Next Article in Journal
A Versatile Axle for the Construction of Disassemblage Rotaxanes
Next Article in Special Issue
Immobilization of Glycoside Hydrolase Families GH1, GH13, and GH70: State of the Art and Perspectives
Previous Article in Journal
In Vitro Antileishmanial Activity of Sterols from Trametes versicolor (Bres. Rivarden)
Previous Article in Special Issue
Solid-State Metalloproteins—An Alternative to Immobilisation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Oxidation of Phenol and 2,4-Dichlorophenol by Using Horseradish Peroxidase Immobilized on Graphene Oxide/Fe3O4

College of Resources and Environmental Science, South Central University for Nationalities, Wuhan 430074, China
*
Author to whom correspondence should be addressed.
Molecules 2016, 21(8), 1044; https://doi.org/10.3390/molecules21081044
Submission received: 12 April 2016 / Revised: 30 July 2016 / Accepted: 3 August 2016 / Published: 10 August 2016
(This article belongs to the Special Issue Enzyme Immobilization 2016)

Abstract

:
Graphene oxide/Fe3O4 (GO/Fe3O4) nanoparticles were synthesized by an ultrasonic-assisted reverse co-precipitation method, and then horseradish peroxidase (HRP) was covalently immobilized onto GO/Fe3O4 with 1-ethyl-3-(3-dimethyaminopropyl)carbodiimide (EDC) as a cross-linking agent. In order to enhance the phenol removal efficiency and prevent the inactivation of the enzyme, the polyethylene glycol with highly hydrophilicity was added in this reaction, because the adsorption capacity for the polymer by degradation was stronger than the HRP. The results showed that the immobilized enzyme removed over 95% of phenol from aqueous solution. The catalytic condition was extensively optimized among the range of pH, mass ratio of PEG/phenol as well as initial concentration of immobilized enzyme and H2O2. The HRP immobilized on GO/Fe3O4 composite could be easily separated under a magnetic field from the reaction solution and reused.

Graphical Abstract

1. Introduction

A great amount of industrial wastewaters containing phenolic compounds are generated from coal conversion, resins and plastics processing, coke combustion, and so on. Many phenolic compounds are a potential danger to human health and suspected to be carcinogen [1,2]. One treatment approach is based on the oxidation of phenolic pollutant by natural enzymes [3]. This treatment has many advantages compared with other biological or chemical/physical methods. It is easier than microorganism manipulation for handling and storage of isolated enzymes. Moreover, the specificity of enzymes is much higher with respect to other conventional catalysts.
Horseradish peroxidase (HRP) is a classic heme enzyme containing a ferric protoporphyrin IX prosthetic group. In the presence of hydrogen peroxide (H2O2), HRP catalyzed the oxidation of phenols to form phenoxy radicals [4]. The free radical products spontaneously formed insoluble high molecular weight polymers, which could be removed from solution by filtration or sedimentation [5]. However, during the enzymatic treatment, the generated phenoxy radicals may interact with the active center of the enzyme, leading to shortened catalytic lifetime of the enzyme [6]. In addition, the generated polymers may entrap HRP, decreasing its catalytic ability. In order to overcome the problem of the inactivation of the enzyme, it was reported that the enzyme lifetime was significantly improved by using polyethylene glycol (PEG) or gelatin with high hydrophilicity, which have a greater affinity than HRP for the hydroxyl groups on the growing polymers [7,8]. Another strategy for overcoming the problem of the inactivation of the enzyme was to immobilize enzymes.
It is demonstrated that enzyme immobilization can endow enzymes with some additional advantageous properties [9]. Compared with free enzyme, immobilized enzymes dispersed on the support surface will suffer from less aggregation, which is another inactivation source [10]. A proper enzyme immobilization may produce a strong rigidification of the enzyme structure, especially if a very intense multipoint covalent attachment of enzymes on the support via short spacer arms is achieved [11]. Because all enzyme residues involved in the enzyme immobilization process should maintain their relative positions during any conformational change induced by more drastic conditions, it reduced any conformational change involved in enzyme inactivation and significantly improved the enzyme stability [12,13]. Besides the immobilized enzymes exhibit higher selectiveness and specificity [14]. Nonetheless, enzymes can be denatured and lose their activity simply when solubilized in the immobilization medium and during the immobilization procedure. The immobilization of enzymes may block the active centers and promote diffusion problems [10]. It is also reported that the diffusional limitation (e.g., the generation of gradients of pH) in the porous support immobilized with very active enzyme might provide a profitable tool to improve the operational parameters of enzymes [15]. It is also suggested that the immobilized enzymes can be used repeatedly or continuously in a variety of reactors for the efficient recovery of costly enzymes, and be easily separated from reaction systems for reuse, which make the work-up simple and the protein of the final product uncontaminated [16].
Nanoparticles are considered to be excellent supports for enzyme immobilization because of their minimized diffusional limitations, high specific surface area and high enzyme loading capability, mechanical stability, which are interesting and important features for the immobilization process [17]. Nevertheless, the purity of the enzyme sample only impairs the volumetric activity when using non-porous nanoparticle supports [10]. Moreover, the recovery by centrifugation and filtration of these nanoparticles is very difficult. Magnetic supports show the obvious feature of being ferromagnetic, which allows recovery of the immobilized enzyme preparation from a reactor by simply applying a magnetic field to the slurry, without the need of tedious decantation, centrifugation, or filtration procedures [18]. Fe3O4 magnetic nanoparticles (MNPs) are known to be biocompatible, display no hemolytic activity and genotoxicity with superparamagnetic properties and have been utilized in various fields [19]. It was found that Fe3O4 MNPs have the peroxidase-like property and could activate H2O2 to degrade organic pollutant [20]. Furthermore, Fe3O4 MNPs can feasibly combine with easily modifiable materials, and modify inherently their surface with functional groups to link interested enzymes. Graphene oxide (GO), a derivation of graphene, having a large specific surface area and a range of reactive oxygen functional groups on the surface, not only renders it a good candidate for supporting metal or metal oxide particles, but also provides an ideal substrate for enzyme immobilization [21]. In this study, GO/Fe3O4 MNPs were synthesized with an ultrasonic-assisted reverse co-precipitation method, and then HRP was covalently immobilized onto the GO/Fe3O4 as an enzyme carrier. This composite system could highly efficiently catalyze the reduction of H2O2 to remove phenol and 2,4-dichlorophenol (2,4-DCP).

2. Results and Discussion

2.1. Characterization of Materials

The magnetization curve of GO/Fe3O4 nanoparticles was measured at room temperature, as shown in Figure 1a, which demonstrated that the GO/Fe3O4 exhibited negligible coercivity (1.17 Oe) and remanence (73.06 × 10−3 emu/g), being typical of superparamagnetic materials. The specific saturation magnetization, Ms of the sample, was 25.96 emu/g. This value was smaller than the reported value of bulk Fe3O4 of 68.80 emu·g−1 [19]. Fe3O4 was estimated as 37.7 wt % calculated from the content of GO/Fe3O4. X-ray powder diffraction was used to identify the produced powders as shown in Figure 1b. The broad small characteristic peak of GO around 24° can be observed, which was attributed to C(002) reflection of graphite oxide based on the short-range order in stacked sheets [22]. The diffraction peaks at 11.0° was assigned to the characteristic GO (001) crystal. The characteristic diffraction peaks (2θ) of GO/Fe3O4 at 30.4, 35.6, 43.3, 53.0, 57.1, and 62.8° were ascribed to the (220), (311), (400), (422), (511) and (440) planes of Fe3O4, being similar to the previously reported data [23,24]. For magnetic particles with bound HRP, the same six characteristic peaks were observed. This suggested that the binding process did not result in the phase change of GO/Fe3O4. As a result, the magnetic particles could preserve their magnetic properties during the separation process, which was suitable for application in bioseparation. A TEM image of the prepared immobilized enzyme indicated GO sheets uniformly decorated with nanoparticles with a narrow size distribution, as shown in Figure 1c. The morphology is similar to GO/Fe3O4 nanoparticles synthesized in our previous work [21].

2.2. Reaction System and Effect of PEG

The procedures for the synthesis of GO/Fe3O4 particles and the immobilization of HRP onto GO/Fe3O4 with EDC as a cross-linking agent are illustrated in Scheme 1. The amount of immobilized enzyme was about 5.3 mg/g on the magnetic GO/Fe3O4 nanoparticles. The initial HRP specific activity was 128 U/mg and the measured activity of the immobilized HRP was about 440 U/g, while retained about 65% of initial HRP activity.
As it was known, inactivation is due to the adsorption of enzyme onto the polymers formed during the reaction [8]. In order to reduce the amount of the enzyme, several researchers have suggested that highly hydrophilic additives such as PEG and gelatin can be used to protect the enzyme from inactivation during phenol removal. In Figure 2a, when there is an absence of PEG, the phenol removal efficiency is only about 9%. Within a 3-h reaction period, the removal of phenol achieved 95% in the GO/Fe3O4-HRP system, being much greater than the GO/Fe3O4 and GO/Fe3O4 + PEG system (almost no reduction). Correspondingly, in Figure 2c, about 78% of the total organic carbon (TOC) was found to be removed. As we all know, Fe3O4 and GO are proved that possess enzyme mimetic activity [20,21]. However, the phenol was not removed, because the amount of the H2O2 was fairly small compared to other reports [20].
As shown in Figure 2b, before the mass ratio of PEG/phenol reaches 0.4, they are significantly enhanced with the increase of the amount of PEG added. The further addition of PEG would result in slightly reduction of the phenol removal efficiency, which was in agreement with the result reported by Cheng et al. [25]. Moreover, the phenol removal efficiency was over 95% at this time. Both GO and Fe3O4 possess intrinsic peroxidase-like activity, and we studied the activity of the GO/Fe3O4. The concentration of H2O2 changed when there is only GO/Fe3O4. As the amount of H2O2 is increased, the phenol removal grows in quantity (Figure 2d). It is evident from the Figure 2d that about 76% of the phenol is found to be removed because of the GO/Fe3O4 catalyzed reaction.

2.3. Effect of the Immobilized Enzyme Dose

The catalyst effect of the immobilized enzyme dose is studied, as shown in Figure 3. As the amount of the immobilized enzyme is increased, the phenol removal is significantly promoted until the concentration is about 90 mg/L. The increase of the phenol removal slowed when the dose of immobilized enzyme is higher than 90 mg/L. With HRP concentration increase, more inert intermediate product, phenoxyl radicals, which increase the inactivation chance, will be produced. Therefore, the load of immobilized enzyme was optimized at 90 mg/L.

2.4. Effect of H2O2 Concentration

According to the experimental results in the Figure 4, a conclusion can be made that there is a relationship between total phenols removed and a function of H2O2 provided during the experiment when the molar ratio of H2O2 to phenol is 1.5, regardless of PEG, which is the optimal conditions in this experiment. Some others reported that the optimum molar ratio of H2O2 to phenol was 1.0 or 2.0, which might result from polymers larger than dimmers being produced during the catalytic process [26,27]. Enzymes can be inactivated by too much hydrogen peroxide.

2.5. Degradation of Phenol and 2,4-Dichlorophenol in the Mixture

As shown in Figure 5a, 2,4-dichlorophenol has almost been completely removed totally, but 30% of phenol remains. When phenol and 2,4-dichlorophenol are mixed, about 94% of the phenol was found to be removed, similar to the amount previously reported [28], which mean that the easily removed compound (2,4-dichlorophenol) can facilitate the removal of the hard one (phenol). This phenomenon is probably due to the different phenoxy radicals produced by the enzyme. The radicals produced from easy-to-remove phenols can react with the hard-to-remove ones. Therefore, phenol can be removed more easily in presence of 2,4-dichlorophenol than individual one.

2.6. Reusability

In practical application one attractive advantage of the immobilized enzyme is that it can be easily separated from the reaction system and reused, which greatly decreases the cost of the enzyme. The efficiency of phenol removal reduced with the cycle of the immobilized enzyme as shown in Figure 6. After the first cycle of the catalytic reaction with the HRP immobilized on GO/Fe3O4, the immobilized enzyme was recovered by centrifuging in order to remove the supernatant, and the immobilized HRP was repeatedly washed with double distilled water and the activity was assayed again. After four cycles, enzyme activity dropped to 40% of its initial activity. The reusability of the immobilized enzyme is dependent on solid support. The activity decline of the immobilized HRP on GO/Fe3O4 can probably be attributed to the polymers produced during the enzymatic reaction, which may cover the enzyme and affect the reaction of the next cycle. Further investigations on improvement of reusability and application of the immobilized enzyme in a continuous-flow device are in process.

3. Experimental Section

3.1. Reagents and Materials

HRP was purchased from Tianyuan Biologic Engineering Corp. (Wuhan, China) and stored at −20 °C. FeCl3·6H2O, FeSO4·7H2O, NH3·H2O, H2O2 (30%, w/w), phenol, 2,4-DCP, EDC, PEG, NaH2PO4·2H2O, Na2HPO4·12H2O, potassium ferricyanide and 4-aminoantipine (4-AAP) were purchased from Sinopharm Chemical Reagent Co, Ltd. (Shanghai, China). All the chemical reagents were of analytical grade and used without further purification. The water used was double distilled water.

3.2. Preparation of GO/Fe3O4 Nanoparticles

Oxidized graphite was prepared from graphite powder through a modified Hummers method [29]. The GO/Fe3O4 nanoparticles were synthesized with an ultrasonic-assisted reverse co-precipitation method [30]. In a typical experiment, 0.2 g GO (water suspension), FeSO4·7H2O (0.48 g) and FeCl3·6H2O (0.466 g) were dissolved in 50 mL of distilled water. The mixed solution was added dropwise into 20 mL of distilled water in which 3 mL of ammonia water (30%) was dissolved. The reaction was carried out at 60 °C in an ultrasound clean bath operating at 25 kHz with a power of 140 W (KQ-200KDE, Kunshan, China). After 60 min for the reaction, the generated black GO/Fe3O4 nanoparticles were collected by magnetic separation, washed with water to neutral pH, and then dried under vacuum at 50 °C.

3.3. Immobilization of HRP on GO/Fe3O4

The above obtained GO/Fe3O4 (0.2 g) was dispersed in 20 mL of PBS solution (pH 6.0) containing 1 mg HRP. One milliliter of EDC (25 mg/L) was added into the mixture. The immobilization was carried out with shaking for 12 h at 25 °C. The nanoparticles were separated and washed with PBS solution and water to remove the free enzyme. The enzyme-immobilized nanoparticles were then redispersed in PBS (pH 7.0) solution and stored at 4 °C.

3.4. Characterization

UV-visible absorption spectra were recorded on a Cary 50 UV-Vis spectrophotometer (Varian, Palo Alto, CA, USA). The magnetic properties were performed using ADE 4HF vibrating sample magnetometer (Lowell, MA, USA) at 300 K. X-ray diffraction patterns were recorded on an X’Pert PRO X-ray diffractometer (PANalytical, Almelo, The Netherlands) with a Cu Ka radiation source generated at 40 kV and 30 mA.

3.5. Enzyme Activity Measurements

The colorimetric procedure was used to determine the activity of enzyme [8], in which phenol, 4-AAP and H2O2 were used as substrates. A mixture was obtained by adding 0.1 mL 0.2 mol/L phenol, PBS (pH 7.4), 0.1 mL 4.8 × 10−2 mol/L 4-AAP, 20–100 μL solution of free or immobilized HRP and 20 μL 2 × 10−3 mol/L H2O2 into a cuvette. The total volume was adjusted by water to 1.8 mL. The catalytic reaction was monitored by recording the absorbance of its red product at 510 nm. One unit of the activity (U) of HRP was defined as the amount of HRP required to hydrolyze 1 × 10−6 mol of H2O2 converted per minute under the conditions stated above.

3.6. Experiments for Removing Phenol and the Mixture of Phenol and 2,4-Dichlorophenol

The experiment for removing phenol was carried out in 100 mL conical flasks in a temperature-controlled water bath shaker (SHZ-82A) at temperature of 25 °C and a constant shaking speed of 160 rpm. The typical concentration in the reaction solution (pH 7.4) contained phenol (1 mmol/L), immobilized enzyme (90 mg/L), PBS (25 mmol/L) and PEG (the mass ratio of PEG/phenol was set at 0.4). After 15 min, to achieve of adsorption–desorption equilibrium of phenol on the GO/Fe3O4 particles, the degradation was initiated by rapid adding H2O2 (1.5 mmol/L) to the reaction solution, the total volume was rapidly adjusted to 40 mL with distilled water. At given time intervals, 2 mL aliquots of the reaction solution were sampled, and immediately centrifuged at 14,000 rpm for 3 min with an EBA-21 centrifugal (Hettich, Germany) to remove any sediment.
The experiment for removing the mixture of phenol and 2,4-dichlorophenol was carried out under almost the same conditions as described for removing the mixture of phenol. The difference was that a mixture of phenol (0.5 mmol/L) and 2,4-dichlorophenol (0.5 mmol/L) was used instead of phenol alone, and the load of PEG was 42 mg/L. Each run of the experiments was replicated at least three times.
The detection of the concentrations of phenol and 2,4-dichlorophenol were conducted on high performance liquid chromatography (U-3000 HPLC, Thermo scientific, Waltham, MA, USA) equipped with a C-18 column (5 μm,150 × 4 mm). The mobile phase was methanol/water (60/40) at a flow rate of 0.8 mL/min and detection wavelength was 270 nm for phenol and 288 nm for 2,4-dichlorophenol. Total organic carbon was measured by a TOC (Vario TOC SELECT, Elementar, Hanau, Germany).

4. Conclusions

In this study, GO/Fe3O4 nanoparticles were synthesized by an ultrasonic-assisted reverse co-precipitation method that was fast and simple, compared with others reported. HRP was successfully immobilized on the GO/Fe3O4 via interaction between functional groups of GO/Fe3O4 and HRP with EDC. We investigated the catalytic property in phenolic compound removal. Compared with soluble HRP, HRP immobilized on GO/Fe3O4 exhibited better thermal stability. PEG improved the enzymatic treatment efficiency of synthetic wastewater significantly. However, the reusability of the immobilized HRP was not satisfactory, possibly because polymers generated during reaction adsorption on the enzyme surface led to enzyme reduction during the reaction cycle. Further investigations on improvement of reusability and application of the immobilized enzyme are in process.

Acknowledgments

This work was supported by the National Science Foundation of China (Grant Nos. 21107143, 21177044 and 21377169) the Fundamental Research Funds for the Central Universities, South-Central University for Nationalities (CZY15003) and the National High Technology Research and Development Program of China (863 Program) (Grant No. 2012AA06A304).

Author Contributions

Qing Chang and Heqing Tang planed the study and wrote the manuscript. Jia Huang and Yaobin Ding carried out the experiments. All authors have given approval to the final version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, M.; Huang, Y.; Cao, H.J.; Lin, Y.H.; Li, X.; Li, H.M. Adsorption of phenol from aqueous solutions using a hydrophilic acrylic ester resin. Environ. Eng. Sci. 2015, 32, 881–890. [Google Scholar] [CrossRef]
  2. Zhou, L.C.; Meng, X.G.; Fu, J.W.; Yang, Y.C.; Yang, P.; Mi, C. Highly efficient adsorption of chlorophenols onto chemically modified chitosan. Appl. Surf. Sci. 2014, 292, 735–741. [Google Scholar] [CrossRef]
  3. Chen, Z.; Li, H.; Gao, Y.Z.; Peng, A.P. Removal of phenanthrene and acenaphthene from aqueous solution by enzyme-catalyzed phenol coupling reaction. Chem. Eng. J. 2015, 265, 27–33. [Google Scholar] [CrossRef]
  4. Nanayakkara, S.; Zhao, Z.Y.; Patti, A.F.; He, L.Z.; Saito, K. Immobilized horseradish peroxidase (I-HRP) as biocatalyst for oxidative polymerization of 2,6-dimethylphenol. ACS Sustain. Chem. Eng. 2014, 2, 1947–1950. [Google Scholar] [CrossRef]
  5. Dong, S.P.; Mao, L.; Luo, S.Q.; Zhou, L.; Feng, Y.P.; Gao, S.X. Comparison of lignin peroxidase and horseradish peroxidase for catalyzing the removal of nonylphenol from water. Environ. Sci. Pollut. Res. 2014, 21, 2358–2366. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, J.L.; Zhang, F.; Yang, H.J.; Huang, X.L.; Liu, H.; Zhang, J.Y.; Guo, S.W. Graphene oxide as a matrix for enzyme immobilization. Langmuir 2010, 26, 6083–6085. [Google Scholar] [CrossRef] [PubMed]
  7. Huang, Q.; Huang, Q.G.; Pinto, R.A.; Griebenow, K.; Stenner, R.S.; Weber, W.J. Inactivation of horseradish peroxidase by phenoxyl radical attack. J. Am. Chem. Soc. 2005, 127, 1431–1437. [Google Scholar] [CrossRef] [PubMed]
  8. Nicell, J.A.; Saadi, K.W.; Buchanan, I.D. Phenol polymerization and precipitation by horseradish peroxidase enzyme and an additive. Bioresour. Technol. 1995, 54, 5–16. [Google Scholar] [CrossRef]
  9. Rodrigues, R.C.; Ortiz, C.; Berenguer-Murcia, Á.; Torres, R.; Fernández-Lafuente, R. Modifying enzyme activity and selectivity by immobilization. Chem. Soc. Rev. 2013, 42, 6290–6307. [Google Scholar] [CrossRef] [PubMed]
  10. Garcia-Galan, C.; Berenguer-Murcia, Á.; Fernandez-Lafuente, R.; Rodrigues, R.C. Potential of different enzyme immobilization strategies to improve enzyme performance. Adv. Synth. Catal. 2011, 353, 2885–2904. [Google Scholar] [CrossRef]
  11. Mateo, C.; Palomo, J.M.; Fuentes, M.; Betancor, L.; Grazu, V.; López-Gallego, F.; Pessela, B.C.C.; Hidalgo, A.; Fernández-Lorente, G.; Fernández-Lafuente, R.; et al. Glyoxyl agarose: A fully inert and hydrophilic support for immobilization and high stabilization of proteins. Enzyme Microb. Technol. 2006, 39, 274–280. [Google Scholar] [CrossRef]
  12. Mateo, C.; Palomo, J.M.; Fernandez-Lorente, G.; Guisan, J.M.; Fernandez-Lafuente, R. Improvement of enzyme activity, stability and selectivity via immobilization techniques. Enzyme Microb. Technol. 2007, 40, 1451–1463. [Google Scholar] [CrossRef]
  13. Sheldon, R.A.; van Pelt, S. Enzyme immobilisation in biocatalysis: Why, what and how. Chem. Soc. Rev. 2013, 42, 6223–6235. [Google Scholar] [CrossRef] [PubMed]
  14. Guzik, U.; Hupert-Kocurek, K.; Wojcieszyńska, D. Immobilization as a strategy for improving enzyme properties-application to oxidoreductases. Molecules 2014, 19, 8995–9018. [Google Scholar] [CrossRef] [PubMed]
  15. Gusan, J.M.; Alvaro, G.; Fernandez-Lafuente, R. Industrial design of enzymic processes catalysed by very active immobilized derivatives: Utilization of diffusional limitations (gradients of pH) as a profitable tool in enzyme engineering. Biotechnol. Appl. Biochem. 1994, 20, 357–369. [Google Scholar]
  16. Chang, Q.; Tang, H.Q. Immobilization of horseradish peroxidase on NH2-modified magnetic Fe3O4/SiO2 particles and its application in removal of 2,4-dichlorophenol. Molecules 2014, 19, 15768–15782. [Google Scholar] [CrossRef] [PubMed]
  17. Garlet, T.B.; Weber, C.T.; Klaic, R.; Foletto, E.L.; Jahn, S.L.; Mazutti, M.A.; Kuhn, R.C. Carbon nanotubes as supports for inulinase immobilization. Molecules 2014, 19, 14615–14624. [Google Scholar] [CrossRef] [PubMed]
  18. Zucca, P.; Sanjust, E. Inorganic materials as supports for covalent enzyme immobilization: methods and mechanisms. Molecules 2014, 19, 14139–14194. [Google Scholar] [CrossRef] [PubMed]
  19. Chang, Q.; Deng, K.J.; Zhu, L.H.; Jiang, G.D.; Yu, C.; Tang, H.Q. Determination of hydrogen peroxide with the aid of peroxidase-like Fe3O4 magnetic nanoparticles as the catalyst. Microchim. Acta 2009, 165, 299–305. [Google Scholar] [CrossRef]
  20. Wang, M.Q.; Wang, N.; Tang, H.Q.; Cao, M.J.; She, Y.B.; Zhu, L.H. Surface modification of nano-Fe3O4 with EDTA and its use in H2O2 activation for removing organic pollutants. Catal. Sci. Technol. 2012, 2, 187–194. [Google Scholar] [CrossRef]
  21. Chang, Q.; Tang, H.Q. Optical determination of glucose and hydrogen peroxide using a nanocomposite prepared from glucose oxidase and magnetite nanoparticles immobilized on graphene oxide. Microchim. Acta 2014, 181, 527–534. [Google Scholar] [CrossRef]
  22. Si, Y.C.; Samulski, E.T. Exfoliated graphene separated by platinum nanoparticles. Chem. Mater. 2008, 20, 6792–6797. [Google Scholar] [CrossRef]
  23. Bai, L.Z.; Zhao, D.L.; Xu, Y.; Zhang, J.M.; Gao, Y.L.; Zhao, L.Y.; Tang, J.T. Inductive heating property of graphene oxide-Fe3O4 nanoparticles hybrid in an AC magnetic field for localized hyperthermia. Mater. Lett. 2012, 68, 399–401. [Google Scholar] [CrossRef]
  24. Yan, H.; Zhang, J.; You, C.; Song, Z.; Yu, B.; Shen, Y. Influences of different synthesis conditions on properties of Fe3O4 nanoparticles. Mater. Chem. Phys. 2009, 113, 46–52. [Google Scholar] [CrossRef]
  25. Cheng, J.; Yu, S.M.; Zuo, P. Horseradish peroxidase immobilized on aluminum-pillared interlayered clay for the catalytic oxidation of phenolic wastewater. Water Res. 2006, 40, 283–290. [Google Scholar] [CrossRef] [PubMed]
  26. Nicell, J.A.; Bewtra, J.K.; Taylor, K.E.; Biswas, N.; St Pierre, C. Enzyme catalyzed polymerization and precipitation of aromatic compounds from wastewater. Water Sci. Technol. 1992, 25, 157–164. [Google Scholar]
  27. Liu, J.Z.; Song, H.Y.; Weng, L.P.; Ji, L.N. Increased thermostability and phenol removal efficiency by chemical modified horseradish peroxidase. J. Mol. Catal. B Enzym. 2002, 18, 225–232. [Google Scholar] [CrossRef]
  28. Entezaria, M.H.; Petrier, C. A combination of ultrasound and oxidative enzyme: Sono-enzyme degradation of phenols in a mixture. Ultrason. Sonochem. 2005, 12, 283–288. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, C.M.; Yang, Q.H.; Yang, Y.G.; Lv, W.; Wen, Y.F.; Hou, P.X.; Wang, M.Z.; Cheng, H.M. Self-assembled free-standing graphite oxide membrane. Adv. Mater. 2009, 21, 3007–3011. [Google Scholar] [CrossRef]
  30. Wang, N.; Zhu, L.; Wang, D.; Wang, M.; Lin, Z.; Tang, H. Sono-assisted preparation of highly-efficient peroxidase-like Fe3O4 magnetic nanoparticles for catalytic removal of organic pollutants with H2O2. Ultrason. Sonochem. 2010, 17, 526–533. [Google Scholar] [CrossRef] [PubMed]
  • Sample Availability: Samples of the compounds Fe3O4 are available from the authors.
Figure 1. (a) Magnetization curve of GO/Fe3O4; (b) The X-ray diffraction patterns of GO, GO/Fe3O4 and GO/Fe3O4 + HRP; (c) TEM image of GO/Fe3O4 + HRP.
Figure 1. (a) Magnetization curve of GO/Fe3O4; (b) The X-ray diffraction patterns of GO, GO/Fe3O4 and GO/Fe3O4 + HRP; (c) TEM image of GO/Fe3O4 + HRP.
Molecules 21 01044 g001
Scheme 1. Schematic illustration of HRP immobilized on GO/Fe3O4 by EDC.
Scheme 1. Schematic illustration of HRP immobilized on GO/Fe3O4 by EDC.
Molecules 21 01044 sch001
Figure 2. (a) The removals of phenol in different systems; (b) Effect of mass ratio of PEG/phenol dose on phenol removal; (c) The time dependence of removal TOC; (d) Effect of H2O2 concentration on phenol removal in Fe3O4/GO system without HRP.
Figure 2. (a) The removals of phenol in different systems; (b) Effect of mass ratio of PEG/phenol dose on phenol removal; (c) The time dependence of removal TOC; (d) Effect of H2O2 concentration on phenol removal in Fe3O4/GO system without HRP.
Molecules 21 01044 g002
Figure 3. Effect of immobilized HRP dose on phenol removal.
Figure 3. Effect of immobilized HRP dose on phenol removal.
Molecules 21 01044 g003
Figure 4. Effect of H2O2 concentration on phenol removal in the absence and presence of PEG.
Figure 4. Effect of H2O2 concentration on phenol removal in the absence and presence of PEG.
Molecules 21 01044 g004
Figure 5. Time profiles of the removals of phenol and 2,4-DCP during the oxidation of: (a) phenol and 2,4-DCP alone; and (b) their mixture.
Figure 5. Time profiles of the removals of phenol and 2,4-DCP during the oxidation of: (a) phenol and 2,4-DCP alone; and (b) their mixture.
Molecules 21 01044 g005
Figure 6. Reusability of the immobilized enzyme in phenol removal.
Figure 6. Reusability of the immobilized enzyme in phenol removal.
Molecules 21 01044 g006

Share and Cite

MDPI and ACS Style

Chang, Q.; Huang, J.; Ding, Y.; Tang, H. Catalytic Oxidation of Phenol and 2,4-Dichlorophenol by Using Horseradish Peroxidase Immobilized on Graphene Oxide/Fe3O4. Molecules 2016, 21, 1044. https://doi.org/10.3390/molecules21081044

AMA Style

Chang Q, Huang J, Ding Y, Tang H. Catalytic Oxidation of Phenol and 2,4-Dichlorophenol by Using Horseradish Peroxidase Immobilized on Graphene Oxide/Fe3O4. Molecules. 2016; 21(8):1044. https://doi.org/10.3390/molecules21081044

Chicago/Turabian Style

Chang, Qing, Jia Huang, Yaobin Ding, and Heqing Tang. 2016. "Catalytic Oxidation of Phenol and 2,4-Dichlorophenol by Using Horseradish Peroxidase Immobilized on Graphene Oxide/Fe3O4" Molecules 21, no. 8: 1044. https://doi.org/10.3390/molecules21081044

Article Metrics

Back to TopTop