Selective synthesis of 2, 5-furandicarboxylic acid by oxidation of 5-hydroxymethylfurfural over MnFe2O4 catalyst

Selective synthesis of 2, 5-furandicarboxylic acid by oxidation of 5-hydroxymethylfurfural over MnFe2O4 catalyst

Catalysis Today 309 (2018) 119–125 Contents lists available at ScienceDirect Catalysis Today journal homepage: www.elsevier.com/locate/cattod Selec...

805KB Sizes 1 Downloads 26 Views

Catalysis Today 309 (2018) 119–125

Contents lists available at ScienceDirect

Catalysis Today journal homepage: www.elsevier.com/locate/cattod

Selective synthesis of 2, 5-furandicarboxylic acid by oxidation of 5hydroxymethylfurfural over MnFe2O4 catalyst Anil B. Gawade, Akhil V. Nakhate, Ganapati D. Yadav

T



Department of Chemical Engineering, Institute of Chemical Technology, Nathalal Parekh Marg, Matunga, Mumbai, 400 019, India

A R T I C L E I N F O

A B S T R A C T

Keywords: 5-Hydroxymethylfurfural 2, 5-Furandicarboxylic acid Oxidation Ferrites Green chemistry Biomass conversion

Development of green catalytic processes using biobased feedstock for valuable chemicals such as 5-hydroxymethylfurfural (HMF) is hotly pursued. 2, 5-Furandicarboxylic acid (FDCA), a bio-based alternative to terephthalic acid was efficiently synthesized by oxidation of HMF using MnFe2O4 spinel structured magnetic nanoparticles (MNPs) and tert-butyl hydroperoxide as an oxidant. MnFe2O4 catalyst showed the highest activity and selectivity and gave 85% yield of FDCA at 100 °C in 5 h. The higher activity of MnFe2O4 catalyst is due to the variable oxidation state of manganese. The combination of MnFe2O4 catalyst and TBHP oxidant requires less time and energy compared to other reported processes of FDCA synthesis. Also, many reported methods have used a homogeneous base for FDCA synthesis which is totally avoided in the current process. The concentration profiles of reactants and products were established and kinetics determined. The effects of various reaction parameters were studied to validate kinetic model. The catalyst was easily recycled due to its magnetic property and showed good catalytic activity up to four cycles. All metal ferrites were characterized by different analytical techniques. The catalyst maintained its fidelity.

1. Introduction Nowadays, the use of biomass is gaining importance for the synthesis of value added chemicals due to the depletion of fossil raw materials [1–3]. In this regard, 5-hydroxymethylfurfural (HMF) and its derivatives play a vital role. HMF is derived from sugars like fructose and glucose after triple dehydration by using acid catalysts [4,5]. It can be further valorized into various useful chemicals such as levulinic acid, 2, 5- dimethylfuran (DMF), ɣ-valerolactone, 2, 5-diformylfuran (DFF) and 2, 5-furandicarboxyllic acid (FDCA) [6–9]. From these chemicals, the oxidation product of HMF i.e. FDCA is one of the promising chemicals because of its potential use in polymer industry. It acts as an alternative for terephthalic acid, which is used in the synthesis of polyethylene terephthalate (PET) [10,11]. PET is widely used in synthesis of plastic bottles, food containers, fibers, photographic films, agrochemicals, pharmaceuticals and cosmetics [12,13]. Similar to terephthalic acid, FDCA can also transformed into poly (ethylene 2, 5-furandicarboxylate) (PEF) polymer, which has similar properties to PET [14]. In accordance with U.S. department of energy, FDCA can act as one of the twelve key platform chemicals obtained from biomass [15]. Till now, many processes are reported for production of FDCA by using homogenous and heterogeneous catalysts. Homogenous catalytic processes involved the use of KMnO4, K2Cr2O7, Co(OAc)2/Zn(OAc)2/



Corresponding author. E-mail addresses: [email protected], [email protected] (G.D. Yadav).

http://dx.doi.org/10.1016/j.cattod.2017.08.061 Received 8 April 2017; Received in revised form 15 August 2017; Accepted 24 August 2017 Available online 01 September 2017 0920-5861/ © 2017 Elsevier B.V. All rights reserved.

Br− [11] and metal/bromide salts [16]. All these methods produce toxic metal waste after their use and thus are not convenient for FDCA production. Several heterogeneous catalytic processes are also reported for FDCA synthesis. Usually these methods involve the use of noble metals like Pt, Au or Ru and addition of liquid base. Supported platinum catalysts consist of Pt/Al2O3 [17], Pt/ZrO2 [11], Pt/C with Bi as a promoter [6] and Pt nanoparticles with PVP as a stabilizer [18]. Gold catalysts used for HMF oxidation consist of Au/TiO2 [19], Au3Cu1/TiO2 [20] and Au-Pd alloy supported on carbon nanotube [21]. These catalysts give good yield of FDCA but are expensive and require addition of liquid base NaOH and high oxygen pressure. Recently, Gupta et al. demonstrated base-free oxidation of HMF into the FDCA by using Au/ HT catalyst [13]. However, magnesium is found to leach out from the support hydrotalcite (HT) due to interaction of acidic product FDCA and basic hydrotalcite. Davis et al. used carbon nanofibres (CNF) supported Pt and Au catalyst for HMF oxidation [22]. Gorbanev et al. used ruthenium based catalysts in which Ru(OH)X was supported on various supports like TiO2, ZrO2, CeO2, MgO2, La2O3, Al2O3, HT and hydroxyapetite by using dioxygen as oxidant [23]. The catalytic processes other than noble metal such as, FeIII-POP-1 [24] and CuCl [25] are also reported for FDCA production. Recently Neatu et al. reported FDCA synthesis by using Mn/Fe mixed oxide in presence of NaOH base [26]. FDCA can also be directly obtained in one pot from fructose by using

Catalysis Today 309 (2018) 119–125

A.B. Gawade et al.

2.4. Method of analysis

cobalt acetylacetonate [27] and various other catalysts [28]. Many of above reported method includes addition of liquid base, high temperature and high pressure for oxidation of HMF into FDCA. The problem associated with usages of base at high temperature is the degradation of HMF to undesirable products and hence making the process uneconomical. So there is a need to develop a heterogeneous catalytic process for HMF oxidation at ambient reaction conditions. Also, the use of stoichiometric oxidants like hydrogen peroxide (H2O2) and tert-butyl hydroperoxide (t-BuOOH) with heterogeneous catalyst is limited for this reaction. Only a single homogeneous method is reported in which Hansen et al. used t-BuOOH and CuCl in presence of nitrogen containing promoters [25]. This method produces large amount of wastes and so it is not clean and economical. Here we have developed a mild and energy efficient process for FDCA synthesis. Various metal ferrite MFe2O4 (M = Cu, Mn, Co, Mg) were synthesized by hydrothermal method and used for HMF oxidation along with t-BuOOH as a stoichiometric oxidant. Because of the nano size and magnetism of spinel ferrites, their use as catalysts is gaining impetus in various organic reaction transformations [29–31]. Spinel ferrites also exhibit other advantages such as stability at higher temperature, easy separation from reaction mixture, monodispersity, and low cost [32]. TBHP is a highly active oxygen containing cheaper and environmental friendly oxidant because its reduction product is tertbutyl alcohol which can be easily removed from reaction mixture [33]. Among all metal ferrites, MnFe2O4 was found to be most active and selective for FDCA synthesis. This catalytic system is better than other reported processes as no external base was added during the reaction. The catalyst was simply separated by using external magnet from reaction mixture and so it prevents catalyst loss which occurs in filtration and makes the process more economic. The concentrations of all intermediates observed in this reaction were determined and kinetic data obtained. Various reaction parameters were studied and a kinetic model successfully developed.

The reaction samples were analyzed by using HPLC (Agilent 1260 infinity) equipped with Agilent Hi-plex H column (300 × 7.7 mm).A 0.005 M H2SO4 solution was used as mobile phase with 0.5 mL/min flow rate. All compounds were identified by calibrating authentic commercial samples on UV and RI detectors. 3. Results and discussion 3.1. Catalyst characterization The characterization of MFe2O4 was performed by various analytical techniques like XRD, BET surface area, GA, FTIR, SEM and TEM. 3.1.1. X-ray powder diffraction (XRD) The textural properties and crystallinity of MFe2O4 spinel ferrites were studied by using XRD. The powder XRD patterns of all catalysts were performed on X-ray diffractometer (Bruker AXS, D8 Discover, USA) with Cu Kα radiation. The crystal structure of the mesoporous spinel ferrite nanospheres were analyzed via X-ray powder diffraction measurements. Fig. 1 shows typical XRD patterns of various MFe2O4 (Co, Mg, Mn, Cu & Fe) nanospheres. The XRD analysis reveals that all samples show the spinel cubic structure, exhibiting six prominent peaks 2θ = 30.1, 35.4, 43.1, 50.3, 53.2, 62.6, and 74.08 which originate from the (220), (311), (400), (422), (511) and (440) planes, which reflects cubic spinel structure of all MFe2O4. The diffraction patterns perfectly match with the expected cubic spinel structure of Co, Mg, Mn, Cu and Fe ferrites and have been identified as (JCPDS 22-1086), (JCPDS 893084) (JCPDS 38-0430), (JCPDS 34-0425), and (JCPDS 19-0629), respectively. 3.1.2. Fourier transform infrared spectroscopy (FTIR) FTIR spectra for all catalysts were obtained on a Perkin Elmer Spectrophotometer in the range of 400–4000 cm−1 (Fig. 2). The IR spectra for all MFe2O4 catalyst show two peaks in the range of 400 to 650 cm−1, The absorption band at 446 cm−1 is that of tetrahedral M3+ (M-O mode) and at 634 cm−1 represents stretching vibration frequency of octahedral M2+ (M-O mode) complexes. The peak near 1100 cm−1 designates the CeOeC stretching vibration of ethylene glycol. The band near 1600 and 3500 cm−1 indicate the bending and stretching vibrations of water respectively.

2. Experimental section 2.1. Chemicals All chemicals were purchased from commercial sources except HMF which was procured from Acros Organics (India) and no further purification was done before the use of all chemicals. 2.2. Catalyst synthesis All metal nano particles (MNPs) were synthesized by hydrothermal method as reported earlier [31,34,35]. A brief process is given in which 10 mmol FeCl3·6H2O and 5 mmol MCl2·4H2O (M = Cu, Co, Mn, Mg, Fe) were dissolved in 150 mL ethylene glycol. Then, in this clear solution were added 7 g NaOAc and 2 g polyethylene glycol and the resulting solution was stirred for 1 h. Then it was transferred into teflon lined stainless steel bomb and heated at 200° C for 10 h to produce a black material. After cooling at room temperature, the black material was washed with ethanol and dried at 100° C for 5 h. Ethylene glycol was used as a solvent for catalyst preparation due to its surfactant property which prevents the agglomeration of MFe2O4 MNPs. 2.3. Experimental setup All reactions were carried out in 100 cm3 autoclave (Amar autoclave, Mumbai) equipped with a pitched turbine impeller. A standard reaction was carried out with 1 mmol of HMF, 9 mmol of TBHP and 0.015 g/cm3 of catalyst loading which was made up to 20 mL by adding acetonitrile as solvent. The reaction mixture was heated at 100° C and agitation speed maintained at 1000 rpm. Sampling was done at regular intervals for further analysis.

Fig. 1. XRD of different metal ferrite spinel catalysts (a) CoFe2O4 (b) MgFe2O4 (c) MnFe2O4 (d) CuFe2O4 (e) Fe3O4.

120

Catalysis Today 309 (2018) 119–125

A.B. Gawade et al.

Fig. 3. TGA analysis of MnFe2O4.

of oxidant, a control experiment was carried out without addition of oxidant and it was found that the reaction did not proceed. 3.3. Effect of various solvent

Fig. 2. FT-IR spectra of catalysts (a) MnFe2O4 (b) MgFe2O4 (c) CuFe2O4 (d) MgFe2O4 (e) Fe3O4.

Various solvents like toluene, 1,4 dioxane, acetonitrile and water were used for liquid phase oxidation of HMF (Fig. 6). The results show the polar aprotic solvents like 1, 4 dioxane and acetonitrile are better than polar protic solvents like water and ethanol. The solubility of FDCA in solvent used in the reaction is important. It is found that the solubility of FDCA in acetonitrile is more than other solvents. The nonpolar solvents like toluene gave poor yield of FDCA and hence were not suitable for this reaction. We got the highest yield of FDCA in acetonitrile solvent and hence it is used for further reactions.

3.1.3. BET surface area analysis The surface properties of all catalysts were determined by BrunauerEmmett-Teller (BET) method by using Micromeritics ASAP 2000 instrument. The surface area and pore volume of all catalysts are given in Table 1. 3.1.4. Thermo gravimetric analysis (TGA) TGA of MnFe2O4 was performed by using Perkin Elmer Pyris Diamond instrument (Fig. 3). TGA of MnFe2O4 indicates regions of two weight loss steps (Fig. 5). The first mass loss of 6 wt.% (between 34 °C and 150 °C) indicates the removal of moisture from ferrite surface of catalyst. The second loss above 150 °C and below 550 °C (11 wt.%) is due to decomposition of surface bonded polyethylene glycol molecules, which is responsible for the stability of MnFe2O4 in the aqueous medium.

3.4. Catalyst screening Various ferrites such as CuFe2O4, CoFe2O4, MnFe2O4 and MgFe2O4 along with Fe3O4 were screened for oxidation of HMF (Fig. 7). It was observed that MnFe2O4 is more active and gave the highest yield of FDCA. The activity of all these ferrites is dependent on variable valance of metal cations. Mn in MnFe2O4 has more variable valence than other metals and hence shows more activity in FDCA synthesis. In MnFe2O4 catalyst both metals take part in the reaction because transition of Fe2+ to Fe3+ and Mn2+ to Mn3+ generates highly active peroxy radicals in the presence of peroxide such as TBHP, H2O2, etc.

3.1.5. SEM and TEM The SEM image of prepared MnFe2O4 nanoparticles as shown in Fig. 4(a) shows that they are of spherical shape and mostly uniform in size. The EDX of MnFe2O4 shows that samples only contain manganese, iron and oxygen. TEM image reveal that MnFe2O4 is spherical and uniform. Approximately 15–35 nm in range.

3.5. Effect of speed of agitation The agitation speed was varied from 800 to 1200 rpm (Fig. 8). There was no distinct change in the rate of FDCA formation at different speeds which also proved the absence of external mass transfer resistance in reaction. So, further reactions were conducted at 1000 rpm.

3.2. Effect of different oxidants The oxidation of HMF was carried out by using various oxidizing agents like molecular oxygen, hydrogen peroxide (H2O2), and tert-butyl hydroperoxide (TBHP) at 100° C (Fig. 5). The reaction with molecular oxygen at 20 bar of O2 pressure gave lower yield of FDCA. However, the use of peroxides in the oxidation of HMF increases the yield of FDCA significantly. TBHP gave higher yield of FDCA than H2O2; hence it is more suitable oxidant for this reaction. The higher yield of FDCA by using TBHP than H2O2 is due to a more stable radical species of TBHP than H2O2. To confirm that the reaction did not proceed in the absence

3.6. Effect of catalyst loading On the basis of total reaction volume the loading of catalyst was varied in the range of 0.01–0.02 g/cm3 (Fig. 9). It was observed that the initial rate of oxidation of HMF to FDCA is directly proportional to catalyst loading. This is mainly due to proportional increase in number of catalytic active sites available for reaction as the loading of catalyst is increased. Above 0.015 g/cm3 catalyst loading, there was no significant increase in FDCA formation because the number of sites was more than that required for adsorption of reacting species and hence further experiments were carried out at 0.015 g/cm3catalyst loading. Conversion of HMF increases with increase in catalyst loading (Fig. S1) and the plot of initial rate vs catalyst loading shows a linear behaviour which confirms that there is no mass transfer resistance present (Fig. S2).

Table 1 BET surface area and pore volume of MFe2O4. Material 2

BET surface area (m /g) Pore volume (cm3/g)

CuFe2O4

CoFe2O4

MnFe2O4

MgFe2O4

95.49 0.33

81.04 0.33

31.29 0.11

38.31 0.17

121

Catalysis Today 309 (2018) 119–125

A.B. Gawade et al.

Fig. 4. SEM and TEM images of MnFe2O4.

Fig. 5. Effect of different oxidants on yield of FDCA. Reaction conditions: HMF 1 mmol, speed of agitation 1000 rpm, catalyst loading 0.015 g/cm3, oxidant 9 mmol, temperature 100° C, total volume 20 mL, reaction time 300 min. Catalyst: MnFe2O4.

Fig. 7. Effect of different catalysts on yield of FDCA. Reaction conditions: HMF 1 mmol, speed of agitation 1000 rpm, catalyst loading 0.015 g/cm3, TBHP 9 mmol, temperature 100° C, total volume 20 mL, reaction time 300 min.

Fig. 8. Effect of speed of agitation on yield of FDCA. Reaction conditions: HMF 1 mmol, catalyst loading 0.015 g/cm3, TBHP 9 mmol, temperature 100° C, total volume 20 mL, reaction time 300 min. Speed of agitation ( ) 800 rpm, ( ) 1000 rpm, ( ) 1200 rpm,

Fig. 6. Effect of different solvents on the yield of FDCA. Reaction conditions: HMF 1 mmol, speed of agitation 1000 rpm, catalyst loading 0.015 g/cm3, TBHP 9 mmol, temperature 100° C, total volume 20 mL, reaction time 300 min. Catalyst: MnFe2O4.

Catalyst: MnFe2O4.

3.8. Effect of temperature 3.7. Effect of mole ratio of HMF to TBHP The temperature effect was studied from 80° to 110 °C (Fig. 11). The rate of conversion of HMF and formation of FDCA was increased with increase in temperature which indicated that the reaction was kinetically controlled. After 100 °C temperature, there was no significant increase in FDCA formation and so 100 °C was taken as an optimum temperature for this reaction.

The mole ratio of HMF to TBHP was varied in the range of 1:5–1:11 (Fig. 10). It was observed that with increase in concentration of TBHP the rate of formation of FDCA increases. This is mainly due to the generation of more O2 for oxidation on increasing the concentration of TBHP. This helps in faster oxidation of HMF into FDCA. After 1:9 mol ratio, there was no much increase in FDCA formation and so further reactions were carried out at this mole ratio. 122

Catalysis Today 309 (2018) 119–125

A.B. Gawade et al.

Fig. 12. Reusability study. ( ) Conversion of HMF ( ) Selectivity of FDCA. Reaction conditions: HMF 1 mmol, speed of agitation 1000 rpm, catalyst loading 0.015 g/cm3, TBHP 9 mmol, temperature 100° C, total volume 20 mL, reaction time 300 min. Catalyst: MnFe2O4.

Fig. 9. Effect of catalyst loading on yield of FDCA. Reaction conditions: HMF 1 mmol, TBHP 9 mmol, speed of agitation 1000 rpm, temperature 100° C,total volume 20 mL, reaction time 300 min. Catalyst loading ( ) 0.01 g/cm3, ( ) 0.015 g/cm3, ( ) 0.02 g/ cm3, Catalyst: MnFe2O4.

again it was separated and dried at 120° C for 3 h. The loss in the catalyst amount obtained after filtration and drying (∼4–5%) was made up by adding virgin catalyst. The catalyst activity was practically the same in all experiments. Also when the reused catalyst was used without make-up but the volume of reaction mass was reduced to get the same catalyst loading, similar rate and conversion were obtained. Thus, the catalyst was robust, active and reusable up to 4 cycles. We have also carried out the ICP-AES analysis of reaction mass and it was found that leaching of manganese is below detection level. 4. Reaction mechanism and kinetic model The reaction for conversion of HMF to FDCA can be divided into five sub-reactions which take place in parallel and series as shown in Scheme 1. Fig. 10. Effect of mole ratio of HMF to TBHP. Reaction conditions: HMF 1 mmol, speed of agitation 1000 rpm, catalyst loading 0.015 g/cm3, temperature 100° C,total volume 20 mL, reaction time 300 min. HMF to TBHP mole ratio ( ) 1:5 mmol, ( ) 1:7 mmol, ( ) 1:9 mmol, ( ) 1:11 mmol. Catalyst: MnFe2O4.

k1

A+F→ B k2

A+F→ C k3

B+F→ D k4

C+F→ D k5

D+F→ E

(1) (2) (3) (4) (5)

Where, A – HMF B – DFF C – HMFCA D – FFCA E – FDCA F – TBHP The rate of formation of the species A–E can be written as,

−dCA = (k1 + k2) CA CF w dt

(6)

Fig. 11. Effect of Temperature. Reaction conditions: HMF 1 mmol, speed of agitation 1000 rpm, catalyst loading 0.015 g/cm3, TBHP 9 mmol, total volume 20 mL, reaction time 300 min. Temperature ( ) 80° C, ( ) 90° C, ( ) 100° C, ( ) 110° C. Catalyst:

dCB = k1 CA CF w − k3 CB CF w dt

(7)

MnFe2O4.

dCC = k2 CA CF w − k 4 CC CF w dt

(8)

dCD = k3 CB CF w + k 4 CC CF w − k5 CD CF w dt

(9)

3.9. Reusability of catalyst The catalyst reusability study of MnFe2O4was performed for 4 runs (Fig. 12). After completion of each reaction, the catalyst was separated by using external magnet. Then to remove any adsorbed material from surface of catalyst it was refluxed in 50 mL acetonitrile for 3 h. Then

dCE = k5 CD CF w dt

(10)

To evaluate the values of the reaction rate constants for all the 123

Catalysis Today 309 (2018) 119–125

A.B. Gawade et al.

Scheme 1. Oxidation of HMF to FDCA using MnFe2O4.

Table 2 Reaction rate constants for various steps. Sr. No

T (K)

k1 × 103 (L2 mol−1 g−1 s−1)

k2 × 103 (L2 mol−1 g−1 s−1)

k3 × 103 (L2 mol−1 g−1 s−1)

k4 × 103 (L2 mol−1 g−1 s−1)

k5 × 103 (L2 mol−1 g−1 s−1)

1. 2. 3. 4.

353 363 373 383

0.90 1.34 1.98 2.21

1.12 1.77 2.20 2.54

2.12 2.58 3.24 4.32

1.35 2.09 2.62 3.20

2.46 3.83 4.53 5.37

Bose National Fellowship of DST-GOI.

Table 3 Activation Energy for various steps. E1 E2 E3 E4 E5

Appendix A. Supplementary data

7.02 kcal/mol 6.08 kcal/mol 5.33 kcal/mol 6.40 kcal/mol 5.71 kcal/mol

Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.cattod.2017.08.061. References

above reactions, we need to develop an algorithm to extract these values from the available reaction rates. The algorithm is discussed below. From Eq. (10), we can obtain values of rate constant k5. Using this value of k5, we can then evaluate values of reaction rate constants k3 and k4 by solving Eq. (9). This is done using Polymath. The value of k3 can then be used to obtain the value of k1 by solving Eq. (7). Using k1, we can evaluate the value of k2 by solving Eq. (6). This loop can be verified by comparing the value of k2 obtained by solving Eq. (6) and its value obtained by solving Eq. (8). The values for all the reaction rate constants are presented in Table 2. The reaction rate constants of all steps show that the oxidation of alcohol to aldehyde is faster than oxidation of aldehyde to acid. From the values of reaction rate constant the activation energy for each step was calculated (Table 3).

[1] A. Corma Canos, S. Iborra, A. Velty, Chemical routes for the transformation of biomass into chemicals, Chem. Rev. 107 (2007) 2411–2502, http://dx.doi.org/10. 1021/cr050989d. [2] B. Kamm, Production of platform chemicals and synthesis gas from biomass, Angew. Chem. Int. Ed. 46 (2007) 5056–5058, http://dx.doi.org/10.1002/anie. 200604514. [3] C. Okkerse, H. van Bekkum, From fossil to green, Green Chem. 1 (1999) 107–114, http://dx.doi.org/10.1039/a809539f. [4] X. Qi, M. Watanabe, T.M. Aida, R.L. Smith Jr., Catalytic dehydration of fructose into 5-hydroxymethylfurfural by ion-exchange resin in mixed-aqueous system by microwave heating, Green Chem. 10 (2008) 799, http://dx.doi.org/10.1039/ b801641k. [5] G. Yong, Y. Zhang, J.Y. Ying, Efficient catalytic system for the selective production of 5-hydroxymethylfurfural from glucose and fructose, Angew. Chem. Int. Ed. Engl. 47 (2008) 9345–9348, http://dx.doi.org/10.1002/anie.200803207. [6] H. Ait Rass, N. Essayem, M. Besson, Selective aqueous phase oxidation of 5-hydroxymethylfurfural to 2, 5-furandicarboxylic acid over Pt/C catalysts: influence of the base and effect of bismuth promotion, Green Chem. 15 (2013) 2240, http://dx. doi.org/10.1039/c3gc40727f. [7] P. Gallezot, Conversion of biomass to selected chemical products, Chem. Soc. Rev. 41 (2012) 1538–1558, http://dx.doi.org/10.1039/c1cs15147a. [8] T. Thananatthanachon, T.B. Rauchfuss, Efficient production of the liquid fuel 2, 5dimethylfuran from fructose using formic acid as a reagent, Angew. Chem. Int. Ed. 49 (2010) 6616–6618, http://dx.doi.org/10.1002/anie.201002267. [9] A.B. Gawade, M.S. Tiwari, G.D. Yadav, Biobased green process: selective hydrogenation of 5-hydroxymethylfurfural to 2, 5-dimethyl furan under mild conditions using Pd-Cs 2.5 H 0.5 PW 12 O 40/K-10 clay, ACS Sustain. Chem. Eng. 4 (2016) 4113–4123, http://dx.doi.org/10.1021/acssuschemeng.6b00426. [10] C. Méalares, A. Gandini, Polymeric schiff bases bearing furan moieties 2. polyazines and polyazomethines, Polym. Int. 40 (1996) 33–39, http://dx.doi.org/10.1002/ (SICI)1097-0126(199605)40:1<33::AID-PI512>3.0.CO;2-S. [11] B. Saha, S. Dutta, M.M. Abu-Omar, Aerobic oxidation of 5-hydroxylmethylfurfural with homogeneous and nanoparticulate catalysts, Catal. Sci. Technol. 2 (2012) 79–81, http://dx.doi.org/10.1039/C1CY00321F. [12] C. Triebl, V. Nikolakis, M. Ierapetritou, Simulation and economic analysis of 5hydroxymethylfurfural conversion to 2, 5-furandicarboxylic acid, Comput. Chem. Eng. 52 (2013) 26–34. [13] N.K. Gupta, S. Nishimura, A. Takagaki, K. Ebitani, Hydrotalcite-supported goldnanoparticle-catalyzed highly efficient base-free aqueous oxidation of 5-hydroxymethylfurfural into 2,5-furandicarboxylic acid under atmospheric oxygen pressure, Green Chem. 13 (2011) 824, http://dx.doi.org/10.1039/c0gc00911c. [14] A.J.J.E. Eerhart, A.P.C. Faaij, M.K. Patel, Replacing fossil based PET with biobased PEF; process analysis, energy and GHG balance, Energy Environ. Sci. 5 (2012) 6407, http://dx.doi.org/10.1039/c2ee02480b. [15] J.J. Bozell, G.R. Petersen, Technology development for the production of biobased products from biorefinery carbohydrates—the US Department of Energy’s Top 10 revisited, Green Chem. 12 (2010) 539, http://dx.doi.org/10.1039/b922014c. [16] W. Partenheimer, V.V. Grushin, Synthesis of 2,5-diformylfuran and furan-2,5-

5. Conclusions The oxidation of HMF to FDCA was successfully achieved with good conversion and high selectivity by using MnFe2O4 catalyst and TBHP as an oxidant. The reaction requires less time and energy compared to other reported methods for production of FDCA. The process is green as no homogenous base was used in reaction. The concentration of all intermediates observed in this reaction was determined and kinetic data was obtained. The activation energy for each step was calculated by plotting Arrhenius plot. Various reaction parameters were studied and kinetic model successfully developed. The catalyst was characterized by different analytical techniques such as XRD, SEM, TEM, FT-IR, BETASAP, and TGA. The catalyst is highly stable and active towards the desired product and recycled four runs without loss of its activity and selectivity. Acknowledgements A.B.G. thanks University Grants Commission (UGC) for awarding the Senior Research Fellowship under its Green Technology special meritorious fellowship program. A.V.N. acknowledges CSIR for awarding the Senior Research Fellowship. G.D.Y. acknowledges support from R.T. Mody Distinguished Professor Endowment of ICT and J.C. 124

Catalysis Today 309 (2018) 119–125

A.B. Gawade et al.

[17]

[18]

[19]

[20]

[21]

[22]

[23]

[24]

[25]

[26] F. Neațu, R.S. Marin, M. Florea, N. Petrea, O.D. Pavel, V.I. Pârvulescu, Selective oxidation of 5-hydroxymethyl furfural over non-precious metal heterogeneous catalysts, Appl. Catal. B Environ. 180 (2016) 751–757, http://dx.doi.org/10.1016/ j.apcatb.2015.07.043. [27] M.L. Ribeiro, U. Schuchardt, Cooperative effect of cobalt acetylacetonate and silica in the catalytic cyclization and oxidation of fructose to 2,5-furandicarboxylic acid, Catal. Commun. 4 (2003) 83–86, http://dx.doi.org/10.1016/s1566-7367(02) 00261-3. [28] Z. Zhang, K. Deng, Recent advances in the catalytic synthesis of 2, 5-furandicarboxylic acid and its derivatives, ACS Catal. 5 (2015) 6529–6544, http://dx. doi.org/10.1021/acscatal.5b01491. [29] A.S. Burange, S.R. Kale, R. Zboril, M.B. Gawande, R.V. Jayaram, Magnetically retrievable MFe2O4 spinel (M = Mn, Co, Cu, Ni, Zn) catalysts for oxidation of benzylic alcohols to carbonyls, RSC Adv. 4 (2014) 6597, http://dx.doi.org/10. 1039/c3ra45327h. [30] K. Sreekumar, T. Mathew, B.M. Devassy, R. Rajgopal, R. Vetrivel, B.S. Rao, Vaporphase methylation of pyridine with methanol to 3-picoline over Zn1-xCoxFe2O4 (x = 0, 0.2, 0.5, 0.8 and 1.0)-type ternary spinels prepared via a low temperature method, Appl. Catal. A Gen. 205 (2001) 11–18, http://dx.doi.org/10.1016/s0926860x(00)00542-1. [31] J. Feng, L. Su, Y. Ma, C. Ren, Q. Guo, X. Chen, CuFe2O4 magnetic nanoparticles: a simple and efficient catalyst for the reduction of nitrophenol, Chem. Eng. J. 221 (2013) 16–24, http://dx.doi.org/10.1016/j.cej.2013.02.009. [32] S. Sun, H. Zeng, D.B. Robinson, S. Raoux, P.M. Rice, S.X. Wang, G. Li, Monodisperse MFe 2 O 4 (M = Fe, Co, Mn) nanoparticles, J. Am. Chem. Soc. 126 (2004) 273–279, http://dx.doi.org/10.1021/ja0380852. [33] J. Młochowski, H. Wójtowicz-Młochowska, Developments in synthetic application of selenium(IV) oxide and organoselenium compounds as oxygen donors and oxygen-transfer agents, Molecules 20 (2015) 10205–10243, http://dx.doi.org/10. 3390/molecules200610205. [34] A.V. Nakhate, G.D. Yadav, Hydrothermal synthesis of CuFe 2 O 4 magnetic nanoparticles as active and robust catalyst for N- arylation of indole and imidazole with aryl halide, Chem. Sel. 2 (2017) 2395–2405, http://dx.doi.org/10.1002/slct. 201601846. [35] A.V. Nakhate, G.D. Yadav, Palladium nanoparticles supported carbon based graphene oxide monolith as catalyst for sonogashira coupling and hydrogenation of nitrobenzene and alkenes, Chem. Sel. 1 (2016) 3954–3965, http://dx.doi.org/10. 1002/slct.201600819.

dicarboxylic acid by catalytic air-oxidation of 5-hydroxymethylfurfural. Unexpectedly selective aerobic oxidation of benzyl alcohol to benzaldehyde with metal/bromide catalysts, Adv. Synth. Catal. 343 (2001) 102–111. P. Vinke, W. van der Poel, H. van Bekkum, On the oxygen tolerance of noble metal catalysts in liquid phase alcohol oxidations the influence of the support on catalyst deactivation, Stud. Surf. Sci. Catal. (1991) 385–394, http://dx.doi.org/10.1016/ S0167-2991(08)61145-3. S. Siankevich, G. Savoglidis, Z. Fei, G. Laurenczy, D.T.L. Alexander, N. Yan, P.J. Dyson, A novel platinum nanocatalyst for the oxidation of 5-hydroxymethylfurfural into 2,5-furandicarboxylic acid under mild conditions, J. Catal. 315 (2014) 67–74, http://dx.doi.org/10.1016/j.jcat.2014.04.011. Y.Y. Gorbanev, S.K. Klitgaard, J.M. Woodley, C.H. Christensen, A. Riisager, Goldcatalyzed aerobic oxidation of 5-hydroxymethylfurfural in water at ambient temperature, ChemSusChem 2 (2009) 672–675, http://dx.doi.org/10.1002/cssc. 200900059. S. Albonetti, T. Pasini, A. Lolli, M. Blosi, M. Piccinini, N. Dimitratos, J.A. LopezSanchez, D.J. Morgan, A.F. Carley, G.J. Hutchings, F. Cavani, Selective oxidation of 5-hydroxymethyl-2-furfural over TiO2-supported gold–copper catalysts prepared from preformed nanoparticles: effect of Au/Cu ratio, Catal. Today 195 (2012) 120–126, http://dx.doi.org/10.1016/j.cattod.2012.05.039. X. Wan, C. Zhou, J. Chen, W. Deng, Q. Zhang, Y. Yang, Y. Wang, Base-free aerobic oxidation of 5- hydroxymethyl-furfural to 2, 5- furandicarboxylic acid in water catalyzed by functionalized carbon nanotube-supported Au-Pd alloy nanoparticles, ACS Catal. 4 (2014) 2175–2185, http://dx.doi.org/10.1021/cs5003096. S.E. Davis, A.D. Benavidez, R.W. Gosselink, J.H. Bitter, K.P. de Jong, A.K. Datye, R.J. Davis, Kinetics and mechanism of 5-hydroxymethylfurfural oxidation and their implications for catalyst development, J. Mol. Catal. A Chem. 388–389 (2014) 123–132, http://dx.doi.org/10.1016/j.molcata.2013.09.013. Y.Y. Gorbanev, S. Kegnæs, A. Riisager, Effect of support in heterogeneous ruthenium catalysts used for the selective aerobic oxidation of HMF in water, Top. Catal. 54 (2011) 1318–1324, http://dx.doi.org/10.1007/s11244-011-9754-2. B. Saha, D. Gupta, M.M. Abu-Omar, A. Modak, A. Bhaumik, Porphyrin-based porous organic polymer-supported iron(III) catalyst for efficient aerobic oxidation of 5hydroxymethyl-furfural into 2,5-furandicarboxylic acid, J. Catal. 299 (2013) 316–320, http://dx.doi.org/10.1016/j.jcat.2012.12.024. T.S. Hansen, I. Sádaba, E.J. García-Suárez, A. Riisager, Cu catalyzed oxidation of 5hydroxymethylfurfural to 2,5-diformylfuran and 2,5-furandicarboxylic acid under benign reaction conditions, Appl. Catal. A Gen. 456 (2013) 44–50, http://dx.doi. org/10.1016/j.apcata.2013.01.042.

125