel catalytic property for the dehydrogenation of alcohols.35,36 Its nucleation and crystal growth kinetics based on mass crystallization experiments was also investigated.37 However, to the best of our knowledge, there have been seldom reports on the controllable preparation of hierarchical cadmium phosphate and their photocatalytic application.

Herein, hierarchical flowerlike Cd5H2(PO4)4·4H2O microspheres assembled with nanosheets were successfully prepared by addition of phosphates to cadmium ions solution at room temperature without additives assistant. Various cadmium ions including CdCl2, Cd(NO3)2, and Cd(CH3COO)2 could be immobilized by such route to form hierarchical structures. Cd5(PO4)2P2O7 with similar hierarchical structures was readily obtained by calcination of Cd5H2(PO4)4·4H2O and was firstly applied as novel photocatalyst for dye wastewater treatment. Such hierarchical Cd5(PO4)2P2O7 showed excellent photocatalytic efficiency and structural stability during the photocatalytic process. The main active radicals during dye degradation and the possible photocatalytic mechanism were also investigated.

Experimental section

Preparation

Hierarchical cadmium phosphate precursor, Cd5H2(PO4)4·4H2O, was fabricated via a simple precipitation method without the assistance of any template or surfactant. All used analytical grade materials were purchased from the Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China) and were used without further purification. In a typical process, 1.8 g Na2HPO4·12H2O and 1.6 g Cd(CH3COO)2·2H2O were dissolved in 100 mL of deionized water, respectively. Then the Na2HPO4 aqueous solution was added dropwise to the Cd(CH3COO)2 aqueous solution under stirring within 30 min. The mixture was stirred for 1 h and the resultant white precipitates were collected by filtration, washed with distilled water repeatedly, and dried at 60 °C overnight. The as-obtained Cd5H2(PO4)4·4H2O sample was obtained and labeled as T0. The Cd5H2(PO4)4·4H2O sample was heated in a tube furnace at 400, 500 and 700 °C for 4 h in air to yield Cd5(PO4)2P2O7, which was respectively labeled as T400, T500, and T700.

Sample characterizations

The crystal phase of the as-synthesized samples was examined by X-ray diffraction (XRD) on a Rigaku MinFlex II equipped with Cu Kα irradiation (λ = 0.15418 nm). Morphologies of the samples were observed by field emission scanning electron microscope (FESEM, JSM-6700F). Selected area electron diffraction (SAED) and transmission electron micrograph (TEM) were recorded on a JEM-2010. Fourier transform infrared spectra (FTIR) of the samples were recorded on a Perkin-Elmer IR spectrophotometer using a KBr pellet technique. The thermal behavior of the samples was investigated by thermogravimetric analysis using a Netzsch thermoanalyzer STA449C at a heating rate of 15 °C min−1 in air atmosphere. The UV-vis reflection spectra (DRS) were recorded on a Varian Cary 500 UV-vis spectrophotometer using BaSO4 as a reference and converted from reflection to absorbance by the Kubelka–Munk method. The specific surface areas of the samples were measured by nitrogen adsorption–desorption isotherms at 77 K using an automatic ASAP-2020 micromeritics analyzer. A desorption isotherm was used to determine the pore size distribution using the Barrett–Joyner–Halenda (BJH) method. Before each measurement, each sample was degassed for 6 h in helium at 333 K. The Cd2+ content of the residual solution after photocatalytic reaction was carried out by inductively coupled plasma (ICP) spectrometry (Ultima2).

Evaluation of photocatalytic activity

For UV-light photocatalytic activity, four UV lamps (Philips, TUV 4W/G4 T5) with wavelength centers at 254 nm were used. 100 mg powder was suspended in 100 mL rhodamine B (RhB) aqueous solution (10 ppm). Prior to irradiation, the suspensions were magnetically stirred in the dark for 30 min to attain the adsorption/desorption equilibrium between RhB and catalysts at ambient conditions. At varied irradiation time intervals, 3 mL suspension was collected and analyzed on a Perkin-Elmer UV WinLab Lambda 35 spectrophotometer. The degradation percentage is reported as C/C0, where C0 is the initial concentration of RhB, and C represents the corresponding concentration at a certain time interval. Photocatalytic stability was performed as follows: after each degradation experiment, the suspension was filtered and the solids were washed with distilled water and dried at 60 °C in air. Then the regenerated product was employed to degrade a new dye aqueous solution for another test under the same UV light irradiation. The generation of hydroxyl radicals (˙OH) was investigated using the PL (F-4600 type) technique, in which a basic terephthalic acid (TA) solution including 5 × 10−3 M TA and 0.01 M NaOH was added to the reactor. The excitation wavelength was 312 nm. The detection of ˙O2 radicals was performed by a nitroblue tetrazolium (NBT) probe method according to the literature.38

Results and discussion

The synthesis of Cd5H2(PO4)4·4H2O hierarchical architectures was achieved by simply mixing cadmium ions solution and phosphate ions solution at room temperature without additives assistant. The morphology and composition of the as-prepared samples were firstly examined by SEM and TEM. The low-magnification SEM image in Fig. 1a shows that the Cd5H2(PO4)4·4H2O products are mainly composed of numerous of flowerlike 3D microspheres with diameters ranging from 3 to 6 μm. Closer observations by the higher magnification SEM image in Fig. 1b and TEM in Fig. 1c and d expose that each flowerlike microsphere is composed of dozens of nanosheets with thickness of about 30 nm. These nanosheets as building blocks are connected to each other to build the flowerlike architectures. The selected-area electron diffraction (SAED) pattern (Fig. 1e) taken on a single nanosheet suggests the single-crystalline nature of the nanosheet unit.
image file: c5ra07224g-f1.tif
Fig. 1 (a and b) FESEM images (c and d) TEM images with different magnifications, and (e) SAED pattern of the as-synthesized Cd5H2(PO4)4·4H2O precursor.

XRD, shown in Fig. 2a, was used to study the crystal structure and phase composition of the obtained precursor. According to the main diffraction peak locations, the sample could be easily indexed to Cd5H2(PO4)4·4H2O, which matched well with the standard pattern (JCPDS card no. 14-0400). The low-angle reflection at 10.6° with strong diffraction peak confirms the presence of interlayer spacing in the lamellar structure.10 In addition, the FTIR of the Cd5H2(PO4)4·4H2O precursor is shown in Fig. 2b. The broad band from 3000 to 3500 cm−1 can be assigned to the O–H stretching vibrations while the sharp peak located at 1625 cm−1 corresponds to the bending mode vibration of water molecules.39 The observation of two shoulders at about 2440 and 1237 cm−1 can be associated to the (P)–O–H stretching modes and the deformation of P–O–(H) in the plane, respectively.35,39 Considering the studies of other phosphate compounds, the vibrations of the P–O bonds are also observed from about 900 to 1200 cm−1 (νas(P–O)) and from 520 to 600 cm−1 (δas(P–O)).39 As for the peaks at 587, 671, and 735 cm−1, they can be related to the stretching vibration of Cd–O bonds, which show obvious shifts as compared to the Cd–O bonds in pure CdO.40 The shift of these characteristic bands confirms the incorporation of phosphorus into the framework in the form of Cd–O–P bonds.41


image file: c5ra07224g-f2.tif
Fig. 2 (a) XRD pattern and (b) FTIR spectrum of the as-synthesized Cd5H2(PO4)4·4H2O precursor.

To understand the formation mechanism of the hierarchical Cd5H2(PO4)4·4H2O microspheres, time-dependent experiments were performed, in which intermediate products were collected at different intervals once Na2HPO4 was added into the Cd(CH3COO)2 solution media. All of the intermediate products were monitored by SEM and the results are shown in Fig. 3. At the initial stage (1 min) with little Na2HPO4 adding, the obtained precursors were ultrafine nanoparticles with size lower than 100 nm (Fig. 3a). As the reaction proceeded (5 min) and more Na2HPO4 was added, the primary nanoparticles tended to aggregate together to form microscaled spheres (Fig. 3b). It was found that some hierarchical structures assembled with connected nanosheets were observed when Na2HPO4 was completely added after 30 min of reaction (Fig. 3c). However, the hierarchical structures in this case were partial or incomplete. Uniformly flowerlike structures were completely formed after 3 h of reaction (Fig. 3d). For clarity, the morphological evolution process of the flowerlike Cd5H2(PO4)4·4H2O microspheres is illustrated in Fig. 3e, which involves a nucleation–aggregation–ripening process.15 Ultrafine Cd5H2(PO4)4·4H2O nanoparticles were produced by the reaction between HPO42− and Cd2+ at room temperature. Meanwhile, the freshly formed monomers were unstable due to their high surface energy and aggregated together gradually, resulting in the microscaled spheres. These monomers located in the Cd5H2(PO4)4·4H2O microspheres further underwent a dissolution–recrystallization (Ostwald ripening) process to form the nanosheets, which were further assembled into hierarchical Cd5H2(PO4)4·4H2O microspheres.


image file: c5ra07224g-f3.tif
Fig. 3 SEM images of the nanostructures collected at different reaction time once the Na2HPO4 was added: (a) 1 min, (b) 5 min, (c) 30 min, and (d) 3 h. (e) A schematic illustration of the formation process for hierarchical Cd5H2(PO4)4·4H2O microspheres.

Control experiments were conducted to reveal suitable reaction condition for the preparation of flowerlike Cd5H2(PO4)4·4H2O microspheres. As shown from Fig. 4a and b, flowerlike Cd5H2(PO4)4·4H2O microspheres could be prepared when other cadmium ions such as Cd(NO3)2 or CdCl2 were used as starting material instead of Cd(CH3COO)2. The XRD patterns of the collected samples show that they have the composition of Cd5H2(PO4)4·4H2O. However, when NaH2PO4 was replaced by Na2HPO4 or Na3PO4 (Fig. 4c and d), novel hierarchical Cd5H2(PO4)4·4H2O microspheres and irregular nanoparticles with amorphous phase were observed, respectively. The change of morphology and crystal phase of the final products may be explained by the release of H+ ions to solution from various phosphate anion species H2PO4, HPO42− and PO43−, while PO43− ions were consumed for cadmium phosphate formation.42,43 Thus, we also investigated the influence of pH value on the final products. It was found that the optical pH values for the formation of flowerlike Cd5H2(PO4)4·4H2O microspheres were between 5 and 7. A slightly decrease in pH value to about 4 by addition of a small amount of HNO3 solution contributed to Cd5H2(PO4)4·4H2O microspheres assembled by large nanoplates with width of ca. 300 nm (Fig. 4e). These large nanoplates were arranged compactly to form the 3D microspheres and no interleaved pores were constructed in this condition. On the other hand, when the pH value was increased above 9 by addition of NaOH solution, the obtained products were irregular nanoparticles with amorphous characteristics (Fig. 4f), supported by the weak diffraction peaks as shown in the XRD pattern. From the above results, it can be concluded that by simply tuning the phosphates or the pH value, hierarchical Cd5H2(PO4)4·4H2O microspheres could be easily fabricated via in situ immobilization of cadmium ions in solution, which provides a simple and environmental attractive remediation method for wastewater contaminated by cadmium ions.


image file: c5ra07224g-f4.tif
Fig. 4 SEM images and XRD patterns of the obtained products collected from various conditions: (a) using CdCl2 as cadmium source and Na2HPO4 as phosphate source, (b) using Cd(NO3)2 as cadmium source and Na2HPO4 as phosphate source, (c) using Cd(CH3COO)2 as cadmium source and NaH2PO4 as phosphate source, (d) using Cd(CH3COO)2 as cadmium source and Na3PO4 as phosphate source, (e) using Cd(CH3COO)2 as cadmium source and Na2HPO4 as phosphate source at pH value of 4, and (f) using Cd(CH3COO)2 as cadmium source and Na2HPO4 as phosphate source at pH value of 9.

As a typical metal hydrogen phosphate, Cd5H2(PO4)4·4H2O may also be written as Cd5(PO4)2(HPO4)2·4H2O, in which HPO4 and H2O entities are highly temperature-dependent, suffering from condensation and dehydration reaction, respectively.44 Thus, the effect of thermal treatment on microstructures of Cd5H2(PO4)4·4H2O precursor was further investigated. The TG-DSC curve of Cd5H2(PO4)4·4H2O is initially shown in Fig. 5a. It displays two main steps of weight loss in the range of 30–900 °C. The first weight loss of 7.4% below 300 °C can be attributed to the loss of four molecule crystal water of Cd5H2(PO4)4·4H2O leading to Cd5H2(PO4)4. An addition weight loss of 1.8% is observed in the range of 300–600 °C, corresponding to the loss of one molecule of water from the condensation reaction of HPO4 groups, which seems to take place leading to the formation of Cd5(PO4)2P2O7, analogous to U2(PO4)2HPO4·H2O.44 No significant mass loss was detected above 600 °C. The XRD patterns (Fig. 5b) indicate that the resulted samples after calcination at 400, 500, and 700 °C are pure Cd5(PO4)2P2O7 (JCPDS card no. 14-0399). With increasing the calcination temperature, XRD patterns become sharper, which indicates the increased crystallinity and grain growth as well. The morphology of the obtained Cd5(PO4)2P2O7 samples was examined by SEM and TEM. As seen from Fig. 6a–d, the hierarchical structure was still retained upon calcination at 400 °C. However, due to the condensation and dehydration process, the single-crystalline nanosheets converted into many smaller nanoparticles along with numerous pores with size of 10–150 nm were generated within the nanosheets. The corresponding SAED pattern (Fig. 6e) depicts that the nanosheet assembled from nanoparticles is polycrystalline in nature. Further increasing calcination temperature to 500 °C caused the grain growth and particle aggregation (Fig. 6f). When the temperature increased to 700 °C, the nanosheets were completely converted into large nanoparticles with size of ca. 200 nm. The hierarchical flowers finally changed to irregular spheres and no pores were observed due to the aggregation of particles (Fig. 6g).


image file: c5ra07224g-f5.tif
Fig. 5 (a) TG-DSC curves of the Cd5H2(PO4)4·4H2O precursor, and (b) XRD patterns of the obtained Cd5(PO4)2P2O7 samples.

image file: c5ra07224g-f6.tif
Fig. 6 SEM (a and b), TEM (c and d) and SAED pattern (e) of the Cd5(PO4)2P2O7 samples (T400); SEM image (f) of the Cd5(PO4)2P2O7 sample (T500); and SEM image (g) of the Cd5(PO4)2P2O7 sample (T700).

The optical properties of the Cd5H2(PO4)4·4H2O precursor and the calcined Cd5(PO4)2P2O7 were measured using UV-vis diffuse reflectance spectra. Fig. 7a shows the UV absorption spectra of the samples, indicating that the maximal absorbance wavelengths of Cd5H2(PO4)4·4H2O and Cd5(PO4)2P2O7 (T400, T500, and T700) are approximately 233 nm, 255 nm, 266 nm and 268 nm, respectively. The band gap energy (Eg) was evaluated using the equation, α() = A(Eg)n/2, where α, h, Eg, and A are the absorption coefficient, light frequency, band gap energy, and a constant, respectively, and n is determined by the type of optical transition in the semiconductor. The band gap energy (Eg values) of these samples was estimated from a plot of α()2 as a function of the photon energy () (Fig. 7b) to be approximately 5.56 eV, 5.04 eV, 4.97 eV, and 4.95 eV for Cd5H2(PO4)4·4H2O and Cd5(PO4)2P2O7 samples (T400, T500, and T700), respectively.


image file: c5ra07224g-f7.tif
Fig. 7 (a) UV-vis diffuse reflection spectra of the as-synthesized Cd5H2(PO4)4·4H2O precursor (T0) and the calcined Cd5(PO4)2P2O7 samples (T400, T500, and T700), and (b) the plots of (αhν)2 as a function of photon energy ().

The specific surface area and porosity of the hierarchical Cd5H2(PO4)4·4H2O and Cd5(PO4)2P2O7 were measured by N2 adsorption–desorption isotherms. As shown in Fig. 8a, the isotherm of the Cd5H2(PO4)4·4H2O precursor exhibits a hysteresis loop in a relative pressure range of 0.6 to 1.0, implying the presence of mesopores in Cd5H2(PO4)4·4H2O hierarchical structures. The corresponding pore-size distribution (insert in Fig. 8a) shows that most of the pores fall into the size region from 5 to 80 nm. The mesopores are formed most likely due to the intercrossing of nanosheets.45 The BET surface area of the Cd5H2(PO4)4·4H2O precursor was determined to be 7.7 m2 g−1. Note that after 400 °C calcination, similar porosity and more broader pore-size distribution (from 5 to 130 nm) were observed on Cd5(PO4)2P2O7 (T400). The BET surface area of the Cd5(PO4)2P2O7 sample (T400) was calculated to be 7.4 m2 g−1. The broad pore-size distribution for Cd5(PO4)2P2O7 (T400) could be attributed to the generated pores within the original single-crystalline nanosheets (Fig. 6a–d). However, since higher temperature caused the growth and aggregation of particles (Fig. 6f and g), no obvious pores were found in the corresponding Cd5(PO4)2P2O7 samples (T500 and T700) (Fig. S1). The BET surface area of the Cd5(PO4)2P2O7 samples (T500 and T700) was also reduced to 4.5 and 1.6 m2 g−1, respectively.


image file: c5ra07224g-f8.tif
Fig. 8 Nitrogen adsorption–desorption isotherms of (a) Cd5H2(PO4)4·4H2O (T0) and (b) Cd5(PO4)2P2O7 (T400). The inset is the corresponding pore-size distribution.

Degradation of RhB was examined as a probe reaction to evaluate the photocatalytic activities of the as-prepared samples under UV light irradiation. The degradation curves of RhB in the presence/absence of photocatalysts as a function of irradiation time are plotted in Fig. 9a. The blank experiment without photocatalyst shows that about 36% of RhB was decomposed within 100 min by the photolysis due to the strong energy of UV light. The addition of Cd5H2(PO4)4·4H2O precursor into RhB system gave rise to a negligible enhancement in RhB degradation, suggesting that Cd5H2(PO4)4·4H2O with band gap of 5.56 eV can not be excited by the used UV light to generate photoinduced electron–hole pairs. However, Cd5H2(PO4)4·4H2O can act as supporting materials by grafting active semiconductor (such as Ag3PO4) to show superior photocatalytic property under the irradiation of visible light (Fig. S2). In contrast, Cd5(PO4)2P2O7 sample after 400 calcination exhibited an excellent photocatalytic activity for RhB degradation under UV light irradiation. After 100 min irradiation, about 93% of RhB was decomposed and the mineralization yield of RhB reached 100% from TOC measurement. It has also been found that the photocatalytic degradation of RhB in the presence/absence of different photocatalysts followed the pseudo-first-kinetics model, ln(Ct/C0) = −kt, where C0 and Ct are the initial concentration of the dye solution and the concentration at time t, respectively, and k is the kinetic constant. As displayed in Fig. 9b, the Cd5(PO4)2P2O7 (T400) has much higher rate constant (ca. 0.02729 min−1) than Cd5H2(PO4)4·4H2O (ca. 0.006 min−1) and is approximately 6.4 times higher than that of the photolysis (ca. 0.00427 min−1). Considering the perniciousness of Cd2+ in Cd5H2(PO4)4·4H2O and Cd5(PO4)2P2O7, we also tested the Cd2+ content of the residual solution after photocatalytic reaction by ICP. It is found that no Cd2+ was dissolved out from these two materials, suggesting the high stability of these hierarchical cadmium phosphates.


image file: c5ra07224g-f9.tif
Fig. 9 (a) Photocatalytic degradation curves of RhB as a function of the irradiation time in the presence/absence of different photocatalysts, (b) the kinetics over different photocatalysts.

For comparison, the photocatalytic activity of the hierarchical flowerlike Cd5(PO4)2P2O7 was also compared with other reported phosphate photocatalysts such as BiPO4, Ca(PO4)6(OH)2 and Ti2O(PO4)2(H2O)2. As shown in Fig. 10, the activity of Cd5(PO4)2P2O7 was much higher than that of Ca(PO4)6(OH)2 and Ti2O(PO4)2(H2O)2 and comparable to that of BiPO4. The inductive effect of PO43− playing in photocatalysis has been well proposed.27 Cd5(PO4)2P2O7 with mixed PO43− and P2O74− groups may have larger inductive effect than Ca(PO4)6(OH)2, Ti2O(PO4)2(H2O)2, and BiPO4 possessed PO43− group only, and therefore a higher photocatalytic activity. Besides, the hierarchical structure of Cd5(PO4)2P2O7 also benefits the activity. We also compared the activity of flowerlike Cd5(PO4)2P2O7 with irregular nanoparticles of Cd5(PO4)2P2O7 (SEM was given in Fig. S3). As presented in Fig. 10, Cd5(PO4)2P2O7 with hierarchical structure exhibited superior activity to that of the nanoparticles, which might be due to the hierarchical structure benefiting the light utilization efficiency and dye adsorption property. The detailed relationship between microstructures and photocatalytic activity should be further investigated.


image file: c5ra07224g-f10.tif
Fig. 10 Comparative photocatalytic activity of hierarchical flowerlike Cd5(PO4)2P2O7 (T400) with other phosphate photocatalysts and Cd5(PO4)2P2O7 (T400) with other morphology (irregular nanoparticles) under UV (254 nm) light irradiation.

The effect of calcination temperature on the photocatalytic performance of the resulted Cd5(PO4)2P2O7 samples was also investigated in the present work. As shown in Fig. 11a, Cd5(PO4)2P2O7 sample after 400 °C calcination (T400) possessed the highest photocatalytic performance (93%) for RhB degradation. Increasing the calcination temperature from 400 °C to 500 °C resulted in a 16% decrease in removal rate of RhB. The slight decrease may be due to its lower surface area and growth of particle size, though the degree of crystallinity for T500 is much higher than that for T400. Further calcination at 700 °C, a very low removal rate (only 43%) was obtained. Such a great decrease of activity should be attributed to the agglomeration of large particle, the reduced surface area and the destruction of unique hierarchical structure at higher temperature.


image file: c5ra07224g-f11.tif
Fig. 11 (a) Photocatalytic degradation curves of RhB as a function of the irradiation time over Cd5(PO4)2P2O7 samples calcined at different temperatures, (b) the recyclability of the photocatalytic degradation of RhB in the presence of the hierarchical Cd5(PO4)2P2O7 microspheres (T400), (c) XRD patterns of the hierarchical Cd5(PO4)2P2O7 microspheres (T400) before and after five runs of photocatalytic reaction, and (d) SEM image of the hierarchical Cd5(PO4)2P2O7 microspheres (T400) after five runs of photocatalytic reaction.

Except for photocatalytic efficiency, stability is another critical factor for the wide application of photocatalysts, especially for those noxious element-containing photocatalysts like CdSe, Pb3O4 Ag3AsO4, and etc.46–48 Thus, the stability of the hierarchical Cd5(PO4)2P2O7 was further investigated by recycling the degradation of RhB under the same conditions. As shown in Fig. 11b, the hierarchical Cd5(PO4)2P2O7 sample (T400) did not show any apparent decrease in RhB removal even after five successive operations. Both XRD and SEM examinations on Cd5(PO4)2P2O7 further confirm that there were no obvious changes in the crystal structure or in the flowerlike architecture of the catalysts before and after photoreaction (Fig. 11c and d). All these results demonstrate that the hierarchical Cd5(PO4)2P2O7 microspheres were highly stable and are promising candidates for wastewater treatment in solar-driven applications.

The photocatalytic mechanism of hierarchical Cd5(PO4)2P2O7 during RhB degradation was further investigated. It is generally accepted that the dyes can be degraded by a large number of reactive species including photoinduced holes (h+), hydroxyl radicals (˙OH), and superoxide anions (˙O2) involved in the photocatalytic oxidation process. Therefore, the reactive species scavengers, including ammonium oxalate (AO), tert-butyl alcohol (TBA), and benzoquinone (BQ), were employed to investigate the corresponding effect of h+, ˙OH, and ˙O2 on the degradation activity of RhB. From Fig. 12a, it was found that the addition of BQ greatly reduced the photocatalytic efficiency of RhB from 93% to about 44%. This indicates that the ˙O2 is the key factor affecting the photocatalytic performance of the Cd5(PO4)2P2O7 microspheres. The presence of ˙O2 radicals can be proved by a nitroblue tetrazolium (NBT) probe method.38 As shown from Fig. 12b, it was observed that the maximum absorption peak at 259 nm decreased with the prolonging irradiation time, indicating the specific reaction between NBT and ˙O2 radicals. When TBA and AO were added into the reaction system, the photocatalytic efficiency of RhB also decreased to 28 and 66%, respectively. This suggested that the photocatalytic oxidation of RhB may involve the direct oxidation by photoinduced h+ and the indirect oxidation by ˙OH radicals. The ˙OH radicals were also examined by a photoluminescence (PL) technology with terephthalic acid (TA) as a probe molecule. The fluorescence intensity was found to increase steadily with irradiation time (Fig. 12c), implying that ˙OH radicals can be generated in the Cd5(PO4)2P2O7 suspensions under UV light irradiation. In summary, the main reactive species involved in the photocatalytic degradation of RhB over Cd5(PO4)2P2O7 are ˙O2, h+, and ˙OH.


image file: c5ra07224g-f12.tif
Fig. 12 (a) Effects of scavengers on the degradation efficiency of RhB, (b) UV-vis absorption spectra of NBT in Cd5(PO4)2P2O7 (T400) suspension under UV irradiation, (c) ˙OH-trapping photoluminescence spectra of Cd5(PO4)2P2O7 (T400) under UV irradiation, and (d) the possible mechanism of the RhB degradation over Cd5(PO4)2P2O7.

Since the generation of reactive species involved in the photocatalytic reaction requires the band structure of semiconductor to meet the thermodynamic potential for the reaction, it is highly necessary to investigate the oxidizing or reducing ability of photogenerated carries. As for Cd5(PO4)2P2O7, the band position was calculated using the empirical equation, ECB = XEe + 0.5Eg, where ECB is the conduction band edge potentials, X is the electronegativity of the semiconductor, which is the geometric mean of the electronegativity of the constituent atoms, Ee is the energy of the free electrons on the hydrogen scale (approximately 4.5 eV), Eg is the band gap energy of the semiconductor, and EVB can be determined by EVB = ECB + Eg.27 The X value for Cd5(PO4)2P2O7 is 6.44 eV, and the Eg of Cd5(PO4)2P2O7 was set as 5.0 eV from the DRS results. Thus, the ECB and EVB of Cd5(PO4)2P2O7 are estimated to be −0.56 eV (vs. NHE) and 4.44 eV (vs. NHE), respectively. According to the band gap structures of Cd5(PO4)2P2O7 and the effects of scavengers on the photocatalytic reaction, a possible mechanism for the hierarchical Cd5(PO4)2P2O7 microspheres is proposed in Fig. 12d. Cd5(PO4)2P2O7 can be efficiently excited to generate photoinduced electron–hole pairs under UV irradiation. The photogenerated electrons in the CB of Cd5(PO4)2P2O7 can reduce O2 to yield ˙O2 as its potential is more negative than E(O2/O2˙) (−0.33 eV vs. NHE). On the other hand, because the photogenerated holes in the VB is more positive than E(˙OH/OH) (2.38 eV vs. NHE),49 it becomes more favorable for the holes to react with OH or H2O, producing active ˙OH radicals. The formed ˙O2 radicals, ˙OH radicals as well as the photoinduced holes with high oxidation power will decompose the RhB molecules and contribute to the high photocatalytic efficiency for Cd5(PO4)2P2O7.

Conclusions

Hierarchical flowerlike Cd5H2(PO4)4·4H2O microspheres have been simply prepared by addition of phosphates to immobilize various cadmium ions including CdCl2, Cd(NO3)2, and Cd(CH3COO)2 in aqueous solution. Such hierarchical Cd5H2(PO4)4·4H2O microspheres assembled with nanosheets were fabricated via an Ostwald ripening process in the absence of any template or surfactant. Hierarchical flowerlike Cd5(PO4)2P2O7 microspheres, which were achieved by condensation and dehydration of Cd5H2(PO4)4·4H2O, exhibited excellent photocatalytic performance for the degradation of RhB under UV light irradiation due to their hierarchical structure, numerous intra- and inter-pores, and strong oxidation ability. During successive operations for treatment of dye wastewater, the hierarchical Cd5(PO4)2P2O7 microspheres showed not only durable photocatalytic degradation toward RhB, but also highly structural stability the catalyst itself. Except for dye wastewater treatment, the hierarchical Cd5(PO4)2P2O7 microspheres with strong oxidation capability may also have potential applications in environmental purification, especially in efficient removal of persistent volatile organic pollutants such as aromatic hydrocarbons and polychlorinated biphenyls.

Acknowledgements

This work is supported by the National Natural Science Foundation of China (no. 21103193), Doctoral Foundation of Shandong Province (BS 2013NJ013) and Scientic Research Foundation of Qufu Normal University (BSQD20110116).

Notes and references

  1. X. B. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891–2959 CrossRef CAS PubMed.
  2. J. X. Wang, X. J. Wei, J. Y. Shen, X. M. Lu, J. M. Xie and M. Chen, Prog. Chem., 2014, 26, 1460–1470 Search PubMed.
  3. C. Wen, A. Y. Yin and W. L. Dai, Appl. Catal., B, 2014, 160, 730–741 CrossRef PubMed.
  4. H. X. Li, Z. F. Bian, J. Zhu, D. Q. Zhang, G. S. Li, Y. N. Huo, H. Li and Y. F. Lu, J. Am. Chem. Soc., 2007, 129, 8406–8407 CrossRef CAS PubMed.
  5. J. H. Huang, R. Ma, Y. Ebina, K. Fukuda, K. Takada and T. Sasaki, Chem. Mater., 2010, 22, 2582–2587 CrossRef CAS.
  6. F. K. Chen, J. T. Zai, M. Xu and X. F. Qian, J. Mater. Chem. A, 2013, 1, 4316–4323 CAS.
  7. W. A. Lopes and H. M. Jaeger, Nature, 2001, 414, 735–738 CrossRef CAS PubMed.
  8. G. J. D. Soler-Illia, C. Sanchez, B. Lebeau and J. Patarin, Chem. Rev., 2002, 102, 4093–4138 CrossRef PubMed.
  9. R. M. Xing, Y. Xue, X. Q. Liu, B. S. Liu, B. J. Miao, W. Z. Kang and S. H. Liu, CrystEngComm, 2012, 14, 8044–8048 RSC.
  10. H. X. Zhong, Y. L. Ma, X. F. Cao, X. T. Chen and Z. L. Xue, J. Phys. Chem. C, 2009, 113, 3461–3466 CAS.
  11. M. Shang, W. Z. Wang, J. Ren, S. M. Sun and L. Zhang, CrystEngComm, 2010, 12, 1754–1758 RSC.
  12. X. F. Chen, J. B. Liu, H. Wang, Y. L. Ding, Y. X. Sun and H. Yan, J. Mater. Chem. A, 2013, 1, 877–883 CAS.
  13. Z. Wang, J. G. Hou, C. Yang, S. Q. Jiao, K. Huang and H. M. Zhu, Phys. Chem. Chem. Phys., 2013, 15, 3249–3255 RSC.
  14. J. B. Fei, Y. Cui, X. H. Yan, W. Qi, Y. Yang, K. W. Wang, Q. He and J. B. Li, Adv. Mater., 2008, 20, 452–456 CrossRef PubMed.
  15. B. Jiang, P. Zhang, Y. Zhang, L. Wu, H. X. Li, D. Q. Zhang and G. S. Li, Nanoscale, 2012, 4, 455–460 RSC.
  16. C. N. R. Rao, S. Natarajan, A. Choudhury, S. Neeraj and A. A. Ayi, Acc. Chem. Res., 2001, 34, 80–87 CrossRef CAS PubMed.
  17. R. Murugavel, A. Choudhury, M. G. Walawalkar, R. Pothiraja and C. N. R. Rao, Chem. Rev., 2008, 108, 3549–3655 CrossRef CAS PubMed.
  18. P. Olivera-Pastor, P. Maireles-Torres, E. Rodriguez-Castellon, A. Jimenez-Lopez, T. Cassagneau, D. J. Joner and J. Roziere, Chem. Mater., 1996, 8, 1758–1769 CrossRef CAS.
  19. H. Yang, H. J. Zeng, L. J. Hao, N. R. Zhao, C. Du, H. Liao and Y. J. Wang, J. Mater. Chem. B, 2014, 2, 4703–4710 RSC.
  20. X. L. Wang, J. F. Shi, Z. Li, S. H. Zhang, H. Wu, Z. Y. Jiang, C. Yang and C. Y. Tian, ACS Appl. Mater. Interfaces, 2014, 6, 14522–14532 CAS.
  21. X. J. Wang, X. L. Yang, J. H. Cai, T. T. Miao, L. H. Li, G. Li, D. R. Deng, L. Jiang and C. R. Wang, J. Mater. Chem. A, 2014, 2, 6718–6722 CAS.
  22. R. X. Sun, K. Z. Chen, Z. M. Liao and N. Meng, Mater. Res. Bull., 2013, 48, 1143–1147 CrossRef CAS PubMed.
  23. X. Y. Zhao, Y. J. Zhu, F. Chen, B. Q. Lu and J. Wu, CrystEngComm, 2013, 15, 206–212 RSC.
  24. S. D. Jiang, Q. Z. Yao, G. T. Zhou and S. Q. Fu, J. Phys. Chem. C, 2012, 116, 4484–4492 CAS.
  25. T. B. Zhang, Y. C. Lu and G. S. Luo, ACS Appl. Mater. Interfaces, 2014, 6, 14433–14438 CAS.
  26. N. Niu, D. Wang, S. H. Huang, C. X. Li, F. He, S. L. Gai, X. B. Li and P. P. Yang, CrystEngComm, 2012, 14, 1744–1752 RSC.
  27. C. S. Pan and Y. F. Zhu, Environ. Sci. Technol., 2010, 44, 5570–5574 CrossRef CAS PubMed.
  28. A. M. Hu, M. Li, C. G. Chang and D. L. Mao, J. Mol. Catal. A: Chem., 2007, 267, 79–85 CrossRef CAS PubMed.
  29. X. T. Hong, X. H. Wu, Q. Y. Zhang, M. F. Xiao, G. L. Yang, M. R. Qiu and G. C. Han, Appl. Surf. Sci., 2012, 258, 4801–4805 CrossRef CAS PubMed.
  30. Y. Yang, Y. Murakami, A. Y. Nosaka and Y. Nosaka, AZojomo, 2008, 4, 1–5 Search PubMed.
  31. Z. G. Yi, J. H. Ye, N. Kikugawa, T. Kako, S. X. Ouyang, H. Stuart-Williams, H. Yang, J. Y. Cao, W. J. Luo, Z. S. Li, Y. Liu and R. L. Withers, Nat. Mater., 2010, 9, 559–564 CrossRef CAS PubMed.
  32. I.-S. Cho, D. W. Kim, S. Lee, C. H. Kwak, S.-T. Bae, J. H. Noh, S. H. Yoon, H. S. Jung, D.-W. Kim and K. S. Hong, Adv. Funct. Mater., 2008, 18, 2154–2162 CrossRef CAS PubMed.
  33. G. Wang, B. B. Huang, X. B. Ma, Z. Y. Wang, X. Y. Qin, X. Y. Zhang, Y. Dai and M.-H. Whangbo, Angew. Chem., Int. Ed., 2013, 52, 4810–4813 CrossRef CAS PubMed.
  34. S. Y. Guo and S. Han, J. Power Sources, 2014, 267, 9–13 CrossRef CAS PubMed.
  35. P. Yin, R. J. Qu, X. G. Liu, X. Q. Dong and Q. Xu, Food Chem., 2014, 148, 307–313 CrossRef CAS PubMed.
  36. F. Nozaki and Y. Iimori, Bull. Chem. Soc. Jpn., 1976, 49, 567–568 CrossRef CAS.
  37. H. E. L. Madsen, J. Cryst. Growth, 2004, 263, 564–569 CrossRef PubMed.
  38. X. L. Xu, X. Duan, Z. G. Yi, Z. W. Zhou, X. M. Fan and Y. Wang, Catal. Commun., 2010, 12, 169–172 CrossRef CAS PubMed.
  39. V. Brandel, N. Clavier and N. Dacheux, J. Solid State Chem., 2005, 178, 1054–1063 CrossRef CAS PubMed.
  40. H. Gülce, V. Eskuizeybek, B. Haspulat, F. Sarl, A. Gülce and A. Avcl, Ind. Eng. Chem. Res., 2013, 52, 10924–10934 CrossRef.
  41. J. C. Yu, L. Z. Zhang, Z. Zheng and J. C. Zhao, Chem. Mater., 2003, 15, 2280–2286 CrossRef CAS.
  42. M. Manechi, P. A. Maurice and S. J. Traina, Soil Sci., 2000, 165, 920–933 CrossRef PubMed.
  43. K. J. Zhu, K. Yanagisawa, A. Onda and K. Kajiyoshi, J. Solid State Chem., 2004, 177, 4379–4385 CrossRef CAS PubMed.
  44. V. Brandel, N. Clavier and N. Dacheux, J. Solid State Chem., 2005, 178, 1054–1063 CrossRef CAS PubMed.
  45. S. P. Hu, C. Y. Xu, F. X. Ma, L. Cao and L. Zhen, Dalton Trans., 2014, 8439–8445 RSC.
  46. F. A. Frame, E. C. Carroll, D. S. Larsen, M. Sarahan, N. D. Browning and F. E. Osterloh, Chem. Commun., 2008, 2206–2208 RSC.
  47. Y. G. Zhou, J. L. Long, Q. Gu, H. X. Lin, H. Lin and X. X. Wang, Inorg. Chem., 2012, 51, 12594–12596 CrossRef CAS PubMed.
  48. J. T. Tang, Y. H. Liu, H. Z. Li, Z. Tan and D. T. Li, Chem. Commun., 2013, 49, 5498–5500 RSC.
  49. Z. H. Li, T. T. Dong, Y. F. Zhang, L. Wu, J. Q. Li, X. X. Wang and X. Z. Fu, J. Phys. Chem. C, 2007, 111, 4727–4733 CAS.

Footnote

Electronic supplementary information (ESI) available: N2 adsorption–desorption isotherms and pore-size distribution of Cd5(PO4)2P2O7 (T500 and T700), comparative photocatalytic activity of samples under the irradiation of visible light (λ > 400 nm), SEM image of Cd5(PO4)2P2O7 (T400) irregular nanoparticles. See DOI: 10.1039/c5ra07224g

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.