Chang-jiu Tenga,
Dan Xie*a,
Meng-xing Suna,
Jian-long Xub,
Chun-song Zhaoc,
Pu Yanga,
Yi-lin Suna,
Cheng Zhanga and
Xian Lia
aInstitute of Microelectronics, Tsinghua National Laboratory for Information Science and Technology (TNList), Tsinghua University, Beijing 100084, People's Republic of China. E-mail: xiedan@mail.tsinghua.edu.cn
bInstitute of Functional Nano and Soft Materials (FUNSOM), Jiangsu Key Laboratory for Carbon-based Functional Materials and Devices, Soochow University, Suzhou, Jiangsu 215123, P. R. China
cState Key Laboratory of New Ceramics and Fine Processing, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
First published on 20th June 2016
Bismuth ferrite (BiFeO3), which acts as a significant multiferroic material, exhibits unique magnetic and ferroelectric properties. Here, we adopted a sol–gel method to synthesize disordered nanoporous BFO with a sucrose template. By different characterization methods such as powder X-ray diffraction (XRD), transmission electron microscopy (TEM) and N2 physisorption measurements, the large BET (Brunauer–Emmett–Teller) surface area (34.25 m2 g−1) and disordered pores (20–40 nm) of the sucrose-templated nanoporous BFO (SN-BFO) materials were determined. These SN-BFO materials exhibited high magnetic performance (Ms = 6.37 emu g−1) and a favorable magnetic recycling ratio. In addition, SN-BFO displayed appropriate adsorption characteristics and good photocatalytic stability, which demonstrates that SN-BFO is a candidate for promising recoverable multifunctional environment-purifying applications in the future.
In this paper, non-surfactant sucrose was used as a non-solid template to synthesize BFO nanoporous structures with disordered pores by a one-pot sol–gel method. These SN-BFO structures exhibited greatly enhanced magnetic properties with a saturation magnetization value of about 6.0 emu g−1, which is attributed to an increase in the Fe2+ content in the BFO lattice arising from incomplete combustion of the sucrose template during calcination. Moreover, SN-BFO also displayed favorable recoverable photocatalytic performance and adsorption properties in the degradation of rhodamine B (RhB), which make it a promising candidate for future practical multifunctional environment-purifying applications.
Fig. 1(a) shows the XRD pattern recorded for the SN-BFO sample at RT. All the diffraction peaks can be indexed according to the standard rhombohedral distorted perovskite crystal structure of BFO with the R3c space group. The sharp and well-defined XRD peaks reveal that the sample is highly polycrystalline with few impurity phases (if the sample had not been immersed in acetic acid, diffraction peaks of Bi2Fe4O9 and Bi24Fe2O39 phases would have appeared in the XRD patterns). Fig. 1(b) shows the Raman shift of the SN-BFO sample, and all the recorded peaks can be identified according to the standard Raman modes of BFO with the rhombohedral R3c space group. As shown in Fig. 1(b), the Raman mode located at 274 cm−1, which is related to the stretching and bending of the FeO6 octahedra, is slightly weakened compared with that of the pure BFO NPs in our previous work.17 This indicates a reduction in the amount of FeO6 octahedra and an increase in distortion caused by the introduction of Fe2+, which is demonstrated by XPS, as discussed later. Fig. 1(c) shows a typical TEM image of SN-BFO nanopowder, which indicates a disordered nanoporous structure with a three-dimensional interconnecting configuration. Fig. 1(d) shows the lattice fringes of the (012) and (01) planes with interplanar distances of 0.40 nm and 0.29 nm. In addition, the components of the expected composition of SN-BFO, which were quantitatively analyzed from energy-dispersive X-ray spectroscopy (EDS) data, indicate a Bi
:
Fe atomic ratio of 1
:
1.
![]() | ||
Fig. 1 (a) X-ray diffraction pattern of SN-BFO powder; (b) Raman shift of SN-BFO powder; (c) typical TEM image of SN-BFO nanopowder; (d) HRTEM image of SN-BFO powder. |
The N2 adsorption–desorption isotherm of SN-BFO shown in Fig. 2(a) reveals its nanoporous structure, which is consistent with Fig. 1. The BET surface area of SN-BFO was calculated to be 34.25 m2 g−1, which is mainly due to nanoscale pores formed by the disappearance of the sucrose template during calcination. The pore size distribution of SN-BFO can also be determined from the desorption branch by the Barrett–Joyner–Halenda (BJH) method, as shown in Fig. 2(a) (inset). The obvious hysteresis loop that was obtained also demonstrates the disordered nanoporous structure,20 and the irregular porosity was also determined from the values of dV/dR for the BJH pore size distribution. The pore diameter of SN-BFO was thus calculated to be between 20 and 50 nm. At low relative pressures of less than 0.6, the isotherm exhibits no obvious loops, which is possibly due to small-sized pores ranging from 3 to 5 nm, and the hysteresis loop that was formed at high relative pressures of greater than 0.6 can be assigned to pores with larger sizes above the 10 nm range. In this work, the hysteresis loop in the range from 0.7 to 1 is caused by disordered pores with sizes from 20 to 50 nm.20,21 Fig. 2(b) shows the RT curve of magnetization versus magnetic field (M–H) for SN-BFO with a maximum applied magnetic field of 50 kOe. In our previous work,17 BFO NPs prepared in the same conditions without any templates exhibited antiferromagnetic characteristics (shown in Fig. 2(b) (inset)). The magnetization at 50 kOe and the coercive field (2Ec) of BFO NPs were 0.305579 emu g−1 and 493.5 Oe, respectively, as a result of the space-modulated spiral spin structure. However, the saturation magnetization of SN-BFO derived from the M–H curve shown in Fig. 2(b) was 6.37316 emu g−1 at 50 kOe, whereas the coercive field was about 40 Oe for SN-BFO (shown in Fig. 2(b) (inset)), which was even weaker than that of BFO NPs, indicating paramagnetism rather than ferromagnetism or antiferromagnetism in the SN-BFO samples. Compared with those of BFO thin films and other nanostructures, the coercive field of SN-BFO is much weaker and more insignificant, which reveals its soft characteristics and better adaptability for device applications. To determine the oxidation states of magnetic Fe ions in SN-BFO, X-ray photoelectron spectroscopy (XPS) was carried out and the recorded XPS spectra of the Fe 2p region of SN-BFO are plotted in Fig. 2(c). In particular, the Fe 2p3/2 peak is centered at 710.1 eV for SN-BFO. In general, the binding energy of Fe 2p3/2 is about 709.5 eV for Fe2+ and 711.0 eV for Fe3+.22 The broadening of the Fe 2p peak could be attributed to the nanoscale size of the SN-BFO powder and the coexistence of Fe2+ and Fe3+ oxidation states in the Fe 2p region with different binding energies. Data fitting of the Fe2+ and Fe3+ peaks demonstrates that the molar ratio of Fe2+ to Fe3+ is about 4:
6, as for other natural polymer-templated BFO,23 whereas it is usually lower than 1
:
2.24 Such an enhancement in magnetization and transformation of the magnetic characteristics are mainly due to the increased Fe2+ content in SN-BFO, which results from the reducing atmosphere during calcination caused by incomplete combustion of the sucrose templates. The improvement in magnetization can therefore be explained as follows: first, the magnetic moment of Fe2+ is smaller than that of Fe3+, so an appropriate proportion of Fe2+ that is present in SN-BFO may seriously disturb the modulated spiral spin structure along the [110] direction with a period of 62 nm by forming a staggered arrangement of Fe2+ and Fe3+ in some regions of the BFO crystal lattice.18 In perfect BFO lattices, the total magnetic moment of Fe3+ in oxygen octahedra is almost zero, which is shown in Fig. 2(d). However, an increase in the Fe2+ content in the BFO lattice results in a change in the total magnetic moment, and a schematic is shown in Fig. 2(d). Second, the volume of Fe2+ is much larger than that of Fe3+, which causes a distortion of the crystal lattice.19 Moreover, the introduction of Fe2+ changes the Fe–O–Fe double-exchange interaction and further influences the electron cloud distribution of the oxygen octahedra to some extent. Thirdly, the ferromagnetic behavior of BFO may also be affected by the change in oxygen vacancies resulting from the large amount of Fe2+.25
Adsorption data for the pollutant rhodamine B (RhB) on SN-BFO and BFO NPs are shown in Fig. 3. Obviously, SN-BFO is more effective for the adsorption of RhB from water than BFO NPs. The adsorption kinetics, which reflect the uptake rate of the solute dyes, are an important feature of adsorption. Fig. 3(a) shows the adsorption kinetic curves of SN-BFO and BFO NPs for removing RhB. Under the test conditions (pH = 3 and a concentration of RhB of 12 mg L−1), the adsorption capability for RhB increased rapidly in the first 10 minutes and then remained unchanged. After 60 min, the adsorption capacities of SN-BFO and BFO NPs reached 30.0 mg g−1 and 10.2 mg g−1, respectively. The adsorption isotherms of RhB on SN-BFO and BFO NPs at various initial concentrations have also been studied (Fig. 3(b)).
With an increase in the RhB concentration from 0 mg L−1 to 100 mg L−1, the adsorption capacities of SN-BFO and BFO NPs were increased to 34.7 mg g−1 and 12.1 mg g−1, respectively. Apparently, SN-BFO exhibits a higher adsorption capacity for RhB than that of BFO NPs. A possible reason is the presence of more basic sites in SN-BFO, which is caused by its surface area being larger (34.25 m2 g−1) than that of BFO NPs (1.9 m2 g−1).
To further study the adsorption behavior of RhB on SN-BFO and BFO NPs, the Langmuir isotherm models have been applied. The Langmuir isotherm is expressed by eqn (1):
![]() | (1) |
The photocatalytic behaviors of SN-BFO and BFO NPs were estimated via the degradation of the typical organic pollutant RhB (rhodamine B) under visible light at a pH value of 7. To eliminate the influence of the adsorption of RhB on SN-BFO, all the solutions were stirred in dark conditions for 60 min. Because the photocatalytic reaction of the degradation of RhB follows first-order kinetics, the curve of ln(C0/C) vs. time exhibits a linear relationship, as shown in Fig. 4(a), and the photocatalytic reaction was repeated 3 times with the same catalysts separately. After illumination with SN-BFO for 180 min, the RhB concentration was reduced to approximately 0.2C0 and the value of the rate constant k reached an average of about 0.0112 min−1. The first-order rate constants (k) of the three photocatalytic reactions were 0.0113 min−1, 0.0109 min−1 and 0.0115 min−1, respectively, which were greater than that of BFO NPs. The relatively high degradation efficiency may be attributed to the small particle size and large specific surface area of SN-BFO. The concentration of RhB after different periods of illumination was confirmed via UV-visible spectra and the absorption spectra of RhB over 180 min are shown in Fig. 4(b). The intensity of the characteristic peak of RhB at 553 nm decreased as the illumination time continued. The absorption peaks decreased considerably throughout the irradiation time and switched to lower wavelengths based on a N-de-ethylation mechanism.26 The inset of Fig. 4(b) shows a shift from 556 nm to 534 nm for the solution under solar-light illumination. To acquire knowledge of the actual degradation process of RhB, the change in the TOC of the RhB solution was measured and is shown in Fig. 4(c). In 240 min of illumination the TOC decreased by 80% and, with an increase in the illumination time to 420 min, all organic species disappeared from the solution. From a comparison of TOC data and the remaining relative concentration of the RhB solution, we found that although after 300 min the relative concentration of RhB in solution was almost zero, TOC still remained in the solution, which was probably because some RhB molecules were adsorbed on the surface of SN-BFO. The TOC in the solution was mainly derived from phenol derivatives or aromatic esters, which originated from the breakdown products of RhB.27,28 To demonstrate the recovery properties of SN-BFO and BFO NPs, a series of magnetic recycling experiments were carried out. Fig. 4(d) shows the relationship between the mass of SN-BFO/BFO NPs and the recycling times after recovery by magnetic separation by pressing a magnetic stirrer close to the reaction beaker and pouring out the liquid phase slowly. From Fig. 4(d), it can be seen that the recycled mass of BFO NPs decreased from 51.8 mg to 26.2 mg, which was a reduction of about 49.4%, after 3 recycling steps. Compared with this, the recycled mass of SN-BFO decreased from 51.8 to 50.5 mg, which was a decrease of only 2.5%, even in the low-intensity magnetic field of the magnetic stirrer.
As a photocatalyst is irradiated by light with energy that is greater than the band gap energy, electrons are excited from the valence band (VB) to the conduction band (CB), producing holes in EV. The holes may attract electrons from adsorbed pollutants or react with H2O to form hydroxyl radicals, whereas electrons can reduce absorbed molecular oxygen to yield superoxide anion radicals (O2−˙), which further form hydroxyl radicals via chain reactions.32 Both types of radicals may initiate redox reactions with organic molecules adsorbed on the surface of the photocatalyst, which leads to the destruction of pollutants. Furthermore, the dye-sensitized effect, namely, that most of the electrons in the EC of SN-BFO may be derived from RhB*, should be taken into consideration. In fact, different BFO materials prepared by different methods may have diverse energy band structures, especially band gaps (Eg ∼ 2.0–2.9 eV). Because of this, a UV-vis absorbance experiment on SN-BFO was first carried out. As shown in Fig. 5(a), SN-BFO exhibits strong photoabsorption properties in the visible-light region. The inset of Fig. 5(a) shows that the estimated band gap of SN-BFO is about 2.0 eV, which was calculated by using Tauc's equation.29 Based on the estimated energy band structure of SN-BFO, another experiment was designed to test whether RhB* can inject electrons into the conduction band of SN-BFO. Monochromatic light at 620 nm and 0.2 W cm−2 was used to illuminate SN-BFO, as shown in Fig. 5(b). It was easily found that the degradation rate of RhB was much slower than when illuminated by solar light with the same power. From the UV absorption spectra of RhB solution, it can be discovered that only light in the range from 460 to 590 nm can excite RhB from the ground state to the excited state (RhB*). On the basis of the above facts, the degradation of RhB under monochromatic light at 620 nm is mainly due to photogenerated radicals in SN-BFO. However, the obvious decrease in the degradation rate of RhB also reflects the existence of the dye-sensitized effect, which is equally important in SN-BFO photocatalysis. For a fundamental investigation of which radical plays a key role in the photocatalytic process of SN-BFO, KI and t-BuOH were used as scavengers of different radicals. After the addition of KI (a quencher of positive holes and hydroxyl radicals on the catalyst surface30), the inhibition of photocatalysis in Fig. 5(b) indicates that the degradation of RhB was mainly achieved by positive holes and hydroxyl radicals on the surface of SN-BFO. Moreover, t-BuOH was used as a scavenger of hydroxyl radicals and the degradation rate changed little, which indicates that positive holes may play a major role in the photocatalysis process, as shown schematically in Fig. 5(c). On the one hand, the photogenerated holes combine with ground-state RhB to form RhB+, which first undergoes a process of N-de-ethylation and fragments into phenol derivatives or aromatic esters,27 finally degrading to CO2 and H2O. On the other hand, when illuminated by visible light, RhB may enter into the excited state and inject electrons into the conduction band of SN-BFO, being transformed into RhB+, which proceeds via the same degradation mechanism as mentioned above.
From the perspective of environmental purification, SN-BFO may play a dual role, namely, when the concentration of an organic pollutant is relatively low SN-BFO can degrade it by a photocatalytic method and when the concentration of an organic pollutant is high SN-BFO can act as an adsorbent to adsorb it and then be magnetically recycled.
This journal is © The Royal Society of Chemistry 2016 |