Next Article in Journal
Enzymatic Production of Lauroyl and Stearoyl Monoesters of d-Xylose, l-Arabinose, and d-Glucose as Potential Lignocellulosic-Derived Products, and Their Evaluation as Antimicrobial Agents
Next Article in Special Issue
Heterogeneous Photocatalysis: A Solution for a Greener Earth
Previous Article in Journal
Facile Synthesis of Various ZrO2 Phases and ZrO2-MO2 (M = Ti, Hf) by Thermal Decomposition of a Single UiO-66 Precursor for Photodegradation of Methyl Orange
Previous Article in Special Issue
Natural Clay Modified with ZnO/TiO2 to Enhance Pollutant Removal from Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Visible Light Induced Nano-Photocatalysis Trimetallic Cu0.5Zn0.5-Fe: Synthesis, Characterization and Application as Alcohols Oxidation Catalyst

1
Faculty of Pharmacy, Al-Ahliyya Amman University, Amman 19328, Jordan
2
Department of Chemistry, The University of Jordan, Amman 11942, Jordan
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(6), 611; https://doi.org/10.3390/catal12060611
Submission received: 11 April 2022 / Revised: 14 May 2022 / Accepted: 27 May 2022 / Published: 2 June 2022
(This article belongs to the Special Issue Heterogeneous Photocatalysis: A Solution for a Greener Earth)

Abstract

:
Here, we report a visible light-induced-trimetallic catalyst (Cu0.5Zn0.5Fe2O4) prepared through green synthesis using Tilia plant extract. These nanomaterials were characterized for structural and morphological studies using powder x-ray diffraction (P-XRD), scanning electron microscopy (SEM) and thermogravimetric analysis (TGA). The spinel crystalline material was ~34 nm. In benign reaction conditions, the prepared photocatalyst oxidized various benzylic alcohols with excellent yield and selectivity toward aldehyde with 99% and 98%; respectively. Aromatic and aliphatic alcohols (such as furfuryl alcohol and 1-octanol) were photo-catalytically oxidized using Cu0.5Zn0.5Fe2O4, LED light, H2O2 as oxidant, 2 h reaction time and ambient temperature. The advantages of the catalyst were found in terms of reduced catalyst loading, activating catalyst using visible light in mild conditions, high conversion of the starting material and the recyclability up to 5 times without loss of the selectivity. Thus, our study offers a potential pathway for the photocatalytic nanomaterial, which will contribute to the advancement of photocatalysis studies.

Graphical Abstract

1. Introduction

The high demand to use renewable energy sources is required to reduce the dependency on fossil fuels and the possibility to reduce emissions in the atmosphere. Therefore, solar energy can be a good renewable alternative to fossil fuels. In recent years, nano semiconductor photocatalysis, as a “green” technology, has been widely used for conducting various applications. Outstanding stability, good photostability, nontoxicity and low price make spinel iron oxide the photocatalyst of choice for environmental remediation [1,2,3,4]. It is known that the UV region occupies only ~4% of the entire solar spectrum, while 45% of the energy belongs to visible light [5]. Therefore, developing efficient visible-light photocatalysts for environmental remediation has become an active research area in photocatalysis research [6,7,8,9,10,11,12]. In the photoelectric conversion process, the most important reaction involves hydroxyl ions on the semiconductor surface reacting with the holes, forming hydroxyl radicals (·OH), which is the main cause of the photocatalytic activity. The hydroxyl radical is a powerful, non-selective oxidant that can rapidly oxidize many organic compounds [13]. Many perovskite- or spinel-type complex oxides have been found to have selective visible-light-driven photoactivity using peroxides as oxidant. Among them, ferrites have also received considerable attention as magnetic nanoparticles (MNPs) with diverse applications metal oxides (e.g., Fe2O3, ZnFe2O4, NiCo2O4, MnFe2O4 and BiFeO3), as it may show visible light photocatalytic alternative [14,15,16,17,18,19].
Trimetallic iron-based oxides possess attractive properties and widespread applications [20]. One of the essential applications of trimetallic NPs is catalysis [21,22,23,24]. The catalyst efficiency surges exponentially by increasing the number of coordination sites through a significant increase in the surface/volume ratio of NPs. Glucose oxidation produces gluconic acid under mild conditions catalyzed by Ag/Au/Pd NPs [24]. The cyclization reaction between diazodicarbonyl compounds and oxalyl chloride was carried out through green AuFeAg catalyst to afford α,β- or β,β-dichloroenones [25]. Moreover, large-scale sensitive and selective organic transformations achieved using trimetallic NPs in various forms such as core/shell [23], alloys [24] and layer-by-layer [25].
The catalytic activity can be dependent on preparation methods, as it affects the morphology and physiochemical properties of the material. Therefore, several methods have been applied in preparation of multi-metallic NPs. Among the known methods, electrochemical synthesis, chemical and hydrothermal synthesis, coprecipitation, sol-gel, mechanical alloying and solvothermal technologies are main techniques utilized [26,27]. Here we use coprecipitation technique using a plant extract as a reducing agent instead of unnecessary reagents. Generally, natural product extracts obtained from bio-renewable sources have a great potential as reducing, stabilizing material as it contains polyphenolic, glycoside and flavonoid compounds. Furthermore, the waxy material in plant extract can help in preventing the aggregation of powder nanoparticles, thus it can be applied for the greener production of mono- and multi-metallic NPs [28,29].
Oxidation of benzyl alcohol to produce the must-need industrial intermediate benzaldehyde is a crucial process in the industry. Within the aldehyde family, benzaldehyde is an essential building block for many industrial products [30]. The attractive interest of benzyl alcohol and benzaldehyde in the research area is noticeable as increase of preparation protocols for benzaldehydes production in the last decade, as well its applications. The heterogeneous catalytic oxidation of aryl or alkyl alcohol to supply various worthy chemicals such as aldehydes or carboxylic acids has been endorsed beneficial over homogenous catalytic techniques due to their recoverable fashion and the help of eco-friendly conditions like the use of H2O2 or O2 as mild oxidants [31,32,33,34].
Herein, we report a trimetallic iron oxide-based NPs, as green and affordable catalyst using phyto mediated extract (Telia) in preparation of Cu0.5Zn0.5Fe2O4 that was applied as a photo-induced oxidation catalyst of benzyl alcohols and aliphatic alcohols. As shown in Scheme 1, benzyl alcohol oxidation is presented, exploiting mild reaction conditions. The characterization of the synthesized nanoparticles has been carried out in detail, employing the Rietveld refinement method using XRD data, SEM and TGA images.
To study the oxidation catalytic activity of Magnetic nanoparticles (MNP), benzyl alcohol was used as substrate model to optimize the reaction conditions. The present study has been extended to various alcohol substates to explore the possibilities of expanding the application of the catalyst, which shows a promising reactivity and applications. The novelty of the present work lies in the preparation of this magnetic trimetallic nanoparticles in a very simple, cheaper, eco-friendly coprecipitation synthesis process of nano particles, which can be used as a LED photo induced catalyst for oxidation of alcohols to obtain selectively aldehyde products.

2. Results and Discussion

2.1. Trimetallic Nanoparticles

The XRD patterns for the Cu0.5Zn0.5FeO4 sample are shown in Figure 1. The XRD spectra for the samples with the diffraction peaks that appeared at 2θ values of 30.00, 35.40, 43.31, 50.41, 56.94 and 62.53 were assigned to the (220) (311) (400) (511) (440) and (550) planes, which was found to be in good agreement with ZnFe2O4 (PDF no. 01-089-1009) and Fe3O4 (PDF no. 01-088-0866). The signals consistent with the (220), (311), (400), (511), (440) and (533) reflect the spinel crystal structure of CuFe2O4 (PDF no.00-025-0283). These results show the successful loading of the cubic Cu0.5Zn0.5 Fe2O4 spinel ferrites structures JCPDS: JCPDS: 00-051-0386. The lattice constant and crystallite size of the nanoparticles have been deliberated from the most conspicuous peak (311) using the Scherrer formula 35. The X-ray diffraction nanoparticles reveal that Cu0.5Zn0.5Fe2O4 NPs possess a single-phase major cubic (fcc) spinel. The experimentally perceived d arrangement values and proportional strengths agree with those described in the records.
The crystallite size (D) of each sample was calculated using the Stokes–Wilson Equation (1) [35,36]
D = λ β c cos θ
where λ is the X-ray wavelength, θ is the Bragg’s angle of diffraction in degrees and   β c is the integral breadth of the diffraction peak corrected for the instrumental broadening. The diffraction peak used for the determination of the crystallite size was fitted with the relation:
β = A I 0
Here A is the peak area and I0 is the maximum intensity. The most intense peak (311) at 2θ = 35.6° was fitted to provide information regarding crystallite size in the direction perpendicular to the corresponding crystallographic plane. In addition, the size of spinel phase was determined with different crystallographic directions using the (220) reflection at 2θ = 30.0° and the (400) reflection at 43.04°; thus, the crystalline spinel Cu0.5Zn0.5Fe2O4 size was 34 nm.
The morphology of the prepared CuZeFeO4 nanoparticles is explored by SEM image, as shown in Figure 2. The average particle size of the particles is ~31.2 nm.

2.2. TGA

Thermal stability of Cu0.5Zn0.5Fe2O4-NP was investigated by thermogravimetric analysis Figure 3. A gradual weight loss of 14 wt% was observed from 50–250 °C, owing to the reduction of water and CO2 molecules [37]. In addition, the second drop of weight (12 wt%) at 250–450 °C can be attributed to the degradation of -OH ion from organic moieties of plant extract residues and decomposition of inorganic salts, which has two degradation mechanism involves both intermolecular and intramolecular transfer reactions [38]. The final loss of 11% occurred at 500–600 °C due to the formation of pure corresponding metal oxide. Notably, above 700 °C no weight loss was observed, which indicates the formation of Cu0.5Zn0.5Fe2O4-NP.

2.3. Benzyl Alcohol Oxidation

The optimization of the catalytic reaction has been tested using benzyl alcohol as a model substrate, and the solvent effect is shown in Table 1. When using a different solvent, the reactivity changes drastically. To our delight, Acetonitrile has given the higher conversion for oxidizing benzyl alcohol (BA) into the anticipated aldehyde rather than isopropanol, while acetone and water have afforded a negligible amount of benzaldehyde. The high conversion of BA using acetonitrile can be explained via the formation of active radical species in the reaction media [39,40,41].
In terms of optimizing reaction conditions, the influences of catalyst load amount, H2O2, on the BA substate have been shown in (Table 2). The blank experiments in entries 1 and 6 prove the need for catalyst and the oxidizing agent to produce an acceptable percentage of benzaldehyde. Excellent conversion (99%) of BA was obtained when the amount of catalyst load was 10–20 mg. Therefore, we used 10 mg of the catalyst to reduce the catalytic waste. Similarly, the H2O2 amount was optimized and 350 µL has provided the best targeted conversion into benzaldehyde. The role of the peroxide can be explained that the photo-induced electrons in the conduction band would dissolve oxygen and peroxide species which form peroxy and OH radicals intermediates. The reactive OH radicals species can oxidize alcohol into aldehyde.
The effects of reaction time on the photocatalytic transformation of BA have been illustrated in Figure 4. The reaction was carried out up to 180 min, using the optimized conditions. In general, the oxidation rate of benzyl alcohol increased significantly when the reaction time raised for the first hour. The reaction ended up with a 99% yield in 120 min. This result shows that the catalyst is highly active in such conditions.
The catalytic performance is demonstrated by the four different illumination power from the LED source with an intensity of 7, 30, 60 and 120 W/cm2, while the other conditions are set carefully at optimum conversion previous results. The reaction substrates included 10 mg of Cu0.5Zn0.5Fe2O4 catalyst, 1 mmol of benzyl alcohol, 350 mg of H2O2 and 2 mL of acetonitrile. With increasing illumination power by 23 W, the oxidation rate of benzyl alcohol surges dramatically to produce the highest yield 93%, while obviously the best conversion appears under 60 watt light power (Table 3).
Aliphatic and aromatic alcohol derivatives were chosen to explore the selective oxidation of various alcohol derivatives by Cu0.5Zn0.5Fe2O4 catalyst under the optimized conditions (Table 4). To our delight, the catalyst has shown a good to excellent reactivity toward both aliphatic and aromatic alcohols. Apparently, aromatic alcohols have shown high reactivity entries 1–4. The substituted benzylic alcohols have shown no effect on the obtained yield, while cinnamyl alcohol and the influence of allylic bond led to lower reactivity than the other aromatic substrates. This allows us to apply the catalyst to a lignin substrate model such as furfuryl alcohol, and indeed the catalyst has excellent yield and selectivity.
When comparing our results to previously reported results, several studies conveyed to prepare adequate metallic catalysts to achieve this transformation, when most of them either use expensive metals like Pd, Pt and Rh [42,43,44] or electrical sources [45], while the reaction conditions depend on intense light power between 150 to 450 W, or high temperature 80 to 120 °C in both natural and synthetic sources. In addition, the percentage yield of benzaldehyde varies widely from around 28 to 99, shown in Table 5.

2.4. Recyclability

The stability of the prepared materials when exposed to consecutive reaction cycles was tested. Therefore, we examine the recyclability by using benzyl alcohol as a substrate in the presence of 10 mg of the catalyst.
The catalyst was reused up to four successive runs without drastic loss in activity. The catalyst was washed with water/ethanol solution and oven dried pre-use for oxidation of benzyl alcohol (Figure 5). Interestingly, after the 5th run, the weight loss of the catalyst was 28 wt% using gravimetric assay. The drop in the reactivity can be due to the slight deactivation of the catalyst via leaching some of the metals in the nanoparticles, or poising the catalyst surface of unwanted reaction products. After the fifth cycle, the conversion dropped slightly without any loss of selectivity.

3. Experimental

3.1. Materials

The solvents used were purchased from sigma-Aldrich (Taufkirchen, Germany). FeSO4·7H2O, Zn(OAc)2·2H2O CuSO4·5H2O and alcohols used were purchased from Sigma-Aldrich with 99.95% purity, and used as received. Tilia leaves extract was collected from a cultivated area in Amman, Jordan, during the flowering period in early spring of 2021. Sodium chloride (NaCl), acetic acid, sodium hydroxide and distilled water were used for all experiments. The crystallographic structure of MFe2O4 nano-ferrites was determined from the Powder X-ray diffraction (P-XRD, mnochromatic Cu-Kα radiation, nickel filter, 40 kV, 30 mA using Shimadzu XRD-7000, Japan) pattern, while the particle morphology and size distribution was determined with the aid of electron microscopy imaging (SEM, JOEL 6400 and FEI QUANTA 200, Japan).

3.2. Copper-Zinc Ferrite Cu0.5Zn0.5Fe2O4 NP Synthesis

The trimetallic ferrite nanoparticles were synthesized using an aqueous extract of Telia, as described previously [27]. Briefly, 1.12 g of FeSO4.7H2O and 0.48 g of Zn(OAC)2.2H2O and 0.48 g CuSO4·5H2O were dispersed in 20 mL distilled water for 1 h at room temperature under stirring, in order to obtain the homogeneous solution. A 50 mL of the aqueous plant extract was drop-wise added to the solution, and the mixture was stirred for 1 h and later the pH was controlled to obtain pH 10–12. The homogenous solution was then centrifuged, and the obtained solid products were collected and washed with distilled water and ethanol several times. Finally, the samples were dried in a vacuum oven at 60 °C for 6 h, and later calcinated at 1000 °C.

3.3. Characterization

The morphology and crystallinity of the synthesized Cu0.5Zn0.5Fe2O4 nanoparticles were determined via field emission scanning electron microscopy (SEM, FESEM JEOL JSM-6380) and (XRD, X’pert diffractometer using CuK α radiation), respectively. All measurements were performed at room temperature (25 ± 2 °C). Thermogravimetric analysis was accomplished using the System Setaram Setsys 12 TGA instrument (Setaram Instrumentation, Caluire, France), by heating the sample up to 800 °C at a rate of 10 °C·min−1.

3.4. Photocatalytic Oxidation of Benzyl Alcohol to Benzaldehyde

To a stirred solution of 2 mL of acetonitrile containing 0.1 mmol benzyl alcohol, 15 mg of the prepared catalyst and 350 µL of H2O2 was added drop wise pre-irradiation. Subsequently, the suspension was irradiated by a 60 W LED, and the reaction was monitored by the absorbance using UV-vis for each sample.

3.5. Product Analysis

When the reaction ended, the magnetic catalyst was removed by applying an external magnet and the product was extracted by column chromatography and analyzed by 1H-NMR and compared to the commercially available chemicals. The conversion percentage of benzyl alcohol was calculated by using the following Equation (3)
Conversion (%) = [(C0 − Cr)/C0] × 100%
where C0 is the initial concentration of benzyl alcohol and Cr and Ct are the concentrations of benzyl alcohol and benzaldehyde, respectively.

3.6. Experimental Procedure for the Reuse of the Catalyst

The reaction was applied for benzyl alcohol oxidation in a 1.0 mmol scale. Keeping reaction conditions constant, except using the recycled Cu0.5Zn0.5Fe2O4 catalyst rather than fresh catalyst, the reaction was monitored. When the reaction ended, the catalyst was removed using an external magnet and the product was analyzed by either 1H-NMR or UV-Vis. The catalyst was collected and washed with water/ethanol (5 mL) three times and oven dried pre-reuse. The dried catalyst was applied for further catalytic cycle up to 5 times without any severe loss of reactivity.

4. Conclusions

In this study, we synthesized phyto-mediated trimetallic Np of Cu0.5Zn0.5Fe2O4 using a Telia extract. The NP catalyst was characterized by powder XRD, FESEM and TGA studies. The XRD indicated a spinel ferrite with average crystalline diameter ~34 nm and average particle size of 31.2 nm by the SEM. The trimetallic nanoparticles has great potential to act as a visible light photocatalyst using LED light for the oxidation of benzylic and aliphatic alcohols selectively to aldehyde, with high selectivity and reactivity under benign reaction conditions using 10 mg of catalyst, 3 equivalent H2O2, 60 W light and 2 h reaction time. Later, the catalyst was compared to previous works, and it shows a promising reactivity. The novelty of this study lies in the preparation of the magnetic trimetallic nanoparticles in a very simple, cheap and eco-friendly coprecipitation synthesis process of nano particles, which offers a potential pathway for photocatalytic oxidation reaction.

Author Contributions

Conceptualization, A.A.-H.; methodology, A.A.-H. and A.G.; software, L.Y. and A.A.-H.; validation, A.A.-H. and A.G.; formal analysis, A.A.-H. and L.Y.; investigation, A.G. and A.A.-H.; resources, A.A.-H.; writing—original draft preparation, A.G. and A.A.-H.; writing—review and editing, A.A.-H. and A.A.-H.; visualization, A.G. and A.A.-H.; supervision, A.A.-H.; project administration, A.A.-H.; and funding acquisition, A.A.-H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Deanship of Academic Research at the University of Jordan (Grant No. 2364), and Al-Ahliyya Amman University.

Data Availability Statement

Not applicable.

Acknowledgments

The Jordanian Cell Therapy Center (CTC) is acknowledged for the SEM imaging. A.A.-H. acknowledges support by the Deanship of Scientific Research at the University of Jordan. A.G. would like to acknowledges support by Deanship of Scientific Research at Al-Ahliyya Amman University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ng, Y.H.; Lightcap, I.V.; Goodwin, K.; Matsumura, M.; Kamat, P.V. To What Extent Do Graphene Scaffolds Improve the Photovoltaic and Photocatalytic Response of TiO2 Nanostructured Films? J. Phys. Chem. Lett. 2010, 1, 2222–2227. [Google Scholar] [CrossRef]
  2. Zhang, H.; Lv, X.; Li, Y.; Wang, Y.; Li, J. P25-graphene Composite as a High Performance Photocatalyst. ACS Nano 2010, 4, 380–386. [Google Scholar] [CrossRef] [PubMed]
  3. Woan, K.; Pyrgiotakis, G.; Sigmund, W. Photocatalytic Carbon Nanotube TiO2 Composites. Adv. Mater. 2009, 21, 2233–2239. [Google Scholar] [CrossRef]
  4. Kumaresan, L.; Mahalakshmi, M.; Palanichamy, M.; Murugesan, V. Synthesis, Characterization, and Photocatalytic Activity of Sr2þ Doped TiO2 Nanoplates. Ind. Eng. Chem. Res. 2010, 49, 1480–1485. [Google Scholar] [CrossRef]
  5. Zhang, Z.J.; Wang, W.Z.; Yin, W.Z.; Shang, M.; Wang, L.; Sun, S.M. Inducing Photocatalysis by Visible Light Beyond the Absorption Edge: Effect of Upconversion Agent on the Photocatalytic Activity of Bi2WO6. Appl. Catal. B 2010, 101, 68–73. [Google Scholar] [CrossRef]
  6. Liang, Y.Y.; Wang, H.L.; Casalongue, H.S.; Chen, Z.; Dai, H.J. TiO2 Nanocrystals Grown on Graphene as Advanced Photocatalytic Hybrid Materials. Nano Res. 2010, 3, 701–705. [Google Scholar] [CrossRef] [Green Version]
  7. Vijayan, B.K.; Dimitrijevic, N.M.; Wu, J.S.; Gray, K.A. The Effects of Pt Doping on the Structure and Visible Light Photoactivity of Titania Nanotubes. J. Phys. Chem. C 2010, 114, 21262–21269. [Google Scholar] [CrossRef]
  8. Romcevic, N.; Kostic, R.; Hazic, B.; Romcevic, M.; KuryliszynKudelska, I.; Dobrowolski, W.D.; Narkiewicz, U.; Sibera, D. Raman Scattering from ZnO Incorporating Fe Nanoparticles: Vibrational Modes and Low-Frequency Acoustic Modes. J. Alloys Compd. 2010, 507, 386–390. [Google Scholar] [CrossRef]
  9. Cai, L.; Liao, X.; Shi, B. Using Collagen Fiber as a Template to Synthesize TiO2 and Fex/TiO2 Nanofibers and Their Catalytic Behaviors on the Visible Light-Assisted Degradation of Orange II. Ind. Eng. Chem. Res. 2010, 49, 3194–3199. [Google Scholar] [CrossRef]
  10. Guo, R.Q.; Fang, L.A.; Dong, W.; Zheng, F.G.; Shen, M.R. Enhanced Photocatalytic Activity and Ferromagnetism in Gd Doped BiFeO3 Nanoparticles. J. Phys. Chem. C 2010, 114, 21390–21396. [Google Scholar] [CrossRef]
  11. Cai, W.D.; Chen, F.; Shen, X.X.; Chen, L.J.; Zhang, J.L. Enhanced Catalytic Degradation of AO7 in the CeO2-H2O2 System with Fe3þ Doping. Appl. Catal. B 2010, 101, 160–168. [Google Scholar] [CrossRef]
  12. Shu, X.; He, J.; Chen, D. Visible-Light-Induced Photocatalyst Based on Nickel Titanate Nanoparticles. Ind. Eng. Chem. Res. 2008, 47, 4750–4753. [Google Scholar] [CrossRef]
  13. Hirakawa, T.; Nosaka, Y. Properties of O2 and OH Formed in TiO2 Aqueous Suspensions by Photocatalytic Reaction and the Influence of H2O2 and Some Ions. Langmuir 2002, 18, 3247–3325. [Google Scholar] [CrossRef]
  14. Xu, S.H.; Feng, D.L.; Shangguan, W.F. Preparations and Photocatalytic Properties of Visible-Light-Active Zinc Ferrite-Doped TiO2 Photocatalyst. J. Phys. Chem. C 2009, 113, 2463–2467. [Google Scholar] [CrossRef]
  15. Laokul, P.; Amornkitbamrung, V.; Seraphin, S.; Maensiri, S. Characterization and Magnetic Properties of Nanocrystalline CuFe2O4, NiFe2O4, ZnFe2O4 Powders Prepared by the Aloe Vera Extract Solution. Curr. Appl. Phys. 2011, 11, 101–108. [Google Scholar] [CrossRef]
  16. Zhang, B.P.; Zhang, J.L.; Chen, F. Preparation and Characterization of Magnetic TiO2/ZnFe2O4 Photocatalysts by a sol-gel Method. Res. Chem. Intermed. 2008, 34, 375–380. [Google Scholar] [CrossRef]
  17. Gómez-Pastora, J.; Bringas, E.; Ortiz, I. Recent progress and future challenges on the use of high performance magnetic nano-adsorbents in environmental applications. Chem. Eng. J. 2014, 256, 187–204. [Google Scholar] [CrossRef]
  18. Al-Hunaiti, A.; Ghazzy, A.; Sweidan, N.; Mohaidat, Q.; Bsoul, I.; Mahmood, S.; Hussein, T. Nano-Magnetic NiFe2O4 and Its Photocatalytic Oxidation of Vanillyl Alcohol—Synthesis, Characterization, and Application in the Valorization of Lignin. Nanomaterials 2021, 11, 1010. [Google Scholar] [CrossRef]
  19. Al-Hunaiti, A.; Mohaidat, Q.; Bsoul, I.; Mahmood, S.; Taher, D.; Hussein, T. Synthesis and characterization of novel phyto-mediated catalyst, and its application for a selective oxidation of (VAL) into vanillin under visible light. Catalysts 2020, 10, 839. [Google Scholar] [CrossRef]
  20. Sharma, G.; Kumar, D.; Kumar, A.; Ala’a, H.; Pathania, D.; Naushad, M.; Mola, G.T. Revolution from monometallic to trimetallic nanoparticle composites, various synthesis methods and their applications: A review. Mater. Sci. Eng. C 2017, 71, 1216–1230. [Google Scholar] [CrossRef]
  21. Mondal, B.N.; Basumallick, A.; Chattopadhyay, P.P. Magnetic behavior of nanocrystalline Cu–Ni–Co alloys prepared by mechanical alloying and isothermal annealing. J. Alloys Compd. 2008, 457, 10–14. [Google Scholar] [CrossRef]
  22. Roshanghias, A.; Bernardi, J.; Ipser, H. An attempt to synthesize Sn-Zn-Cu alloy nanoparticles. Mater. Lett. 2016, 178, 10–14. [Google Scholar] [CrossRef]
  23. Lan, J.; Li, C.; Liu, T.; Yuan, Q. One-step synthesis of porous PtNiCu trimetallic nanoalloy with enhanced electrocatalytic performance toward methanol oxidation. J. Saudi Chem. Soc. 2019, 23, 43–51. [Google Scholar] [CrossRef]
  24. Venkatesan, P.; Santhanalakshmi, J. Designed synthesis of Au/Ag/Pd trimetallic nanoparticle-based catalysts for Sonogashira coupling reactions. Langmuir 2010, 26, 12225–12229. [Google Scholar] [CrossRef]
  25. Duan, H.; Wang, D.; Li, Y. Green chemistry for nanoparticle synthesis. Chem. Soc. Rev. 2015, 44, 5778–5792. [Google Scholar] [CrossRef]
  26. Iravani, S.; Varma, R. Plant-derived Edible Nanoparticles and miRNAs: Emerging Frontier for Therapeutics and Targeted Drug-delivery. ACS Sustain. Chem. Eng. 2019, 7, 8055–8069. [Google Scholar] [CrossRef]
  27. Imraish, A.; Al-Hunaiti, A.; Abu-Thiab, T.; Ibrahim, A.; Hwaitat, E.; Omar, A. Phyto-Facilitated Bimetallic ZnFe2O4 Nanoparticles via Boswellia carteri: Synthesis, Characterization, and Anti-Cancer Activity. Anti-Cancer Agents Med. Chem. 2021, 21, 1767–1772. [Google Scholar] [CrossRef]
  28. Iravani, S.; Varma, R.S. Plants and plant-based polymers as scaffolds for tissue engineering. Green Chem. 2019, 21, 4839–4867. [Google Scholar] [CrossRef]
  29. Pugh, S.; McKenna, R.; Halloum, I.; Nielsen, D.R. Engineering Escherichia coli for Renewable Benzyl Alcohol Production. Metab. Eng. Commun. 2015, 2, 39–45. [Google Scholar] [CrossRef]
  30. Pillai, U.R.; Sahle-Demessie, E. Oxidation of alcohols over Fe3+/montmorillonite-K10 using hydrogen peroxide. Appl. Catal. A-Gen. 2003, 245, 103–109. [Google Scholar] [CrossRef]
  31. O’Brien, P.; Lopez-Tejedor, D.; Benavente, R.; Palomo, J.M. Pd Nanoparticles-Polyethylenemine-Lipase Bionanohybrids as Heterogeneous Catalysts for Selective Oxidation of Aromatic Alcohols. ChemCatChem 2018, 10, 4992–4999. [Google Scholar] [CrossRef]
  32. Alshammari, H.M.; Alshammari, A.S.; Humaidi, J.R.; Alzahrani, S.A.; Alhumaimess, M.S.; Aldosari, O.F.; Hassan, H. Au-Pd Bimetallic Nanocatalysts Incorporated into Carbon Nanotubes (CNTs) for Selective Oxidation of Alkenes and Alcohol. Processes 2020, 8, 1380. [Google Scholar] [CrossRef]
  33. An, H.; Deng, C.; Sun, Y.; Lv, Z.; Cao, L.; Xiao, S.; Zhao, L.; Yin, Z. Design of Au@ Ag/BiOCl–OV photocatalyst and its application in selective alcohol oxidation driven by plasmonic carriers using O2 as the oxidant. CrystEngComm 2020, 22, 6603–6611. [Google Scholar] [CrossRef]
  34. Smit, J.; Wijn, H.P.J.; Ferrites; Mahmood, S.H. Properties and Synthesis of Hexaferrites. In Hexaferrite Permanent Magnetic Material; Wiley: New York, NY, USA, 1959; p. 54. [Google Scholar]
  35. Mahmood, S.H. Magnetic anisotropy in fine magnetic particles. J. Magn. Magn. Mater. 1993, 118, 359–364. [Google Scholar] [CrossRef]
  36. Narayanasamy, A.; Jeyadevan, B.; Chinnasamy, C.N.; Ponpandian, N.; Greneche, J.M. Structural, magnetic and electrical properties of spinel ferrite nanoparticles. In Proceedings of the Ninth International Conference on Ferrites, San Francisco, CA, USA, 3 January 2005; pp. 867–875. [Google Scholar]
  37. Rincón-Granados, K.L.; Vázquez-Olmos, A.R.; Vega-Jiménez, A.; Ruiz, F.; Garibay-Febles, V.; Ximénez-Fyvie, L. Preparation, characterization and photocatalytic activity of NiO, Fe2O3 and NiFe2O4. Materialia 2021, 15, 177–182. [Google Scholar]
  38. Marmisollé, W.A.; Azzaroni, O. Recent developments in the layer-by-layer assembly of polyaniline and carbon nanomaterials for energy storage and sensing applications. From synthetic aspects to structural and functional characterization. Nanoscale 2016, 8, 9890–9918. [Google Scholar] [CrossRef]
  39. Jiang, C.; Markutsya, S.; Tsukruk, V.V. Compliant, robust, and truly nanoscale free-standing multilayer films fabricated using spin-assisted layer-by-layer assembly. Adv. Mater. 2004, 16, 157–161. [Google Scholar] [CrossRef]
  40. Kroschwitz, J.I.; Howe-Grant, M. Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley and Sons: New York, NY, USA, 1991; p. 127. [Google Scholar]
  41. Al-Hunaiti, A.; Al-Said, N.; Halawani, L.; Haija, M.A.; Baqaien, R.; Taher, D. Synthesis of magnetic CuFe2O4 nanoparticles as green catalyst for toluene oxidation under solvent-free conditions. Arab. J. Chem. 2020, 13, 4945–4953. [Google Scholar] [CrossRef]
  42. Karimi, B.; Abedi, S.; Clark, J.; Budarin, V. Highly efficient aerobic oxidation of alcohols using a recoverable catalyst: The role of mesoporous channels of SBA-15 in stabilizing palladium nanoparticles. Angew. Chem. Int. Ed. 2006, 17, 4776–4779. [Google Scholar] [CrossRef]
  43. Wang, C.; Chen, W.; Chang, H.T. Enzyme Mimics of Au/Ag Nanoparticles for Fluorescent Detection of Acetylcholine. Anal. Chem. 2012, 84, 9706–9712. [Google Scholar] [CrossRef]
  44. Zhu, Z.; Liu, F.; Fan, J.; Li, Q.; Min, Y.; Xu, Q. C2 Alcohol Oxidation Boosted by Trimetallic PtPbBi Hexagonal Nanoplates. ACS Appl. Mater. Interfaces 2020, 12, 52731–52740. [Google Scholar] [CrossRef] [PubMed]
  45. Miedziak, P.; Sankar, M.; Dimitratos, N.; Lopez-Sanchez, J.A.; Carley, A.F.; Knight, D.W.; Brian, T.; Christopher, J.K.; Hutchings, G.J. Oxidation of benzyl alcohol using supported gold–palladium nanoparticles. Catal. Today 2011, 164, 315–319. [Google Scholar] [CrossRef]
  46. Higashimoto, S.; Kitao, N.; Yoshida, N.; Sakura, T.; Azuma, M.; Ohue, H.; Sakata, Y. Selective photocatalytic oxidation of benzyl alcohol and its derivatives into corresponding aldehydes by molecular oxygen on titanium dioxide under visible light irradiation. J. Catal. 2009, 266, 279–285. [Google Scholar] [CrossRef]
  47. Du, M.; Zeng, G.; Huang, J.; Sun, D.; Li, Q.; Wang, G.; Li, X. Green photocatalytic oxidation of benzyl alcohol over noble-metal-modified H2Ti3O7 nanowires. ACS Sustain. Chem. Eng. 2019, 7, 9717–9726. [Google Scholar] [CrossRef]
  48. Hong, Y.; Jing, X.; Huang, J.; Sun, D.; Odoom-Wubah, T.; Yang, F.; Li, Q. Biosynthesized bimetallic Au–Pd nanoparticles supported TiO2 for solvent-free oxidation of benzyl alcohol. ACS Sustain. Chem. Eng. 2014, 2, 1752–1759. [Google Scholar] [CrossRef]
  49. Veisi, H.; Hemmati, S.; Qomi, M. Aerobic oxidation of benzyl alcohols through biosynthesized palladium nanoparticles mediated by Oak fruit bark extract as an efficient heterogeneous nanocatalyst. Tetrahedron Lett. 2017, 58, 4191–4496. [Google Scholar] [CrossRef]
Scheme 1. Benzyl alcohol oxidation by using trimetallic Cu-Zn-Fe NPs in Acetonitrile.
Scheme 1. Benzyl alcohol oxidation by using trimetallic Cu-Zn-Fe NPs in Acetonitrile.
Catalysts 12 00611 sch001
Figure 1. XRD patterns with Rietveld refinement the Cu0.5Zn0.5Fe2O4 sample.
Figure 1. XRD patterns with Rietveld refinement the Cu0.5Zn0.5Fe2O4 sample.
Catalysts 12 00611 g001
Figure 2. (a) FESEM images of the Cu0.5Zn0.5Fe2O4-NP; and (b) particle size distribution with a Gaussian fit average particle size (nm) = 31.2135 ± 0.23159.
Figure 2. (a) FESEM images of the Cu0.5Zn0.5Fe2O4-NP; and (b) particle size distribution with a Gaussian fit average particle size (nm) = 31.2135 ± 0.23159.
Catalysts 12 00611 g002
Figure 3. The thermogravimetric analysis (TGA) of the prepared nanoparticles.
Figure 3. The thermogravimetric analysis (TGA) of the prepared nanoparticles.
Catalysts 12 00611 g003
Figure 4. Effects of reaction time on the conversion of benzyl alcohol.
Figure 4. Effects of reaction time on the conversion of benzyl alcohol.
Catalysts 12 00611 g004
Figure 5. Recyclability test of Cu0.5Zn0.5Fe2O4-NP catalyst. The weight loss of catalyst after 5th runs was 28 wt%.
Figure 5. Recyclability test of Cu0.5Zn0.5Fe2O4-NP catalyst. The weight loss of catalyst after 5th runs was 28 wt%.
Catalysts 12 00611 g005
Table 1. Effect of solvents on the oxidation of benzyl alcohol.
Table 1. Effect of solvents on the oxidation of benzyl alcohol.
EntrySolventYield * %
1Acetonitrile74
2Isopropanol67
3Aceton7
4Water2
* Benzaldehyde yield was analyzed using 1H-NMR and UV-vis using internal standard. Reaction condition: benzyl alcohol = 0.1 mmol, catalyst = 15 mg, H2O2 = 250 mg, solvent 2 mL, light power 60 Watt, reaction time 4 h, temperature 25 °C.
Table 2. Quantitative optimization of Cu-Zn-Fe and oxidant amount.
Table 2. Quantitative optimization of Cu-Zn-Fe and oxidant amount.
EntryCu-Zn-Fe (mg)H2O2 (µL)Yield * %
103501
2735095
31035099
41535099
52035089
61009
71025054
81035099
91050099
* Benzaldehyde yield. Reaction condition: benzyl alcohol = 0.1 mmol, catalyst Cu0.5Zn0.5Fe2O4, ACN 2 mL, light power 60 Watt, reaction time 3 h, temperature 25 °C.
Table 3. Benzyl alcohol oxidation with different light power.
Table 3. Benzyl alcohol oxidation with different light power.
EntryLight Power (W)Yield * %
1730.0
23093.0
36099
412099
* Benzaldehyde yield. Reaction condition: (benzyl alcohol = 0.1 mmol, catalyst = 10 mg, Solvent ACN = 2 mL, H2O2 = 350 mg), reaction time 2 h, Temperature 25 °C.
Table 4. Alcohol derivatives on the selective oxidation of benzyl alcohol.
Table 4. Alcohol derivatives on the selective oxidation of benzyl alcohol.
EntrySubstrateAlcoholAldehydeYield * %
11-Octanol Catalysts 12 00611 i001 Catalysts 12 00611 i00229.2
22-Octanol Catalysts 12 00611 i003 Catalysts 12 00611 i00434.5
3Furfuryl
alcohol
Catalysts 12 00611 i005 Catalysts 12 00611 i00699
4Cinnamyl
alcohol
Catalysts 12 00611 i007 Catalysts 12 00611 i00861.4
53,4-dimethyl benzyl alcohol Catalysts 12 00611 i009 Catalysts 12 00611 i01099
6Benzyl alcohol Catalysts 12 00611 i011 Catalysts 12 00611 i01299
* Benzaldehyde yield. Reaction condition: (alcohol = 0.1 mmol, catalyst = 10 mg, Solvent ACN = 2 mL, H2O2 = 350 µL, light power 60 Watt, reaction time 2 h, Temperature 25 °C.
Table 5. A comparison studies using di- and trimetallic NPs as catalyst to oxidize alcohols.
Table 5. A comparison studies using di- and trimetallic NPs as catalyst to oxidize alcohols.
SourceMetallic NPsReaction ConditionsConversion %Selectivity %Ref.
SynthaticAu-Pd/O-CNTs2 h, 120 °C, Solvent free28.396[42]
SynthaticAu@Ag/BiOCl–OV1 h, Xenon 300 W, CH3CN9299[43]
SynthaticTrimetallic PtPbBiElectrical--[44]
Synthatic2.5% Au + 2.5% Pd/TiO2DP2140 °C, 10 bar O28440[45]
SynthaticPd/H2Ti3O7 NanowiresHalogen 150 W,
90 °C
49.576[46]
SynthaticTiO2LED 450 W, 36 h, O29999[47]
Nature
(Cacumen Platycladi)
Au−Pd/TiO26 h, 90 °C, O274.298[48]
Nature
(Oak fruit bark)
Pd NPs12 h, 80 °C, K2CO395-[49]
This work (Tilia)Trimetallic Cu-Zn-Fe2 h, LED 60 W, H2O29999
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ghazzy, A.; Yousef, L.; Al-Hunaiti, A. Visible Light Induced Nano-Photocatalysis Trimetallic Cu0.5Zn0.5-Fe: Synthesis, Characterization and Application as Alcohols Oxidation Catalyst. Catalysts 2022, 12, 611. https://doi.org/10.3390/catal12060611

AMA Style

Ghazzy A, Yousef L, Al-Hunaiti A. Visible Light Induced Nano-Photocatalysis Trimetallic Cu0.5Zn0.5-Fe: Synthesis, Characterization and Application as Alcohols Oxidation Catalyst. Catalysts. 2022; 12(6):611. https://doi.org/10.3390/catal12060611

Chicago/Turabian Style

Ghazzy, Asma, Lina Yousef, and Afnan Al-Hunaiti. 2022. "Visible Light Induced Nano-Photocatalysis Trimetallic Cu0.5Zn0.5-Fe: Synthesis, Characterization and Application as Alcohols Oxidation Catalyst" Catalysts 12, no. 6: 611. https://doi.org/10.3390/catal12060611

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