Next Article in Journal
CVD Elaboration of 3C-SiC on AlN/Si Heterostructures: Structural Trends and Evolution during Growth
Next Article in Special Issue
Synthesis of g-C3N4 from Various Precursors for Photocatalytic H2 Evolution under the Visible Light
Previous Article in Journal
A Novel Model of Semiconductor Porosity Medium According to Photo-Thermoelasticity Excitation with Initial Stress
Previous Article in Special Issue
Growth and Spectral Properties of Er3+ and Yb3+ Co-Doped Bismuth Silicate Single Crystal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Photocatalytic Activities for Degradation of Dyes and Drugs by Crystalline Bismuth Ferrite-Modified Graphene Hybrid Aerogel

1
Yangtze Delta Region Institute (Huzhou), University of Electronic Science and Technology of China, Huzhou 313001, China
2
School of Physics, University of Electronic Science and Technology of China, Chengdu 610054, China
3
National Key Laboratory of Science and Technology on Advanced Composites in Special Environments, and Center for Composite Materials and Structures, Harbin Institute of Technology, Harbin 150080, China
4
Chongqing Research Institute, Harbin Institute of Technology, Chongqing 401151, China
5
College of Materials Science & Engineering, Beijing Institute of Petrochemical Technology, Beijing 102617, China
*
Authors to whom correspondence should be addressed.
Crystals 2022, 12(11), 1604; https://doi.org/10.3390/cryst12111604
Submission received: 24 October 2022 / Revised: 8 November 2022 / Accepted: 9 November 2022 / Published: 10 November 2022
(This article belongs to the Special Issue Advances in Optoelectric Functional Crystalline Materials)

Abstract

:
Industrial wastewater contains diverse toxic dyes and drugs, which pollute the environment and poison creatures. Utilizing photocatalysts has been accepted to be an effective method to degrade water pollutions using solar light. Crystalline bismuth ferrite (Bi2Fe4O9) with a band gap of 1.9–2.0 eV is expected to be one of the most promising candidates for photocatalysts in the visible light region. Amorphous graphene is also a promising candidate as a photocatalyst owing to its excellent electronic and optical properties. Herein, a composite of Bi2Fe4O9/graphene aerogels (GAs) was prepared with a two-step hydrothermal method. The prepared Bi2Fe4O9 powders were confirmed to be successfully doped into GAs and evenly dispersed between graphene sheets. The Bi2Fe4O9/GA composite was utilized to perform photodegradation for organic dyes and antibiotic drugs under visible light irradiation, yielding efficiencies of 90.22%, 92.3%, 71.8% and 78.58% within 330 min for methyl orange, methylene blue, Rhodamine B and tetracycline hydrochloride, respectively. Such distinct photocatalytic activities overwhelmed the pure Bi2Fe4O9 powders of 14.10%, 22.19%, 13.98% and 48.08%, respectively. Additionally, the composite produced a degradation rate constant of 0.00623 min−1 for methylene blue, which is significantly faster than that of 0.00073 min−1 obtained by the pure powder. These results provide an innovative strategy for designing 3D visible-light-responsive photocatalysts combined with graphene aerogel for water purification.

1. Introduction

The utilization of diverse synthetic dyes and drugs has been promoted by modern industries to gain economic benefits, such as in the paper, textile, printing and dyeing, pharmaceutical, cosmetic, and food and beverage industries. Massive consumer markets continue to stimulate industrial production in modern society. Various antibiotic drugs, about 1–15% of dyes and drugs, are directly discarded and released into wastewater. Such improper handling processes contaminate the surface and groundwater, leading to serious environmental pollution which poison human beings and other creatures [1,2]. Dyes such as methyl orange (MO), methyl blue (MB) and rhodamine B (RhB) and drugs such as tetracycline hydrochloride (TC-HCl) have been confirmed to present potential carcinogenic effects on life. It has been reported that MO dye can cause vomiting and diarrhea, and high levels of exposure to MO dye can result in death [3]; MB dye can induce oxidative damage of DNA when photosensitized by white light [4]; RhB dye can cause respiratory tract infections, skin and gastrointestinal tract irritations and eye infections, with developmental toxicity in animals and humans [5]; and TC-HCl drugs present clastogenic and mutagenic activities in mammalian cells [6]. In order to effectively remove dyes from wastewater, research efforts have been exploring various methods including precipitation, adsorption, and biological treatment [7,8,9]. However, numerous dyes and drugs, including the abovementioned four, present high stability and solubility in water and high stability to light, temperature and chemicals; hence, it is ineffective to remove them from aqueous solutions by common water purification or treatment methods such as adsorption and precipitation.
Photocatalytic degradation has been accepted to be an effective technique for removing organic pollutants [10,11,12,13,14]. Although titanium dioxide (TiO2) acts as the most outstanding photocatalyst for oxidative decomposition under UV irradiation because of its high photocatalytic performance and environmental friendliness [15,16], it encounters some inherent defects that limit its application in photocatalysis, such as a high electron–hole recombination rate and ultraviolet light response. Current research into photocatalysts focuses on developing high activities under visible light irradiation in a wide band gap range. Many strategies have been adopted to improve the visible light absorption performance of TiO2, including doping [17,18], surface modification [19,20], and decoration with plasmonic nanostructures to apply the surface plasmon effect [21,22]. These strategies can greatly improve the photocatalytic performance of TiO2. Especially for photocatalytic degradation, numerous studies have explored treatment methods for wastewater using the oxidation reaction of wide-band gap semiconductor nanomaterials to generate active -OH radicals under UV and visible light irradiation [23,24,25]. The secondary electrons and holes generated by the semiconductor nanomaterials interact with the oxygen molecules and water molecules of the dye to generate free radicals, which leads to the degradation of organic chemicals. Importantly, the higher-energy band gap of semiconductor nanomaterials prevents electron–hole recombination, enhancing their interaction with oxygen and water molecules, and fully mineralizing organic pollutants in water [26,27,28,29,30].
Bi-based compounds have attracted research attention in photocatalysis due to their unique properties, such as suitable optical band gaps and high chemical stability [31]. Bismuth ferrite (Bi2Fe4O9) is a visible-light active photocatalyst with a band gap of 1.9–2.0 eV, and its multiferroic, magnetic and sensing properties have received extensive attention. Bi2Fe4O9 (space group Pbam; lattice constants a = 7.965 Å, b = 8.440 Å, c = 5.994 Å) is paramagnetic at room temperature and transforms into an antiferromagnetic state with the Néel temperature (TN) of 64 K [32]. Single-, nano-, and sub-micron-sized Bi2Fe4O9 crystals have been synthesized via hydrothermal techniques, and Bi2Fe4O9 crystals with different orientations and morphologies exhibit different photocatalytic efficiencies [33,34,35]. Bi2Fe4O9 sub-microcrystals were synthesized with tunable morphologies via the hydrothermal method [36], and Bi2Fe4O9 nanofibers synthesized via the electrospinning method combined with the sol–gel method present antioxidation and exhibit the methyl orange (MO) degradation effect [37]. Bi2Fe4O9 is a promising visible light catalytic material to degrade various pollutants in the environment; however, the degradation efficiency requires further improvement due to the inherent poor electrical conductivity [38].
In recent years, graphene is considered to be the most promising carbon two-dimensional material which is used in the whole industry due to its exceptional properties. Its excellent optical behavior, electrical conductivity, thermal conductivity and mechanical strength and electromagnetic properties enable it to be widely used in energy storage, catalysis and engineering fields [39]. Additionally, graphene exhibits a linear dispersive electronic band structure that contributes to excellent light absorption capacity, which presents an attractive application prospect in the field of photocatalysis. Graphene, reduced graphene oxide (rGO) and graphene-analogous two-dimensional (2D) material have been accepted as effective photocatalysts in photodegradation for pollutants, CO2 conversion and photoenergy conversion [40,41,42,43]. Previous works reported the combination of porous Bi2Fe4O9 microspheres with reduced graphene oxide (rGO) can enhance light reflection, scattering and absorption properties, which is expected to benefit the photocatalytic reaction processes [44], and crystalline Bi2Fe4O9 was used to modify 2D graphene for the photodegradation of MO [45], methyl violet [46,47] and bisphenol A [48]; however, light absorption efficiencies on Bi2Fe4O9-based photocatalysts are still limited on 2D graphenic materials. Graphene aerogels (GAs) are three-dimensional (3D) scaffold materials that show great promise and have been investigated intensely [49,50,51], because GAs show higher electrical conductivity and optical absorbance than those which have been prepared via the dispersion of 2D graphene sheets. Crystalline semiconductor-decorated GAs have been explored as photocatalysts to degrade water pollutants [52,53]. However, the photocatalytic properties of Bi2Fe4O9- decorated GAs on composite have been scarcely reported to date.
In this work, a novelty composite of Bi2Fe4O9 and GAs (denoted Bi2Fe4O9/GA) was constructed and successfully fabricated via a facile hydrothermal method and freeze-drying process for the photocatalytic degradation of dyes and drugs in sewage. On the basis of characterizing the structure, morphology and composition of the photocatalyst, the degradation of dyes and drugs under visible light irradiation was performed. The photocatalytic activities on methyl orange (MO), methylene blue (MB), rhodamine B (RhB) and tetracycline hydrochloride (TC-HCl) were evaluated. The composite of Bi2Fe4O9/GA showed more efficient visible-light-driven photodegradation activities and cycling stability compared with the pure Bi2Fe4O9 powder. The construction of crystalline Bi2Fe4O9 modified 3D graphene aerogels potentially impacts the designs of more efficient photocatalysts for pollution degradation.

2. Experimental Section

2.1. Materials

Natural graphite powder (purity ≥ 99.95%), bismuth nitrate (Bi(NO3)3 H2O), ferric nitrate (Fe(NO3)3 9H2O), nitric acid (HNO3), sodium hydroxide (NaOH), methyl orange (MO), methylene blue (MB), rhodamine B (RhB) and tetracycline hydrochloride (TC-HCl) were purchased from Shanghai Aladdin Industrial Co., Ltd., Shanghai, China.

2.2. Synthesis of Bi2Fe4O9/GA

Bi2Fe4O9 powder was synthesized based on a previously reported procedure [54,55]. A typical process is briefly described below and illustrated in Scheme 1. Prepare solution A: add 2.43 g of Bi(NO3)3 5H2O and 2.02 g of Fe(NO3)3 9H2O into 5 mL of dilute HNO3 (1 M) solution. Prepare solution B: dissolve 6.20 g of NaOH in 20 mL of H2O. Add solution B dropwise to solution A and stir the mixed solution for 10 min. Then, transfer the mixture into a stainless autoclave with a filling degree of 80% and seal the autoclave. Maintain the reaction at 180 °C for 24 h and cool the autoclave to room temperature. Filter the orange–yellow product solution, and wash several times with distilled water and ethanol to remove unwanted components such as sodium ions and nitrate. Then, allow it to dry at 60 °C for 6 h in air to obtain Bi2Fe4O9 powder. It is important to note that the synthesis of pure Bi2Fe4O9 should be carried out in an alkaline condition to avoid the coexistence of Bi25FeO40 and BiFeO3, because Bi25FeO40 forms at pH < 8.5.
The Bi2Fe4O9/GA composite was synthesized based on previous methods [56]. Prepare a graphene aerogel dispersion: add 5 mL of dispersion solution of graphene oxide (5.0 mg mL−1) into a clean beaker, add 62.5 mg of ascorbic acid as the reducing agent, and add 50 mg of pre-synthesized Bi2Fe4O9 powder into the beaker. Maintain the reaction at 140 °C for 10 h to obtain the Bi2Fe4O9/graphene hydrogel composite and wash several times with distilled water. Then, freeze-dry it for 48 h to form the Bi2Fe4O9/GA composite.

2.3. Characterization

X-ray diffraction (XRD) was recorded on a Shimadzu XRD-7000 diffractometer (Shimadzu, Kyoto, Japan). A field-emission scanning electron microscopy (FE-SEM, S-4800, Hitachi, Tokyo, Japan) was used to observe the morphology of the synthesized samples. A Shimadzu UV-3600 spectrophotometer equipped with an integrating sphere was applied to measure the diffuse reflection spectra (DRS), using BaSO4 as a reference.

2.4. Photocatalytic Degradation Performance

The prepared Bi2Fe4O9/GA composite was used to photocatalytically degrade dyes and drugs. Dyes of methyl orange (MO), methylene blue (MB) and Rhodamine B (RhB) and drugs of tetracycline hydrochloride (TC-HCl) were selected as the degradation targets. The prepared dye or drug solutions were detected under the action of visible light (an xenon lamp, 500 W). The concentration of the dye or drug solution was varied to evaluate the photocatalytic activity. The specific experiments were performed as follows: 2.0 mg of the Bi2Fe4O9/GA composite was added into 10 mL of TC-HCl (20 mgL−1) and placed in a dark environment for 30 min to reach the equilibrium on the photocatalyst’s surface. The change in concentration of each dye or drug at the wavelength of the maximum absorption (465 nm for MO, 668 nm for MB, 568 nm for RhB and 365 nm for TC-HCl) was measured with an ultraviolet–visible spectrometer (UV-2600).

3. Results and Discussion

3.1. Structural, Morphological, Optical and Chemical Characterization

Typical XRD patterns of the prepared Bi2Fe4O9 powder and Bi2Fe4O9/GA composite are presented in Figure 1. The diffraction peaks of the Bi2Fe4O9 powder at 2θ = 28.5, 29.3, 34.0, 47.4 and 56.9°, respectively, can be indexed to the (121), (211), (130), (141) and (332) crystal planes of the orthorhombic Pbam phase, which are consistent with the standard data (JCPDS No. 25-0090) [57]. The XRD patterns of the graphene presented a broad peak centered at 24.9°, which was correlated to the inter-planar spacing of 3.68 Å [58]. The obtained wide peak also shows that the graphene was loosely stacked without a strong lattice periodic structure in crystalline graphite [59]. The diffraction peaks of the Bi2Fe4O9/GA composite (red line in Figure 1) were similar to those of the pure Bi2Fe4O9 sample (black line in Figure 1), but new diffraction peaks appeared at 2θ = 26.6 and 44.8°, which could be indexed to the (201) and (132) crystal planes. Such phenomena demonstrate that the crystal structure of Bi2Fe4O9 in the composite did not change even when prepared in the environment containing graphene but increased during the formation of the composite. This suggests that graphene was well coupled to the surface of the Bi2Fe4O9 powder, rather than entering into the interior of Bi2Fe4O9 so as to change the lattice periodicity. It is worth noting that the Bi2Fe4O9/GA composite did not present a broad peak in the range of 10° to 20°, which was possibly caused by the amorphous phase transition of graphene during the hydrothermal process. Additionally, the diffraction peaks of graphene aerogel disappeared in the Bi2Fe4O9/GA composite, which indicated that the graphitic structure was recovered after the hydrothermal reaction because of the reduction of graphene aerogel.
The SEM images of the prepared Bi2Fe4O9/GA composite are shown in Figure 2a–c. Figure 2a presents a hierarchically interconnected porous structure, in which randomly arranged graphene sheets overlap together. Such a hierarchical and porous structure contributes to the processes of multiple reflection and the absorption of light, thus promoting the mass transfer during the photocatalytic reaction and improving the photocatalytic performance. Figure 2b shows that Bi2Fe4O9 particles anchored well on the surface of graphene sheets in the composite aerogel. Figure 2c clearly shows that the Bi2Fe4O9 particles have irregular lamellar morphologies with thicknesses of about 200 nm. The EDX energy spectrum in Figure 2d suggests that Fe and Bi elements exist in the Bi2Fe4O9/GA composite. The element mapping of Figure 2e indicates that the Bi2Fe4O9 particles dispersed uniformly on the graphene surface, demonstrating the existence of Bi2Fe4O9 and graphene in the composite.
The photoluminescence spectra of the prepared Bi2Fe4O9/GA composite are shown in Figure 3, giving three emission peaks at ~491, ~580 and ~621 nm. The UV-vis diffuse reflectance spectrum (UV-vis DRS) can be found from the Supplementary Material (Figure S1), which showed an extinct shielding effect on the visible and near-infrared regions by graphene aerogels. This evidence confirmed that graphene aerogels plays an essential role in modifying the optical properties of the pure Bi2Fe4O9 powders.

3.2. Photodegradation Performance

Figure 4 shows the photocatalytic degradation of MO, MB and RhB dyes and the TC-HCl drug over the Bi2Fe4O9/GA composite. All samples were kept in dark conditions for 30 min to reach an adsorption desorption equilibrium before the illumination. The UV-visible spectra in Figure 4a–d show that under visible light irradiation, the characteristic absorption peaks of dyes and drugs, 465 nm for MO, 668 nm for MB, 568 nm for RhB and 365 nm for TC-HC, decreased rapidly from aqueous solutions until they were completely degraded within 330 min. The spectral characteristics confirmed that the Bi2Fe4O9/GA composite exhibited effective photodegradation activities across a wide wavelength range from 350 to 750 nm. Figure 4e–h show the concentration ratio (C/C0) changes for the four types of dyes and drug on Bi2Fe4O9/GA as a function of irradiation time (t) along with the blank experiment result. On the Bi2Fe4O9 powder, MO, MB, RhB and TC-HCl were only degraded by 14.10%, 22.19%, 13.98% and 48.08%, respectively, within 330 min, whilst on the Bi2Fe4O9/GA composite, these dyes and drug degraded by 90.22%, 92.2%, 71.8% and 78.54% with significantly more rapid photodegradation rates. These results indicate that the photocatalytic performance of the as-prepared Bi2Fe4O9 powder photocatalyst was remarkably improved by combining it with graphene aerogels. In addition, the photodegradation on MO took 1 h to yield a 70% degradation percentage, significantly faster than the reported 5 h by the Bi2Fe4O9/graphene composite, demonstrating the advantage provided by a three-dimensional photocatalyst [45].
Figure 5 shows the kinetic fitting results of the dyes and drug degradation activities over the Bi2Fe4O9/GA composite and the pure Bi2Fe4O9 powder. The fitting curves presented linear relationships for all the dyes and drug, confirming that the photocatalytic degradation processes obeyed the law of first-order kinetic reactions. The degradation rate constant k was evaluated from the fitting curves. Figure 6 and Table 1 illustrate that the k values of the dyes and drug by the Bi2Fe4O9/GA composite (ca. 0.00356~0.00586 min−1) are significantly larger than that by the Bi2Fe4O9 powder (ca. 0.00044~0.00174 min−1). For the three dyes, the Bi2Fe4O9/GA composite yielded more than 8-fold higher k values than the pure Bi2Fe4O9 powder. Especially for MO, the rate constant ratio of the composite to the pure powder was over 13, which demonstrates the graphene aerogels did enhance the photodegradation activity on the Bi2Fe4O9 powder. In addition, the TC-HCl drug exhibited relatively less difference in k between the two photocatalytic systems, which might be ascribed to the special chemical structure of TC-HCl that exhibits more polarity than the three dyes. Further studies on photodegradation mechanisms will be examined by in situ surface-enhanced Raman spectroscopy to explore degradation pathways induced by photocatalysts.
Compared to the pure Bi2Fe4O9 powder, the enhanced photocatalytic degradation properties by the Bi2Fe4O9/GA composite can be attributed to the enhanced light absorption capacity and porous structure. First, the modification of Bi2Fe4O9 on graphene could extend the light absorption range to the visible light region and increase light reflection and absorption between the graphene layers [60], which greatly improved the absorption coefficient of the solar spectrum and thus produced a higher concentration of photogenerated electron–hole pairs under the same irradiation condition (Scheme 2). Second, after introducing graphene, the Bi2Fe4O9/GA composite exhibited a porous and hierarchical structure. Such structure not only enhanced the absorption of the dye or drug molecules on the photocatalyst’s surface, but also accelerated the mass transfer during photocatalytic degradation reaction processes, meaning the reactants could rapidly attach to the photocatalyst surface while the products could rapidly separate from the reaction sites. Additionally, photoexcited electrons from Bi2Fe4O9 could rapidly transfer into graphene aerogels, leaving holes to oxide dye or drug molecules absorbed on the composite’s surface (Scheme 2), thus promoting the photocatalytic performance. Further experiments on surface area and the porosity distribution of Bi2Fe4O9/GA composite will be performed via Brunauer–Emmett–Teller (BET) analysis, and the complex products of photodegradation will be examined via liquid chromatography–mass spectrometry in the future.

4. Conclusions

In this work, based on the controllable synthesis of graphene aerogels, a graphene aerogel–semiconductor photocatalyst composite of Bi2Fe4O9/GA was prepared via a hydrothermal method through the visible-light-available photocatalyst of Bi2Fe4O9. The controllable preparation of the photocatalysts and the related structural and morphology characteristics were carried out and determined via XRD and SEM. The Bi2Fe4O9 particles in the prepared Bi2Fe4O9/GA composite were confirmed to be successfully doped into the graphene aerogel. The Bi2Fe4O9/GA composite was utilized to examine the photocatalytic degradation activities for dyes of MO, MB, RhB and the antibiotic drug of TC-HCl under visible light irradiation. The photocatalytic degradation efficiency of 90.22%, 92.3%, 71.8% and 78.58% for MO, MB, RhB and TC-HCl by the Bi2Fe4O9/GA composite was significantly higher than that by the Bi2Fe4O9 powder within 330 min, whilst by the pure Bi2Fe4O9 powder, the degradation efficiencies were 14.10%, 22.19%, 13.98% and 48.08%, respectively. Additionally, the Bi2Fe4O9/GA composite produced a degradation rate constant of 0.00623 min−1 for MB, which is significantly faster than that of 0.00073 min−1 by the pure Bi2Fe4O9 powder. The enhanced photodegradation abilities can be attributed to the synergistic effect between graphene and the crystalline Bi2Fe4O9 particles with enhanced light absorption and the efficient separation of photo-induced charge carriers by the presence of graphene flakes in the aerogel. This work paved a new way for effective photodegradation for wastewater using 3D graphene-based photocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst12111604/s1, Figure S1: UV-vis DRS spectrum of the prepared Bi2Fe4O9/GA composite.

Author Contributions

Conceptualization G.X.; Supervision, L.Q. and W.H.; Methodology, Y.J. and Y.D.; Investigation, Writing—Original Draft Preparation, Y.Z. and M.X.; Formal Analysis, Y.Z., X.L. and P.X.; Writing—Review and Editing, X.L. and P.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos. 11774044 and 22202033), the Young Leading Talents of Nantaihu Talent Program in Huzhou (2022), Research start-up funding in Yangtze Delta Region Institute (Huzhou) of UESTC (No. U03220089), the Foundation of Sichuan Excellent Young Talents (2021JDJQ0015), Fundamental Research Funds for the Central Universities (ZYGX2020J023) and the Natural Science Foundation of Chongqing (CSTB2022NSCQ-MSX0441).

Data Availability Statement

The data that support the findings of this study are available from the corresponding authors on reasonable request.

Conflicts of Interest

The authors declare no competing interests. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Shindhal, T.; Rakholiya, P.; Varjani, S.; Pandey, A.; Ngo, H.H.; Guo, W.; Ng, H.Y.; Taherzadeh, M.J. A critical review on advances in the practices and perspectives for the treatment of dye industry wastewater. Bioengineered 2021, 12, 70–87. [Google Scholar] [CrossRef] [PubMed]
  2. Muz, M.; Dann, J.P.; Jäger, F.; Brack, W.; Krauss, M. Identification of mutagenic aromatic amines in river samples with industrial wastewater impact. Environ. Sci. Technol. 2017, 51, 4681–4688. [Google Scholar] [CrossRef]
  3. Farhan Hanafi, M.; Sapawe, N. A review on the water problem associate with organic pollutants derived from phenol, methyl orange, and remazol brilliant blue dyes. Mater. Today Proc. 2020, 31, A141–A150. [Google Scholar] [CrossRef]
  4. Olliver, J.R.; Wild, C.P.; Sahay, P.; Dexter, S.; Hardie, L.J. Chromoendoscopy with methylene blue and associated DNA damage in Barrett’s oesophagus. Lancet 2003, 362, 373–374. [Google Scholar] [CrossRef]
  5. Al-Gheethi, A.A.; Azhar, Q.M.; Senthil Kumar, P.; Yusuf, A.A.; Al-Buriahi, A.K.; Radin Mohamed, R.M.S.; Al-shaibani, M.M. Sustainable approaches for removing Rhodamine B dye using agricultural waste adsorbents: A review. Chemosphere 2022, 287, 132080. [Google Scholar] [CrossRef]
  6. Hagiwara, M.; Watanabe, E.; Barrett, J.C.; Tsutsui, T. Assessment of genotoxicity of 14 chemical agents used in dental practice: Ability to induce chromosome aberrations in Syrian hamster embryo cells. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2006, 603, 111–120. [Google Scholar] [CrossRef] [PubMed]
  7. Bagyo, A.N.M.; Arai, H.; Miyata, T. Radiation-induced decoloration and sedimentation of colloidal disperse dyes in water. Appl. Radiat. Isot. 1997, 48, 175–181. [Google Scholar] [CrossRef]
  8. Rai, H.S.; Bhattacharyya, M.S.; Singh, J.; Bansal, T.K.; Vats, P.; Banerjee, U.C. Removal of dyes from the effluent of textile and dyestuff manufacturing industry: A review of emerging techniques with reference to biological treatment. Crit. Rev. Environ. Sci. Technol. 2005, 35, 219–238. [Google Scholar] [CrossRef]
  9. Gibson, L.T. Mesosilica materials and organic pollutant adsorption: Part B removal from aqueous solution. Chem. Soc. Rev. 2014, 43, 5173–5182. [Google Scholar] [CrossRef]
  10. Viswanathan, B. Photocatalytic degradation of dyes: An overview. Curr. Catal. 2018, 7, 99–121. [Google Scholar] [CrossRef]
  11. Antoniadou, M.; Falara, P.P.; Likodimos, V. Photocatalytic degradation of pharmaceuticals and organic contaminants of emerging concern using nanotubular structures. Curr. Opin. Green Sustain. Chem. 2021, 29, 100470. [Google Scholar] [CrossRef]
  12. Velempini, T.; Prabakaran, E.; Pillay, K. Recent developments in the use of metal oxides for photocatalytic degradation of pharmaceutical pollutants in water—A review. Mater. Today Chem. 2021, 19, 100380. [Google Scholar] [CrossRef]
  13. Zhu, L.; Ye, R.; Tang, P.T.; Xia, P. Facile synthesis of Ti3C2 MXene for enhanced adsorption performance. Aust. J. Chem. 2022. accepted. [Google Scholar] [CrossRef]
  14. Khan, B.; Raziq, F.; Bilal Faheem, M.; Umar Farooq, M.; Hussain, S.; Ali, F.; Ullah, A.; Mavlonov, A.; Zhao, Y.; Liu, Z.; et al. Electronic and nanostructure engineering of bifunctional MoS2 towards exceptional visible-light photocatalytic CO2 reduction and pollutant degradation. J. Hazard. Mater. 2020, 381, 120972. [Google Scholar] [CrossRef] [PubMed]
  15. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  16. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  17. Daghrir, R.; Drogui, P.; Robert, D. Modified TiO2 for environmental photocatalytic applications: A review. Ind. Eng. Chem. Res. 2013, 52, 3581–3599. [Google Scholar] [CrossRef]
  18. Zaleska, A. Doped-TiO2: A review. Recent Pat. Eng. 2008, 2, 157–164. [Google Scholar] [CrossRef]
  19. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C Photochem. Rev. 2013, 15, 1–20. [Google Scholar] [CrossRef]
  20. Low, J.; Cheng, B.; Yu, J. Surface modification and enhanced photocatalytic CO2 reduction performance of TiO2: A review. Appl. Surf. Sci. 2017, 392, 658–686. [Google Scholar] [CrossRef]
  21. Suzuki, K.; Li, X.; Wang, Y.; Nagasawa, F.; Murakoshi, K. Active intermediates in plasmon-induced water oxidation at Au nanodimer structures on a single crystal of TiO2. ACS Energy Lett. 2020, 5, 1252–1259. [Google Scholar] [CrossRef]
  22. Raziq, F.; Humayun, M.; Ali, A.; Wang, T.; Khan, A.; Fu, Q.; Luo, W.; Zeng, H.; Zheng, Z.; Khan, B.; et al. Synthesis of S-Doped porous g-C3N4 by using ionic liquids and subsequently coupled with Au-TiO2 for exceptional cocatalyst-free visible-light catalytic activities. Appl. Catal. B Environ. 2018, 237, 1082–1090. [Google Scholar] [CrossRef]
  23. Reza, K.M.; Kurny, A.; Gulshan, F. Parameters affecting the photocatalytic degradation of dyes using TiO2: A review. Appl. Water Sci. 2017, 7, 1569–1578. [Google Scholar] [CrossRef] [Green Version]
  24. Szczepanik, B. Photocatalytic degradation of organic contaminants over clay-TiO2 nanocomposites: A review. Appl. Clay Sci. 2017, 141, 227–239. [Google Scholar] [CrossRef]
  25. Konstantinou, I.K.; Albanis, T.A. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: Kinetic and mechanistic investigations: A review. Appl. Catal. B Environ. 2004, 49, 1–14. [Google Scholar] [CrossRef]
  26. Qi, K.; Cheng, B.; Yu, J.; Ho, W. A review on TiO2-based Z-scheme photocatalysts. Chin. J. Catal. 2017, 38, 1936–1955. [Google Scholar] [CrossRef]
  27. Gong, Y.; Wu, Y.; Xu, Y.; Li, L.; Li, C.; Liu, X.; Niu, L. All-solid-state Z-scheme CdTe/TiO2 heterostructure photocatalysts with enhanced visible-light photocatalytic degradation of antibiotic waste water. Chem. Eng. J. 2018, 350, 257–267. [Google Scholar] [CrossRef]
  28. Zhang, S.; Wang, J.; Wen, S.; Jiang, M.; Xiao, H.; Ding, X.; Wang, N.; Li, M.; Zu, X.; Li, S.; et al. Approaching charge separation efficiency to unity without charge recombination. Phys. Rev. Lett. 2021, 126, 176401. [Google Scholar] [CrossRef]
  29. Qiao, L.; Xiao, H.Y.; Meyer, H.M.; Sun, J.N.; Rouleau, C.M.; Puretzky, A.A.; Geohegan, D.B.; Ivanov, I.N.; Yoon, M.; Weber, W.J.; et al. Nature of the band gap and origin of the electro-/photo-activity of Co3O4. J. Mater. Chem. C 2013, 1, 4628–4633. [Google Scholar] [CrossRef]
  30. Ikram, M.; Haider, A.; Imran, M.; Haider, J.; Naz, S.; Ul-Hamid, A.; Nabgan, W.; Mustajab, M.; Shahzadi, A.; Shahzadi, I.; et al. Facile synthesis of starch and tellurium doped SrO nanocomposite for catalytic and antibacterial potential: In silico molecular docking studies. Int. J. Biol. Macromol. 2022, 221, 496–507. [Google Scholar] [CrossRef]
  31. Nakamura, H.; Ishii, S.i.; Yamada, K.; Matsushima, S.; Arai, M.; Kobayashi, K. First-principles energy band calculation for CaBi2O4 with monoclinic structure. Mater. Chem. Phys. 2010, 121, 385–389. [Google Scholar] [CrossRef]
  32. Pooladi, M.; Sharifi, I.; Behzadipour, M. A review of the structure, magnetic and electrical properties of bismuth ferrite (Bi2Fe4O9). Ceram. Int. 2020, 46, 18453–18463. [Google Scholar] [CrossRef]
  33. Ruan, Q.-J.; Zhang, W.-D. Tunable morphology of Bi2Fe4O9 crystals for photocatalytic oxidation. J. Phys. Chem. C 2009, 113, 4168–4173. [Google Scholar] [CrossRef]
  34. Wu, T.; Liu, L.; Pi, M.; Zhang, D.; Chen, S. Enhanced magnetic and photocatalytic properties of Bi2Fe4O9 semiconductor with large exposed (001) surface. Appl. Surf. Sci. 2016, 377, 253–261. [Google Scholar] [CrossRef]
  35. Ma, H.; Liu, Q.; Zhang, B.; Zhang, R.; Chen, Z.; Li, Y.; Shao, W.; Wu, X.; Zhang, Y.; Ye, W. Hydrothermal synthesis of single-crystalline Bi2Fe4O9 photocatalyst with enhanced visible-light photocatalytic activity. J. Nanosci. Nanotechnol. 2019, 19, 7859–7865. [Google Scholar] [CrossRef]
  36. Zhang, X.; Lv, J.; Bourgeois, L.; Cui, J.; Wu, Y.; Wang, H.; Webley, P.A. Formation and photocatalytic properties of bismuth ferrite submicrocrystals with tunable morphologies. New J. Chem. 2011, 35, 937–941. [Google Scholar] [CrossRef]
  37. Qi, S.; Zuo, R.; Wang, Y.; Chan, H.W.L.-W. Synthesis and photocatalytic performance of the electrospun Bi2Fe4O9 nanofibers. J. Mater. Sci. 2013, 48, 4143–4150. [Google Scholar] [CrossRef]
  38. Wang, W.-S.; Wang, D.-H.; Qu, W.-G.; Lu, L.-Q.; Xu, A.-W. Large ultrathin anatase TiO2 nanosheets with exposed {001} facets on graphene for enhanced visible light photocatalytic activity. J. Phys. Chem. C 2012, 116, 19893–19901. [Google Scholar] [CrossRef]
  39. Lei, Z.; Sathish, C.; Liu, Y.; Karokoti, A.; Wang, J.; Qiao, L.; Vinu, A.; Yi, J. Single metal atoms catalysts—Promising candidates for next generation energy storage and conversion devices. EcoMat 2022, 4, e12186. [Google Scholar] [CrossRef]
  40. Raziq, F.; Aligayev, A.; Shen, H.; Ali, S.; Shah, R.; Ali, S.; Bakhtiar, S.H.; Ali, A.; Zarshad, N.; Zada, A.; et al. Exceptional photocatalytic activities of rGO modified (B,N) co-doped WO3, coupled with CdSe QDs for one photon Z-scheme system: A joint experimental and DFT study. Adv. Sci. 2022, 9, 2102530. [Google Scholar] [CrossRef]
  41. Raziq, F.; He, J.; Gan, J.; Humayun, M.; Faheem, M.B.; Iqbal, A.; Hayat, A.; Fazal, S.; Yi, J.; Zhao, Y.; et al. Promoting visible-light photocatalytic activities for carbon nitride based 0D/2D/2D hybrid system: Beyond the conventional 4-electron mechanism. Appl. Catal. B Environ. 2020, 270, 118870. [Google Scholar] [CrossRef]
  42. Ali, S.; Yasin, G.; Iqbal, R.; Huang, X.; Su, J.; Ibraheem, S.; Zhang, Z.; Wu, X.; Wahid, F.; Ismail, P.M.; et al. Porous aza-doped graphene-analogous 2D material a unique catalyst for CO2 conversion to formic-acid by hydrogenation and electroreduction approaches. Mol. Catal. 2022, 524, 112285. [Google Scholar] [CrossRef]
  43. Raziq, F.; Hayat, A.; Humayun, M.; Baburao Mane, S.K.; Faheem, M.B.; Ali, A.; Zhao, Y.; Han, S.; Cai, C.; Li, W.; et al. Photocatalytic solar fuel production and environmental remediation through experimental and DFT based research on CdSe-QDs-coupled P-doped-g-C3N4 composites. Appl. Catal. B Environ. 2020, 270, 118867. [Google Scholar] [CrossRef]
  44. Lin, Y.; Dai, J.; Yang, H.; Wang, L.; Wang, F. Graphene multilayered sheets assembled by porous Bi2Fe4O9 microspheres and the excellent electromagnetic wave absorption properties. Chem. Eng. J. 2018, 334, 1740–1748. [Google Scholar] [CrossRef]
  45. Xian, T.; Yang, H.; Di, L.J.; Dai, J.F. Graphene-assisted enhancement of photocatalytic activity of bismuth ferrite nanoparticles. Res. Chem. Intermed. 2015, 41, 433–441. [Google Scholar] [CrossRef]
  46. Liu, P.; Sun, H.; Liu, X.; Sui, H.; Zhang, Y.; Zhou, D.; Guo, Q.; Ruan, Y. Enhanced photocatalytic performance of Bi2Fe4O9/graphene via modifying graphene composite. J. Am. Ceram. Soc. 2017, 100, 3540–3549. [Google Scholar] [CrossRef]
  47. Sun, H.; Liu, Y.; Zhang, Y.; Lv, L.; Zhou, J.; Chen, W. Synthesis of Bi2Fe4O9/reduced graphene oxide composite by one-step hydrothermal method and its high photocatalytic performance. J. Mater. Sci. Mater. Electron. 2014, 25, 4212–4218. [Google Scholar] [CrossRef]
  48. Hu, Z.-T.; Liu, J.; Yan, X.; Oh, W.-D.; Lim, T.-T. Low-temperature synthesis of graphene/Bi2Fe4O9 composite for synergistic adsorption-photocatalytic degradation of hydrophobic pollutant under solar irradiation. Chem. Eng. J. 2015, 262, 1022–1032. [Google Scholar] [CrossRef]
  49. Gorgolis, G.; Galiotis, C. Graphene aerogels: A review. 2D Mater. 2017, 4, 032001. [Google Scholar] [CrossRef]
  50. Zhu, C.; Han, T.; Duoss, E.B.; Golobic, A.M.; Kuntz, J.D.; Spadaccini, C.M.; Worsley, M.A. Highly compressible 3D periodic graphene aerogel microlattices. Nat. Commun. 2015, 6, 6962. [Google Scholar] [CrossRef]
  51. Zhang, X.; Sui, Z.; Xu, B.; Yue, S.; Luo, Y.; Zhan, W.; Liu, B. Mechanically strong and highly conductive graphene aerogel and its use as electrodes for electrochemical power sources. J. Mater. Chem. 2011, 21, 6494–6497. [Google Scholar] [CrossRef]
  52. Fan, Y.; Ma, W.; Han, D.; Gan, S.; Dong, X.; Niu, L. Convenient recycling of 3D AgX/graphene aerogels (X= Br, Cl) for efficient photocatalytic degradation of water pollutants. Adv. Mater. 2015, 27, 3767–3773. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, J.-J.; Wu, Y.-H.; Mei, J.-Y.; Zheng, G.-P.; Yan, T.-T.; Zheng, X.-C.; Liu, P.; Guan, X.-X. Synergetic adsorption and photocatalytic degradation of pollutants over 3D TiO2—Graphene aerogel composites synthesized via a facile one-pot route. Photochem. Photobiol. Sci. 2016, 15, 1012–1019. [Google Scholar] [CrossRef] [PubMed]
  54. Xiong, Y.; Wu, M.; Peng, Z.; Jiang, N.; Chen, Q. Hydrothermal synthesis and characterization of Bi2Fe4O9 nanoparticles. Chem. Lett. 2004, 33, 502–503. [Google Scholar] [CrossRef]
  55. Han, J.-T.; Huang, Y.-H.; Wu, X.-J.; Wu, C.-L.; Wei, W.; Peng, B.; Huang, W.; Goodenough, J.B. Tunable synthesis of bismuth ferrites with various morphologies. Adv. Mater. 2006, 18, 2145–2148. [Google Scholar] [CrossRef]
  56. Hu, Z.-T.; Liu, J.-W.; Zhao, J.; Ding, Y.; Jin, Z.; Chen, J.; Dai, Q.; Pan, B.; Chen, Z.; Chen, J. Enhanced BiFeO3/Bi2Fe4O9/H2O2 heterogeneous system for sulfamethoxazole decontamination: System optimization and degradation pathways. J. Colloid Interface Sci. 2020, 577, 54–65. [Google Scholar] [CrossRef]
  57. Zhang, Q.; Gong, W.; Wang, J.; Ning, X.; Wang, Z.; Zhao, X.; Ren, W.; Zhang, Z. Size-dependent magnetic, photoabsorbing, and photocatalytic properties of single-crystalline Bi2Fe4O9 semiconductor nanocrystals. J. Phys. Chem. C 2011, 115, 25241–25246. [Google Scholar] [CrossRef]
  58. Cui, P.; Lee, J.; Hwang, E.; Lee, H. One-pot reduction of graphene oxide at subzero temperatures. Chem. Commun. 2011, 47, 12370–12372. [Google Scholar] [CrossRef]
  59. Shi, Y.; Li, X.; Ye, M.; Hu, C.; Shao, H.; Qu, L. An imperata cylindrical flowers-shaped porous graphene microelectrode for direct electrochemistry of glucose oxidase. J. Electrochem. Soc. 2015, 162, B138. [Google Scholar] [CrossRef]
  60. Lu, M.; Sun, Y.-K.; Yang, S.-H.; Wang, H.-Y.; Guan, X.-H.; Wang, G.-S. Three-dimensional Bi2Fe4O9 nanocubes loaded on reduced graphene oxide for enhanced electromagnetic absorbing properties. Front. Chem. 2020, 8, 608. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration to synthesis Bi2Fe4O9 powder and Bi2Fe4O9/GA composite.
Scheme 1. Schematic illustration to synthesis Bi2Fe4O9 powder and Bi2Fe4O9/GA composite.
Crystals 12 01604 sch001
Figure 1. XRD patterns of GA, Bi2Fe4O9 and Bi2Fe4O9/GA.
Figure 1. XRD patterns of GA, Bi2Fe4O9 and Bi2Fe4O9/GA.
Crystals 12 01604 g001
Figure 2. (ac) SEM images, (d) EDX spectra and (e) mapping images of the prepared Bi2Fe4O9/GA composite.
Figure 2. (ac) SEM images, (d) EDX spectra and (e) mapping images of the prepared Bi2Fe4O9/GA composite.
Crystals 12 01604 g002
Figure 3. Photoluminescence spectrum of the prepared Bi2Fe4O9/GA composite.
Figure 3. Photoluminescence spectrum of the prepared Bi2Fe4O9/GA composite.
Crystals 12 01604 g003
Figure 4. UV-visible spectra for photocatalytic degradation and photodegradation activity of (a,e) MO, (b,f) MB, (c,g) RhB and (d,h) TC-HCl catalyzed by Bi2Fe4O9/GA under visible light irradiation.
Figure 4. UV-visible spectra for photocatalytic degradation and photodegradation activity of (a,e) MO, (b,f) MB, (c,g) RhB and (d,h) TC-HCl catalyzed by Bi2Fe4O9/GA under visible light irradiation.
Crystals 12 01604 g004
Figure 5. Kinetic fitting results of the dye and drug degradation over (a) the Bi2Fe4O9 powder and (b) the Bi2Fe4O9/GA composite.
Figure 5. Kinetic fitting results of the dye and drug degradation over (a) the Bi2Fe4O9 powder and (b) the Bi2Fe4O9/GA composite.
Crystals 12 01604 g005
Figure 6. Rate constants of Bi2Fe4O9/GA and Bi2Fe4O9 for various dyes and drug.
Figure 6. Rate constants of Bi2Fe4O9/GA and Bi2Fe4O9 for various dyes and drug.
Crystals 12 01604 g006
Scheme 2. Schematic illustration of photodegradation mechanism on Bi2Fe4O9/GA composite.
Scheme 2. Schematic illustration of photodegradation mechanism on Bi2Fe4O9/GA composite.
Crystals 12 01604 sch002
Table 1. Evaluated degradation rate constant k (min−1) on the dyes and drug.
Table 1. Evaluated degradation rate constant k (min−1) on the dyes and drug.
PhotocatalystMOMBRhBTC-HCl
Bi2Fe4O9/GA composite0.005860.006230.003560.00461
Bi2Fe4O9 powder0.000440.000730.000440.00174
kcomposite/kpowder13.38.58.12.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, Y.; Xu, M.; Ji, Y.; Dong, Y.; Xing, G.; Xia, P.; Li, X.; He, W.; Qiao, L. Enhanced Photocatalytic Activities for Degradation of Dyes and Drugs by Crystalline Bismuth Ferrite-Modified Graphene Hybrid Aerogel. Crystals 2022, 12, 1604. https://doi.org/10.3390/cryst12111604

AMA Style

Zhao Y, Xu M, Ji Y, Dong Y, Xing G, Xia P, Li X, He W, Qiao L. Enhanced Photocatalytic Activities for Degradation of Dyes and Drugs by Crystalline Bismuth Ferrite-Modified Graphene Hybrid Aerogel. Crystals. 2022; 12(11):1604. https://doi.org/10.3390/cryst12111604

Chicago/Turabian Style

Zhao, Yan, Minghui Xu, Yuanpeng Ji, Yunfa Dong, Guangjian Xing, Pengfei Xia, Xiaowei Li, Weidong He, and Liang Qiao. 2022. "Enhanced Photocatalytic Activities for Degradation of Dyes and Drugs by Crystalline Bismuth Ferrite-Modified Graphene Hybrid Aerogel" Crystals 12, no. 11: 1604. https://doi.org/10.3390/cryst12111604

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop