Next Article in Journal
Plasma-Based Synthesis and Modification of Nanomaterials
Next Article in Special Issue
Effects of Preparation Method on the Physicochemical Properties of Cationic Nanocellulose and Starch Nanocomposites
Previous Article in Journal
Removing Metal Ions from Water with Graphene–Bovine Serum Albumin Hybrid Membrane
Previous Article in Special Issue
Advancing the Use of High-Performance Graphene-Based Multimodal Polymer Nanocomposite at Scale
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nano-Cellulose/MOF Derived Carbon Doped CuO/Fe3O4 Nanocomposite as High Efficient Catalyst for Organic Pollutant Remedy

1
National-Provincial Joint Engineering Research Center of Electromechanical Product Packaging, Jiangsu Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, Nanjing Forestry University, Nanjing 210037, China
2
Department of Chemical Engineering, University of New Brunswick, Fredericton, NB E3B 5A3, Canada
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(2), 277; https://doi.org/10.3390/nano9020277
Submission received: 20 January 2019 / Revised: 7 February 2019 / Accepted: 11 February 2019 / Published: 16 February 2019

Abstract

:
Metal–organic framework (MOF)-based derivatives are attracting increased interest in various research fields. In this study, nano-cellulose MOF-derived carbon-doped CuO/Fe3O4 nanocomposites were successfully synthesized via direct calcination of magnetic Cu-BTC MOF (HKUST-1)/Fe3O4/cellulose microfibril (CMF) composites in air. The morphology, structure, and porous properties of carbon-doped CuO/Fe3O4 nanocomposites were characterized using SEM, TEM, powder X-ray diffraction (PXRD), X-ray photoelectron spectroscopy (XPS), and vibrating sample magnetometry (VSM). The results show that the as-prepared nanocomposite catalyst is composed of Fe3O4, CuO, and carbon. Compared to the CuO/Fe3O4 catalyst from HKUST-1/Fe3O4 composite and CuO from HKUST-1, this carbon-doped CuO/Fe3O4 nanocomposite catalyst shows better catalytic efficiency in reduction reactions of 4-nitrophenol (4-NP), methylene blue (MB), and methyl orange (MO) in the presence of NaBH4. The enhanced catalytic performance of carbon-doped CuO/Fe3O4 is attributed to effects of carbon preventing the aggregation of CuO/Fe3O4 and providing high surface-to-volume ratio and chemical stability. Moreover, this nanocomposite catalyst is readily recoverable using an external magnet due to its superparamagnetic behavior. The recyclability/reuse of carbon-doped CuO/Fe3O4 was also investigated.

1. Introduction

Recently, metal nanoparticles (NPs) were widely used in the fields of biomedicine and chemical reactions due to their high selectivity and catalytic efficiency [1,2,3]. Noble-metal nanoparticles (gold, silver, etc.) [4,5,6] and non-noble-metal nanoparticles (copper, zinc, and their oxides, sulfides, etc.) [7,8,9,10] are particularly noticeable. For example, Jiang et al. [11] reported that CuO and Au domains could greatly improve the photocatalytic activity and stability of Cu2O cubes in the photocatalytic degradation of methyl orange (MO). Rodríguez et al. [12] reported that potassium poly(heptazine imide) (PHIK)/Ti-based metal–organic framework (MIL-125-NH2) composites had superior photocatalytic activity in rhodamine B (RhB) degradation under blue-light irradiation. Among the applications, metal nanoparticles can also be used for treating wastewater and drinking water due to their large surface areas and high activity [13,14]. With growing focus on the development of cost-effective, efficient catalysts, more attention is being paid to non-noble-metal catalysts, such as metal-oxide catalysts.
Metal NPs with nanometer-scale dimensions are unstable and tend to aggregate due to their high surface energy and surface area, which can lead to the loss of catalytic efficiency [15,16,17]. The concept of immobilization/stabilization of metal NPs onto support/substrates is one of the effective methods to overcome aggregation of NPs [18,19]. There are many types of substrates/matrices that can be used to support metal NPs, such as carbon [20,21], silica [22], metal oxide [23], polymers, etc. [24,25,26]. Carbon, as a support for metal nanoparticles, provides multiple accessible channels for diffusion/transport to take advantage of the excellent catalytic functionalities of metal nanoparticles [27].
In past decades, metal–organic frameworks (MOFs) attracted much attention due to their porous structures and potential applications in gas storage, molecule separation, chemical sensing, drug delivery, and catalysis [28,29,30]. Recently, MOF-based derivative catalysts received more attention because MOF-derived materials have advantages in terms of tailorable morphologies, hierarchical porosity, and easy functionalization with other heteroatoms and metal oxides [31,32]. For instance, Ji et al. used ZIF-67 as a precursor to synthesize a Pt@Co3O4 composite to improve the catalytic activity of CO oxidation [33]. Yang et al. reported that ZnO nanoparticles prepared via calcination of MOF-5 in air at 600 °C showed excellent photocatalytic degradation of rhodamine B [34]. Niu et al. synthesized a hybrid catalyst consisting of Cu/Cu2O NPs supported on porous carbon for the catalytic reduction of 4-nitrophenol (4-NP) using HKUST-1 as a precursor [27].
The direct pyrolysis/thermolysis treatment of MOFs is a simple and controllable method to prepare various metal oxides in a one-step process. By following it, we can successfully synthesize carbon-doped CuO/Fe3O4 composite catalysts for organic pollutant reduction. Herein, we prepared carbon-doped CuO/Fe3O4 composite catalysts via direct calcination of HKUST-1/Fe3O4/CMF composites under air. We then applied the as-prepared carbon-doped CuO/Fe3O4 composite catalysts for the catalytic reduction of 4-NP. Its catalytic performance, in comparison with a CuO/Fe3O4 composite from HKUST-1/Fe3O4 composite and CuO, is remarkably better, which is attributed to the fact that carbon doping can (1) minimize the aggregation of CuO/Fe3O4, (2) provide high surface-to-volume ratio and chemical stability for the catalyst in contact with the target pollutants, and (3) enhance the catalytic activity of the CuO/Fe3O4 catalyst. Furthermore, the carbon-doped CuO/Fe3O4 composite catalyst has excellent efficiency in the reduction of methylene blue (MB) and methyl orange (MO). The features of the carbon-doped CuO/Fe3O4 composite catalyst are as follows: CuO acts as the effective catalyst, with the doped carbon having the three functions discussed, and the magnetic Fe3O4 supports easy reuse/recycling of the catalyst using a magnet.

2. Experimental

2.1. Materials

Cellulose microfibrils (CMF) (~4.0 wt.%) from Cellulose Lab; trimesic acid (H3BTC), copper (II) acetate monohydrate, ethanol, and hydrochloric acid (HCl, 37%, w/w) from Sigma-Aldrich; methylene blue, ferric nitrate nonahydrate (Fe(NO3)3·9H2O), ferrous sulfate (FeSO4·7H2O), citric acid anhydrous, and sodium hydroxide (NaOH, 50%, w/w) from Fisher Scientific were used in this study. All other chemicals were of analytical grade and used without further purification.

2.2. Synthesis of Magnetic Carbon-Doped CuO/Fe3O4 Nanocomposite Catalyst

Fe3O4/cellulose microfibril (CMF) composites were prepared using chemical co-precipitation of aqueous ferrous and ferric ions, and the procedures were similar to those described in the literature [35]. CMF (7.5 g, ~4%, w/w) was diluted to 0.3 wt.% with distilled water and stirred for 5 min at 1000 rpm. The diluted CMF suspension was further treated in a sonicator (QSON-ICA) for 5 min to disperse the individual nanocellulose, followed by heating at 65 °C and bubbling with nitrogen for 10 min under magnetic stirring (500 rpm) to remove the dissolved oxygen from the CMF suspension. Then, citric acid (1 mg) and diluted HCl solution (1 mL, 2 mol/L) were added to the above solution, followed by the successive additions of 1.3 mmol Fe(NO3)3·9H2O and 0.65 mmol FeSO4·7H2O solid samples with nitrogen bubbling and magnetic stirring. The function of citric acid was to protect the as-prepared Fe3O4 NPs from being oxidized by the dissolving oxygen in water. A dilute NaOH solution (10 mL, 2 mol/L) was added drop-wise into the above system, followed by bubbling with nitrogen and magnetic stirring (500 rpm) for 120 min at room temperature. The obtained Fe3O4/CMF composites (using a magnet) were washed three times with distilled water in a centrifugation step (4000 rpm, 10 min).
Copper (II) acetate monohydrate was dissolved in distilled water. Fe3O4/CMF nanocomposites and H3BTC were dispersed in ethanol and then treated in a sonicator (QSON-ICA) for 5 min. The ethanol solution containing Fe3O4/CMF composite and H3BTC was added to the aqueous system. The color of the solution changed from light blue to blue-black immediately, and the system was stirred continuously for 4 h. The final HKUST-1/Fe3O4/CMF composite was washed in a centrifugation step (4000 rpm, 10 min) with distilled water and absolute ethanol three times. Then, the drying was done at 50 °C under vacuum.
The synthesized HKUST-1/Fe3O4/CMF composites were placed in a ceramic boat, then transferred to a muffle roaster, which was operated under the conditions of 500 °C, air atmosphere, and a heating rate of 20 °C/min, before being held at 500 °C for 5 h; the procedures were similar to those described in the literature [36]. After cooling to room temperature, the sample was washed several times with distilled water and absolute ethanol in a centrifugation step (4000 rpm, 10 min), and finally dried in a vacuum at 60 °C for 5 h.

2.3. Characterization

Transmission electron microscope (TEM) analyses were conducted with a JEOL JEM 2011 transmission electron microscope at 200 kV. Scanning electron microscope (SEM) analyses were conducted with a JEOL JSM 6400 scanning electron microscope. The powder X-ray diffraction (PXRD) patterns of the prepared samples were collected using an X-ray diffractometer with Cu-Kα radiation (40 kV, 30 mA). The patterns were recorded in the region of 2θ from 10° to 80° with a scan step of 0.02°. The chemical binding energies of the respective ions in the samples were measured using X-ray photoelectron spectroscopy (XPS, ESCALa-b220i-XL electron spectrometer, Thermo Fisher Scientific K-Alpha, UK) under 300-W Al-Kα radiation. The magnetic properties were measured with a vibrating sample magnetometer (VSM), a physical property measurement system, at 300 K, as a function of the applied magnetic field between −80 and 80 kOe. The ultraviolet–visible (UV–Vis) diffuse reflectance data were collected with a UV–Vis spectrophotometer (Evolution 201, Thermo Scientific) equipped with an integrated sphere.

2.4. Evaluation of Catalytic Performance

The catalytic reduction of 4-nitrophenol (4-NP) using NaBH4 was chosen as the model reaction for investigating the catalytic performance of magnetic carbon-doped CuO/Fe3O4 nanocomposite catalysts. Typically, catalytic reduction of a 30-mL 4-NP solution (1 mmol/L) was carried out in a beaker (100 mL) by continuously stirring at room temperature. Upon the addition of NaBH4 (10 mM) into the 4-NP solution, its color changed immediately from light yellow to dark yellow due to the formation of 4-nitrophenolate ions (formed from the high alkalinity of NaBH4) [37]. Then, the dark-yellow color faded with time (due to the conversion of 4-NP to 4-aminophenol (4-AP)) after the addition of carbon-doped CuO/Fe3O4 nanocomposite catalyst (5 mg) (Scheme 1). UV–Vis adsorption spectra were recorded using a UV–Vis spectrophotometer in the range of 250–550 nm. When the reaction completed, the catalyst was easily separated from the solution using an external magnet due to its good magnetic performance.
During the degradation process of organic dyes, 30 mL of dye solution (10 mg/L) and NaBH4 (10 mM) were mixed in a beaker (100 mL) by continuously stirring at room temperature. Then, 5 mg of carbon-doped CuO/Fe3O4 nanocomposite catalyst was added to the reaction system, and the catalytic process was monitored using the UV–Vis spectrophotometer. When the reaction ended, the catalyst was readily separated from the solution using an external magnet.

3. Results and Discussion

3.1. Structure and Morphological Characterization

MOFs can be readily converted to metal-oxide composites, which take advantage of their original morphology and porosity. The as-prepared HKUST-1, HKUST-1/Fe3O4, and HKUST-1/Fe3O4/CMF composite samples were then calcinated to obtain the nanoporous metal-oxide particles, generating nanoporous CuO, CuO/Fe3O4, and carbon-doped CuO/Fe3O4 composite catalysts, respectively. Pyrolysis/thermolysis of HKUST-1/Fe3O4/CMF composites led to the formation of porous carbon-doped CuO/Fe3O4 composites.
The size and the morphology of the synthesized materials were investigated using a transmission electron microscope (TEM) and scanning electron microscope (SEM). Figure 1 shows the SEM and TEM images of HKUST-1/Fe3O4/CMF composites and carbon-doped CuO/Fe3O4 composites after calcination in air. HKUST-1 crystals and Fe3O4 nanoparticles were uniformly loaded onto the surface of CMF (Figure 1). The nanocellulose MOF-derived porous carbon-doped CuO/Fe3O4 composite was then obtained through a calcination process under air at 500 °C for 5 h. The Brunauer–Emmett–Teller surface area (SBET) of the carbon-doped CuO/Fe3O4 sample was 38.7 m2/g, which was much higher than that of the CuO/Fe3O4 sample (0.042 m2/g) obtained from the calcination of HKUST-1/Fe3O4. The resultant carbon-doped CuO/Fe3O4 composite inherited largely the original porous morphology of HKUST-1. Based on the mapping images, the conclusion was that both Cu and Fe elements dispersed well in the carbon-doped CuO/Fe3O4 composite catalyst.
XPS analysis was employed to investigate the elemental composition on the surface of different composites. For the XPS spectrum of carbon-doped CuO/Fe3O4 composite catalyst (Figure 2a), the main peaks were C1s, O1s, Fe2p, and Cu2p, centered at around 285 eV, 530 eV, 720 eV, and 930 eV, respectively. The C1s XPS pattern of the sample is shown in Figure 2b. The spectrum contains two peaks at 285 eV and 288.5 eV, which may be attributed to carbon (sp2 hybridization) in the sample, and the Cu–O–C bonds or Fe–O–C bonds, respectively. Figure 2c shows the Cu2p core-level XPS spectrum of the composite catalyst. The Cu2p1 and Cu2p3 binding energies for the composite catalyst were 952.8 and 932.7 eV, respectively, indicating the presence of CuO in the composite catalyst. Similar results were reported in the literature [38,39]. The Fe2p3 and Fe2p1 binding energies (Figure 2d) for the composite catalyst were 710.7 and 725.4 eV, respectively, which agrees with published results [35,40], confirming the presence of Fe3O4 in the composite catalyst.
The crystalline nature and the composition of the as-synthesized products were characterized using PXRD. The crystalline phases of CuO, such as (110), (11-1), (111), (20-2), (020), (202), (11-3), (31-1), and (220) are shown in Figure 3a, which are consistent with those reported in the literature [41,42]. These results support the conclusion that HKUST-1 was transformed into CuO via calcination under the present conditions. In the XRD pattern of carbon-doped CuO/Fe3O4 composite, the diffraction peaks and relative intensities were indexed to Fe3O4 (JCPD NO. 19-0629) and CuO (JCPD NO. 05-0661). This indicates that CuO and Fe3O4 were both obtained via calcining HKUST-1/Fe3O4/CMF composites under air. No other C-related impurity peaks were detected; a similar result was reported in the previous research [43].
The magnetic behavior of the carbon-doped Fe3O4/CuO composite catalyst sample was evaluated at 300 K. Its magnetization saturation value (Ms) was 15.1 emu∙g−1, suggesting a good magnetic property (Figure 3b). Thus, the magnetic carbon-doped Fe3O4/CuO composite catalyst could be readily separated from the reaction system using a magnet for the recycling process.

3.2. Catalytic Reduction

In this study, we investigated the application of carbon-doped CuO/Fe3O4 composite catalysts for the catalytic reduction of 4-nitrophenol. In Figure 4a, the results are shown for the catalytic reduction of 4-NP to 4-AP in the presence of NaBH4 and carbon-doped CuO/Fe3O4 composite catalysts. The reduction process was followed based on the UV–Vis spectrophotometry. It shows that the absorbance at 400 nm (4-NP) decreased gradually as a function of time, while the absorbance at 290 nm (due to 4-AP) increased, confirming the catalytic reduction of 4-NP to 4-AP [44]. The catalytic reduction was almost complete within 10 min at room temperature, and the color of the solution changed from yellow to colorless. Similar results were also reported in the literature; Bordbar et al. [45] found that, using CuO/ZnO nanocomposites, catalytic reduction of 4-NP to 4-AP (using NaBH4 as the reducing agent) was completed in several minutes.
The kinetics of calcined CMF, CuO, CuO/Fe3O4, and carbon-doped CuO/Fe3O4 composites are shown in Figure 4b. In the absence of catalyst, the reduction of 4-NP by NaBH4 was negligible. In each case, the pseudo-first-order kinetic prevailed. The calcined CMF also had a negligible effect on the reduction of 4-NP. The rate constants (k) for CuO, CuO/Fe3O4, and carbon-doped CuO/Fe3O4 composite samples were 1.3 × 10−3 s−1, 3.6 × 10−3 s−1 and 6.5 × 10−3 s−1, respectively. In the case of CuO, the catalytic activity was the lowest due to the aggregation of CuO nanoparticles, while the catalytic efficiency of carbon-doped CuO/Fe3O4 composites was much better than that of CuO/Fe3O4, demonstrating that carbon doping is effective for enhancing the catalytic activity of the catalysts.
The catalytic reduction of 4-NP by NaBH4 using metal-oxide nanoparticles (CuO) has two steps [46]: (1) borohydride ions are adsorbed onto the nanoparticle surface, forming active surface-hydrogen, while 4-NP is also adsorbed onto the nanoparticle surface; (2) active hydrogen attacks the positively charged nitrogen in the nitro group of 4-NP, followed by the addition of two hydrogen atoms, producing 4-AP.
Cationic and anionic organic dyes were chosen to further investigate the catalytic properties of carbon-doped CuO/Fe3O4 composite catalyst. As shown in Figure 5a, for cationic dye (methylene blue), in the presence of NaBH4 and carbon-doped CuO/Fe3O4 composite catalyst, the absorbance at 660 nm (MB) gradually decreased as a function of time; furthermore, the catalytic reduction was completed within 6 min at room temperature (the color of the solution was colorless).
The results from methyl orange (an anionic dye) are shown in Figure 5c. Under otherwise the same conditions, the color change (from orange to colorless) was slower than that for MB (Figure 5a, from blue to colorless). The pseudo-first-order rate law was also valid here (Figure 5d). For the carbon-doped CuO/Fe3O4 composite catalyst, the rate constant (k) was 2.4 × 10−3 s−1 for MO, while it was 12.9 × 10−3 s−1 for MB.
We compared the catalytic performance of the carbon-doped CuO/Fe3O4 composite catalyst for the reduction of 4-NP, MB, and MO, with other related ones from the literature (Table 1). As shown, the carbon-doped CuO/Fe3O4 composite catalyst showed much improved results, and the pseudo-first-order rate constant (k) for the nanocomposite catalyst from the present study was indeed consistently higher than that reported in the literature. The improved results may be attributed to the unique original morphologies associated with HKUST-1.
From the viewpoint of practical application, the recycling/reuse of the catalyst is of critical importance. In the present study, after the catalytic degradation experiments, the magnetic carbon-doped CuO/Fe3O4 composite catalyst was readily separated from the reaction system using an external magnet. The used catalysts were collected, and rinsed with distilled water several times. After a thorough washing process, the recovered magnetic catalyst was reused in the subsequent run of catalytic reduction of 4-NP under identical conditions, and the same process was repeated five times. The results are shown in Figure 6. The catalytic performance of the magnetic carbon-doped CuO/Fe3O4 composite catalyst decreased only slightly (the 4-NP reduction ratio decreased from 100% to 96%) after five cycles. Similar results were obtained in the reuse/recycling experiments of the as-prepared magnetic carbon-doped CuO/Fe3O4 composite catalyst during the catalytic reduction of MB and MO. Therefore, the as-prepared magnetic carbon-doped CuO/Fe3O4 composite catalyst is a promising system for practical applications.

4. Conclusions

In this study, a nano-cellulose/MOF-derived carbon-doped CuO/Fe3O4 composite catalyst was successfully fabricated through pyrolysis/thermolysis of the HKUST-1/Fe3O4/CMF composite. The resultant carbon-doped CuO/Fe3O4 composite catalyst took advantage of the original porous morphology of HKUST-1; consequently, the carbon-doped CuO/Fe3O4 composite catalyst exhibited high catalytic activity for the reduction of 4-NP and organic dyes (MB and MO). In addition, the carbon-doped CuO/Fe3O4 composite catalyst showed good reusability/recyclability after five cycles. Notably, this strategy can be extended to the preparation of other functional MOF-based derivatives.

Author Contributions

Conceptualization, H.L. and L.Z.; Methodology, H.L.; Software, H.L. and N.A.; Validation, N.A. and J.M.; Formal Analysis, H.L. and L.Z.; Investigation, H.L.; Resources, X.Z. and Y.N.; Data Curation, J.M.; Writing-Original Draft Preparation, H.L.; Writing-Review & Editing, X.Z. and Y.N.; Visualization, L.Z. and J.M.; Supervision, Y.N.; Project Administration, X.Z. and Y.N.; Funding Acquisition, X.Z. and Y.N.

Funding

This research was funded by the National Natural Science Foundation of China (31570576), the financial support from the Natural Science Foundation of Jiangsu Provincial University (16KJA220005), the Canada Research Chairs program of the Government of Canada, the Postgraduate Research and Practice Innovation Program of Jiangsu Province (KYCX17_0835), the Doctorate Fellowship Foundation of Nanjing Forestry University, and the Priority Academic Program Development (PAPD) of Jiangsu Higher Education Institutions. The APC was funded by the Natural Science Foundation of Jiangsu Provincial University (16KJA220005).

Acknowledgments

The authors acknowledge Mike Johnson (Department of Chemistry and Institute for Research in Materials, Dalhousie University, Halifax, NS, Canada) for the acquisition and interpretation of the PPMS data.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Almasi, H.; Jafarzadeh, P.; Mehryar, L. Fabrication of novel nanohybrids by impregnation of CuO nanoparticles into bacterial cellulose and chitosan nanofibers: Characterization, antimicrobial and release properties. Carbohydr. Polym. 2018, 186, 273–281. [Google Scholar] [CrossRef] [PubMed]
  2. Palomo, J.M.; Filice, M. Biosynthesis of Metal Nanoparticles: Novel Efficient Heterogeneous Nanocatalysts. Nanomaterials 2016, 6, 84. [Google Scholar] [CrossRef] [PubMed]
  3. Tran, C.D.; Makuvaza, J.; Munson, E.; Bennett, B. Biocompatible Copper Oxide Nanoparticle Composites from Cellulose and Chitosan: Facile Synthesis, Unique Structure, and Antimicrobial Activity. ACS Appl. Mater. Interfaces 2017, 9, 42503–42515. [Google Scholar] [CrossRef] [PubMed]
  4. Müller, A.; Peglow, S.; Karnahl, M.; Kruth, A.; Junge, H.; Brüser, V.; Scheu, C. Morphology, optical properties and photocatalytic activity of photo-and plasma-deposited au and Au/Ag core/shell nanoparticles on titania layers. Nanomaterials 2018, 8, 502. [Google Scholar] [CrossRef] [PubMed]
  5. Corma, A.; Garcia, H. Supported gold nanoparticles as catalysts for organic reactions. Chem. Soc. Rev. 2008, 37, 2096–2126. [Google Scholar] [CrossRef] [PubMed]
  6. Gangula, A.; Podila, R.; Karanam, L.; Janardhana, C.; Rao, A.M. Catalytic Reduction of 4-Nitrophenol using Biogenic Gold and Silver Nanoparticles Derived from Breynia rhamnoides. Langmuir 2011, 27, 15268–15274. [Google Scholar] [CrossRef] [PubMed]
  7. Huang, Z.; Cui, F.; Kang, H.; Chen, J.; Zhang, X.; Xia, C. Highly Dispersed Silica-Supported Copper Nanoparticles Prepared by Precipitation−Gel Method: A Simple but Efficient and Stable Catalyst for Glycerol Hydrogenolysis. Chem. Mater. 2008, 20, 5090–5099. [Google Scholar] [CrossRef]
  8. Ma, J.; Sun, Z.; Wang, Z.; Zhou, X. Preparation of ZnO–cellulose nanocomposites by different cellulose solution systems with a colloid mill. Cellulose 2016, 23, 3703–3715. [Google Scholar] [CrossRef]
  9. Aguirre, M.E.; Zhou, R.; Eugene, A.J.; Guzman, M.I.; Grela, M.A. Cu2O/TiO2 heterostructures for CO2 reduction through a direct Z-scheme: Protecting Cu2O from photocorrosion. Appl. Catal. B Environ. 2017, 217, 485–493. [Google Scholar] [CrossRef]
  10. Zhou, R.; Guzman, M.I. Photocatalytic Reduction of Fumarate to Succinate on ZnS Mineral Surfaces. J. Phys. Chem. C 2016, 120, 7349–7357. [Google Scholar] [CrossRef]
  11. Jiang, D.; Zhang, Y.; Li, X. Synergistic effects of CuO and Au nanodomains on Cu2O cubes for improving photocatalytic activity and stability. Chin. J. Catal. 2019, 40, 105–113. [Google Scholar] [CrossRef]
  12. Rodríguez, N.A.; Savateev, A.; Grela, M.A.; Dontsova, D. Facile Synthesis of Potassium Poly(heptazine imide) (PHIK)/Ti-Based Metal–Organic Framework (MIL-125-NH2) Composites for Photocatalytic Applications. ACS Appl. Mater. Interfaces 2017, 9, 22941–22949. [Google Scholar] [CrossRef] [PubMed]
  13. Hu, J.; Chen, G.; Lo, I.M.C. Removal and recovery of Cr(VI) from wastewater by maghemite nanoparticles. Water Res. 2005, 39, 4528–4536. [Google Scholar] [CrossRef] [PubMed]
  14. Simeonidis, K.; Mourdikoudis, S.; Kaprara, E.; Mitrakas, M.; Polavarapu, L. Inorganic engineered nanoparticles in drinking water treatment: A critical review. Environ. Sci. Water Res. Technol. 2016, 2, 43–70. [Google Scholar] [CrossRef]
  15. Keller, A.A.; Wang, H.; Zhou, D.; Lenihan, H.S.; Cherr, G.; Cardinale, B.J.; Miller, R.; Ji, Z. Stability and Aggregation of Metal Oxide Nanoparticles in Natural Aqueous Matrices. Environ. Sci. Technol. 2010, 44, 1962–1967. [Google Scholar] [CrossRef] [PubMed]
  16. Marambio-Jones, C.; Hoek, E.M.V. A review of the antibacterial effects of silver nanomaterials and potential implications for human health and the environment. J. Nanopart. Res. 2010, 12, 1531–1551. [Google Scholar] [CrossRef]
  17. Su, X.; Kanjanawarut, R. Control of Metal Nanoparticles Aggregation and Dispersion by PNA and PNA−DNA Complexes, and Its Application for Colorimetric DNA Detection. ACS Nano 2009, 3, 2751–2759. [Google Scholar] [CrossRef]
  18. Zhang, Q.; Lee, I.; Ge, J.; Zaera, F.; Yin, Y. Surface-Protected Etching of Mesoporous Oxide Shells for the Stabilization of Metal Nanocatalysts. Adv. Funct. Mater. 2010, 20, 2201–2214. [Google Scholar] [CrossRef]
  19. Zhu, Q.-L.; Li, J.; Xu, Q. Immobilizing Metal Nanoparticles to Metal–Organic Frameworks with Size and Location Control for Optimizing Catalytic Performance. J. Am. Chem. Soc. 2013, 135, 10210–10213. [Google Scholar] [CrossRef]
  20. Bak, W.; Soo Kim, H.; Chun, H.; Cheol Yoo, W. Facile synthesis of metal/metal oxide nanoparticles inside a nanoporous carbon matrix (M/MO@C) through the morphology-preserved transformation of metal–organic framework. Chem. Commun. 2015, 51, 7238–7241. [Google Scholar] [CrossRef]
  21. Xia, W.; Zou, R.; An, L.; Xia, D.; Guo, S. A metal–organic framework route to in situ encapsulation of Co@Co3O4@C core@bishell nanoparticles into a highly ordered porous carbon matrix for oxygen reduction. Energy Environ. Sci. 2015, 8, 568–576. [Google Scholar] [CrossRef]
  22. Ennas, G.; Falqui, A.; Marras, S.; Sangregorio, C.; Marongiu, G. Influence of Metal Content on Size, Dispersion, and Magnetic Properties of Iron−Cobalt Alloy Nanoparticles Embedded in Silica Matrix. Chem. Mater. 2004, 16, 5659–5663. [Google Scholar] [CrossRef]
  23. Obena, R.P.; Lin, P.C.; Lu, Y.W.; Li, I.C.; del Mundo, F.; Arco, S.D.; Nuesca, G.M.; Lin, C.-C.; Chen, Y.-J. Iron Oxide Nanomatrix Facilitating Metal Ionization in Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry. Anal. Chem. 2011, 83, 9337–9343. [Google Scholar] [CrossRef] [PubMed]
  24. Kaushik, M.; Moores, A. Review: Nanocelluloses as versatile supports for metal nanoparticles and their applications in catalysis. Green Chem. 2016, 18, 622–637. [Google Scholar] [CrossRef]
  25. Budarin, V.L.; Clark, J.H.; Luque, R.; Macquarrie, D.J.; White, R.J. Palladium nanoparticles on polysaccharide-derived mesoporous materials and their catalytic performance in C–C coupling reactions. Green Chem. 2008, 10, 382–387. [Google Scholar] [CrossRef]
  26. Zhang, Z.; Wang, Z. Diatomite-Supported Pd Nanoparticles: An Efficient Catalyst for Heck and Suzuki Reactions. J. Org. Chem. 2006, 71, 7485–7487. [Google Scholar] [CrossRef]
  27. Niu, H.; Liu, S.; Cai, Y.; Wu, F.; Zhao, X. MOF derived porous carbon supported Cu/Cu2O composite as high performance non-noble catalyst. Microporous Mesoporous Mater. 2016, 219, 48–53. [Google Scholar] [CrossRef]
  28. Burtch, N.C.; Jasuja, H.; Walton, K.S. Water Stability and Adsorption in Metal–Organic Frameworks. Chem. Rev. 2014, 114, 10575–10612. [Google Scholar] [CrossRef]
  29. Duan, C.; Meng, J.; Wang, X.; Meng, X.; Sun, X.; Xu, Y.; Zhao, W.; Ni, Y. Synthesis of novel cellulose- based antibacterial composites of Ag nanoparticles@ metal-organic frameworks@ carboxymethylated fibers. Carbohydr. Polym. 2018, 193, 82–88. [Google Scholar] [CrossRef]
  30. Leo, P.; Orcajo, G.; Briones, D.; Calleja, G.; Sánchez-Sánchez, M.; Martínez, F. A recyclable Cu-MOF-74 catalyst for the ligand-free O-arylation reaction of 4-nitrobenzaldehyde and phenol. Nanomaterials 2017, 7, 149. [Google Scholar] [CrossRef]
  31. Shen, K.; Chen, X.; Chen, J.; Li, Y. Development of MOF-Derived Carbon-Based Nanomaterials for Efficient Catalysis. ACS Catal. 2016, 6, 5887–5903. [Google Scholar] [CrossRef]
  32. Li, H.; Ke, F.; Zhu, J. MOF-derived ultrathin cobalt phosphide nanosheets as efficient bifunctional hydrogen evolution reaction and oxygen evolution reaction electrocatalysts. Nanomaterials 2018, 8, 89. [Google Scholar] [CrossRef] [PubMed]
  33. Ji, W.; Xu, Z.; Liu, P.; Zhang, S.; Zhou, W.; Li, H.; Zhang, T.; Li, L.; Lu, X.; Wu, J.; et al. Metal–Organic Framework Derivatives for Improving the Catalytic Activity of the CO Oxidation Reaction. ACS Appl. Mater. Interfaces 2017, 9, 15394–15398. [Google Scholar] [CrossRef] [PubMed]
  34. Yang, S.J.; Im, J.H.; Kim, T.; Lee, K.; Park, C.R. MOF-derived ZnO and ZnO@C composites with high photocatalytic activity and adsorption capacity. J. Hazard. Mater. 2011, 186, 376–382. [Google Scholar] [CrossRef] [PubMed]
  35. An, X.; Cheng, D.; Dai, L.; Wang, B.; Ocampo, H.J.; Nasrallah, J.; Jia, X.; Zou, J.; Long, Y.; Ni, Y. Synthesis of nano-fibrillated cellulose/magnetite/titanium dioxide (NFC@Fe3O4@TNP) nanocomposites and their application in the photocatalytic hydrogen generation. Appl. Catal. B Environ. 2017, 206, 53–64. [Google Scholar] [CrossRef]
  36. Zhang, Y.-F.; Qiu, L.-G.; Yuan, Y.-P.; Zhu, Y.-J.; Jiang, X.; Xiao, J.-D. Magnetic Fe3O4@C/Cu and Fe3O4@CuO core–shell composites constructed from MOF-based materials and their photocatalytic properties under visible light. Appl. Catal. B Environ. 2014, 144, 863–869. [Google Scholar] [CrossRef]
  37. An, X.; Long, Y.; Ni, Y. Cellulose nanocrystal/hexadecyltrimethylammonium bromide/silver nanoparticle composite as a catalyst for reduction of 4-nitrophenol. Carbohydr. Polym. 2017, 156, 253–258. [Google Scholar] [CrossRef] [PubMed]
  38. Boruban, C.; Esenturk, E.N. Activated carbon-supported CuO nanoparticles: A hybrid material for carbon dioxide adsorption. J. Nanopart. Res. 2018, 20, 59. [Google Scholar] [CrossRef]
  39. Gu, C.; Qi, R.; Wei, Y.; Zhang, X. Preparation and performances of nanorod-like inverse CeO2–CuO catalysts derived from Ce-1,3,5-Benzene tricarboxylic acid for CO preferential oxidation. React. Kinet. Mech. Catal. 2018, 124, 651–667. [Google Scholar] [CrossRef]
  40. Dong, X.-L.; Mou, X.-Y.; Ma, H.-C.; Zhang, X.-X.; Zhang, X.-F.; Sun, W.-J.; Ma, C.; Xue, M. Preparation of CdS–TiO2/Fe3O4 photocatalyst and its photocatalytic properties. J. Sol-Gel Sci. Technol. 2013, 66, 231–237. [Google Scholar] [CrossRef]
  41. Avgouropoulos, G.; Ioannides, T.; Matralis, H. Influence of the preparation method on the performance of CuO–CeO2 catalysts for the selective oxidation of CO. Appl. Catal. B Environ. 2005, 56, 87–93. [Google Scholar] [CrossRef]
  42. Kumar, R.V.; Diamant, Y.; Gedanken, A. Sonochemical Synthesis and Characterization of Nanometer-Size Transition Metal Oxides from Metal Acetates. Chem. Mater. 2000, 12, 2301–2305. [Google Scholar] [CrossRef]
  43. Zhang, H.; Zheng, L.; Ouyang, X.; Ni, Y. Carbon doping of Ti0.91Co0.03La0.06O2 nanoparticles for enhancing room-temperature ferromagnetism using carboxymethyl cellulose as carbon source. Ceram. Int. 2018, 44, 15754–15763. [Google Scholar] [CrossRef]
  44. Li, J.; Liu, C.; Liu, Y. Au/graphene hydrogel: Synthesis, characterization and its use for catalytic reduction of 4-nitrophenol. J. Mater. Chem. 2012, 22, 8426–8430. [Google Scholar] [CrossRef]
  45. Bordbar, M.; Negahdar, N.; Nasrollahzadeh, M. Melissa Officinalis L. leaf extract assisted green synthesis of CuO/ZnO nanocomposite for the reduction of 4-nitrophenol and Rhodamine B. Sep. Purif. Technol. 2018, 191, 295–300. [Google Scholar] [CrossRef]
  46. You, J.-G.; Shanmugam, C.; Liu, Y.-W.; Yu, C.-J.; Tseng, W.-L. Boosting catalytic activity of metal nanoparticles for 4-nitrophenol reduction: Modification of metal naoparticles with poly(diallyldimethylammonium chloride). J. Hazard. Mater. 2017, 324, 420–427. [Google Scholar] [CrossRef] [PubMed]
  47. Nabil, B.; Ahmida, E.A.; Christine, C.; Julien, V.; Abdelkrim, A. Polyfunctional cotton fabrics with catalytic activity and antibacterial capacity. Chem. Eng. J. 2018, 351, 328–339. [Google Scholar] [CrossRef]
  48. Bouazizi, N.; El achari, A.; Campagne, C.; Vieillard, J.; Azzouz, A. Copper oxide coated polyester fabrics with enhanced catalytic properties towards the reduction of 4-nitrophenol. J. Mater. Sci. Mater. Electron. 2018, 29, 10802–10813. [Google Scholar] [CrossRef]
  49. Islam, M.T.; Dominguez, N.; Ahsan, M.A.; Dominguez-Cisneros, H.; Zuniga, P.; Alvarez, P.J.J.; Noveron, J.C. Sodium rhodizonate induced formation of gold nanoparticles supported on cellulose fibers for catalytic reduction of 4-nitrophenol and organic dyes. J. Environ. Chem. Eng. 2017, 5, 4185–4193. [Google Scholar] [CrossRef]
  50. Gan, Z.; Zhao, A.; Zhang, M.; Tao, W.; Guo, H.; Gao, Q.; Mao, R.; Liu, E. Controlled synthesis of Au-loaded Fe3O4@C composite microspheres with superior SERS detection and catalytic degradation abilities for organic dyes. Dalton Trans. 2013, 42, 8597–8605. [Google Scholar] [CrossRef]
  51. Afshar, S.; Samari Jahromi, H.; Jafari, N.; Ahmadi, Z.; Hakamizadeh, M. Degradation of malachite green oxalate by UV and visible lights irradiation using Pt/TiO2/SiO2 nanophotocatalyst. Sci. Iran. 2011, 18, 772–779. [Google Scholar] [CrossRef]
  52. Ganapuram, B.R.; Alle, M.; Dadigala, R.; Dasari, A.; Maragoni, V.; Guttena, V. Catalytic reduction of methylene blue and Congo red dyes using green synthesized gold nanoparticles capped by salmalia malabarica gum. Int. Nano Lett. 2015, 5, 215–222. [Google Scholar] [CrossRef] [Green Version]
  53. Sallam, S.A.; El-Subruiti, G.M.; Eltaweil, A.S. Facile Synthesis of Ag–γ-Fe2O3 Superior Nanocomposite for Catalytic Reduction of Nitroaromatic Compounds and Catalytic Degradation of Methyl Orange. Catal. Lett. 2018, 148, 3701–3714. [Google Scholar] [CrossRef]
  54. Momeni, M.M.; Mirhosseini, M.; Mohammadi, N. Efficient photo catalytic degradation of methyl orange over Ag–CuO nanostructures grown on copper foil under visible light irradiation. J. Mater. Sci. Mater. Electron. 2016, 27, 6542–6551. [Google Scholar] [CrossRef]
  55. Alzahrani, S.A.; Malik, M.A.; Al-Thabaiti, S.A.; Khan, Z. Seedless synthesis and efficient recyclable catalytic activity of Ag@Fe nanocomposites towards methyl orange. Appl. Nanosci. 2018, 8, 255–271. [Google Scholar] [CrossRef]
Scheme 1. Catalytic reduction of 4-nitrophenol to 4-aminophenol.
Scheme 1. Catalytic reduction of 4-nitrophenol to 4-aminophenol.
Nanomaterials 09 00277 sch001
Figure 1. SEM images of (a) HKUST-1/Fe3O4/cellulose microfibril (CMF) composites and (b) carbon-doped CuO/Fe3O4 composite catalysts after calcination; TEM images of (c) carbon-doped CuO/Fe3O4 composite catalysts; (df) energy-dispersive X-ray spectroscopy (EDX) mapping of carbon-doped CuO/Fe3O4 composite catalysts.
Figure 1. SEM images of (a) HKUST-1/Fe3O4/cellulose microfibril (CMF) composites and (b) carbon-doped CuO/Fe3O4 composite catalysts after calcination; TEM images of (c) carbon-doped CuO/Fe3O4 composite catalysts; (df) energy-dispersive X-ray spectroscopy (EDX) mapping of carbon-doped CuO/Fe3O4 composite catalysts.
Nanomaterials 09 00277 g001
Figure 2. (a) High-resolution X-ray photoelectron spectroscopy (XPS) survey spectra of the carbon-doped CuO/Fe3O4 composite catalysts. (b) High-resolution XPS scans over C1s peaks of the carbon-doped CuO/Fe3O4 composite catalysts. (c) High-resolution XPS scans over Cu2p peaks of carbon-doped CuO/Fe3O4 composite catalysts. (d) High-resolution XPS scans over Fe2p peaks of carbon-doped CuO/Fe3O4 composite catalysts.
Figure 2. (a) High-resolution X-ray photoelectron spectroscopy (XPS) survey spectra of the carbon-doped CuO/Fe3O4 composite catalysts. (b) High-resolution XPS scans over C1s peaks of the carbon-doped CuO/Fe3O4 composite catalysts. (c) High-resolution XPS scans over Cu2p peaks of carbon-doped CuO/Fe3O4 composite catalysts. (d) High-resolution XPS scans over Fe2p peaks of carbon-doped CuO/Fe3O4 composite catalysts.
Nanomaterials 09 00277 g002
Figure 3. (a) Powder X-ray diffraction (PXRD) patterns of carbon-doped Fe3O4/CuO composite catalyst. (b) Magnetization curve at 300 K of carbon-doped Fe3O4/CuO composite catalyst.
Figure 3. (a) Powder X-ray diffraction (PXRD) patterns of carbon-doped Fe3O4/CuO composite catalyst. (b) Magnetization curve at 300 K of carbon-doped Fe3O4/CuO composite catalyst.
Nanomaterials 09 00277 g003
Figure 4. (a) Time-dependent ultraviolet–visible (UV–Vis) absorption spectra of 4-nitrophenol (4-NP) reduced by NaBH4 in the presence of carbon-doped CuO/Fe3O4 composite catalysts; (b) the first-order kinetic plot (absorbance at 400 nm, ln(At/A0)) versus reaction time for the reduction of 4-nitrophenol; At and A0 represent the absorbance values of 4-NP at 400 nm at designated time t and t = 0, respectively.
Figure 4. (a) Time-dependent ultraviolet–visible (UV–Vis) absorption spectra of 4-nitrophenol (4-NP) reduced by NaBH4 in the presence of carbon-doped CuO/Fe3O4 composite catalysts; (b) the first-order kinetic plot (absorbance at 400 nm, ln(At/A0)) versus reaction time for the reduction of 4-nitrophenol; At and A0 represent the absorbance values of 4-NP at 400 nm at designated time t and t = 0, respectively.
Nanomaterials 09 00277 g004
Figure 5. (a) Time-dependent UV–Vis absorption spectra of methylene blue (MB) with the carbon-doped CuO/Fe3O4 composite catalyst in the presence of NaBH4; (b) the corresponding first-order kinetic plot (absorbance at 660 nm, ln(At/A0)) versus reaction time for the reduction of MB; At and A0 represent the absorbance of MB (660 nm) at designated time t and t = 0, respectively; (c) time-dependent UV–Vis absorption spectra of methyl orange (MO) with the carbon-doped CuO/Fe3O4 composite catalyst in the presence of NaBH4; (d) the corresponding first-order kinetic plot (absorbance at 460 nm, ln(At/A0)) versus reaction time for the reduction of MO; At and A0 represent the absorbance of MO (460 nm) at designated time t and t = 0, respectively.
Figure 5. (a) Time-dependent UV–Vis absorption spectra of methylene blue (MB) with the carbon-doped CuO/Fe3O4 composite catalyst in the presence of NaBH4; (b) the corresponding first-order kinetic plot (absorbance at 660 nm, ln(At/A0)) versus reaction time for the reduction of MB; At and A0 represent the absorbance of MB (660 nm) at designated time t and t = 0, respectively; (c) time-dependent UV–Vis absorption spectra of methyl orange (MO) with the carbon-doped CuO/Fe3O4 composite catalyst in the presence of NaBH4; (d) the corresponding first-order kinetic plot (absorbance at 460 nm, ln(At/A0)) versus reaction time for the reduction of MO; At and A0 represent the absorbance of MO (460 nm) at designated time t and t = 0, respectively.
Nanomaterials 09 00277 g005
Figure 6. Reduction conversion ratio of 4-NP, MB, and MO after five successive cycles using the carbon-doped CuO/Fe3O4 catalyst.
Figure 6. Reduction conversion ratio of 4-NP, MB, and MO after five successive cycles using the carbon-doped CuO/Fe3O4 catalyst.
Nanomaterials 09 00277 g006
Table 1. Comparison of the catalytic performance for the reduction of 4-nitrophenol, methylene blue (MB), and methyl orange (MO) using the carbon-doped CuO/Fe3O4 composite catalyst and other reported catalysts in the literature. NP—nanoparticle.
Table 1. Comparison of the catalytic performance for the reduction of 4-nitrophenol, methylene blue (MB), and methyl orange (MO) using the carbon-doped CuO/Fe3O4 composite catalyst and other reported catalysts in the literature. NP—nanoparticle.
PollutantSamplesConcentration of PollutantTime (min)k (s−1)Reference
4-NitrophenolAgnp-PSAC0.02 mM253.9 × 10−3[47]
CF-AuNPs1 mM110.17 × 10−3[48]
CuO0.1 mM325.8 × 10−3[49]
CuO/Fe3O4/C1 mM106.5 × 10−3This work
Methylene BlueAu/Fe3O4@C0.01 mM105.5 × 10−3[50]
AuNPs1 mM94.0 × 10−3[51]
Fe3O4@polydopamine-Ag0.1 mM-7.2 × 10−3[52]
CuO/Fe3O4/C0.03 mM612.9 × 10−3This work
Methyl OrangeAg–γ-Fe2O30.3 mM301.53 × 10−3[53]
Ag-CuO0.006 mM1200.3 × 10−3[54]
Ag@Fe0.1 mM141.6 × 10−3[55]
CuO/Fe3O4/C0.03 mM252.4 × 10−3This work

Share and Cite

MDPI and ACS Style

Lu, H.; Zhang, L.; Ma, J.; Alam, N.; Zhou, X.; Ni, Y. Nano-Cellulose/MOF Derived Carbon Doped CuO/Fe3O4 Nanocomposite as High Efficient Catalyst for Organic Pollutant Remedy. Nanomaterials 2019, 9, 277. https://doi.org/10.3390/nano9020277

AMA Style

Lu H, Zhang L, Ma J, Alam N, Zhou X, Ni Y. Nano-Cellulose/MOF Derived Carbon Doped CuO/Fe3O4 Nanocomposite as High Efficient Catalyst for Organic Pollutant Remedy. Nanomaterials. 2019; 9(2):277. https://doi.org/10.3390/nano9020277

Chicago/Turabian Style

Lu, Hailong, Lili Zhang, Jinxia Ma, Nur Alam, Xiaofan Zhou, and Yonghao Ni. 2019. "Nano-Cellulose/MOF Derived Carbon Doped CuO/Fe3O4 Nanocomposite as High Efficient Catalyst for Organic Pollutant Remedy" Nanomaterials 9, no. 2: 277. https://doi.org/10.3390/nano9020277

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