Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Fuel 290 (2021) 119947

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Production of biofuel 2,5-dimethylfuran using highly efficient single-step


selective hydrogenation of 5-hydroxymethylfurfural over novel Pd-Co/
Al-Zr mixed oxide catalyst
Devendra S. Pisal, Ganapati D. Yadav *
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: Sustainable and cleaner production of liquid biofuel 2,5-dimethylfuran (DMF) is in great demand. Because of its
2,5-Dimethylfuran water insolubility, high boiling point, and high energy density, DMF seems to be a promising liquid biofuel.
5‑Hydroxymethylfurfural (HMF) Selective hydrogenation of biomass-derived 5-hydroxymethylfurfural (HMF) to 2,5-DMF can be achieved using
Biofuel
novel catalysts and is presented here. In this study, a series of Al-Zr mixed oxide catalysts (AZMO) were prepared
Hydrogenation
by co-precipitation and hydrothermal techniques. AZMO catalyst prepared by the co-precipitation method gave
Sustainability
Green Chemistry and Engineering higher acidity and surface area. Furthermore, mono and bimetallic M/AZMOCP (M = Ni, Co, Cu, and Pd) catalysts
were developed by the impregnation and used for hydrogenation of HMF. The catalytic activity was found in the
order of: 2%Pd-5%Co/AZMOCP > 1%Pd-5%Co/AZMOCP > 2%Pd-3%Co/AZMOCP > 1%Pd-3%Co/AZMOCP > 5%
Ni-5%Co/AZMOCP > 1%Pd/AZMOCP > 5%Co/AZMOCP > 5%Cu-5%Co/AZMOCP > 3%Co/AZMOCP. It was found
that 2%Pd-5%Co/AZMOCP resulted in complete conversion of HMF with 97% yield of DMF at 100 ◦ C and 10 atm
H2 pressure. The catalysts were characterized by sophisticated techniques such as FESEM, EDS, XRD, NH3-TPD,
ATR-FTIR, BET analysis, HRTEM, TPR, XPS and TGA-DSC. The reaction mechanism was developed based on
dual-site Langmuir-Hinshelwood-Hougen-Watson (LHHW) model. The catalyst was highly selective and reusable.
The overall process is clean and green.

instance, Wu et al. reported excellent review on different MOF-derived


catalysts and their use for valorization of sugars into furans and alco­
1. Introduction
hols and hydrogenation of furans into different fine chemicals [7].
Similarly cellulose based materials showed major use in energy storage,
Fossil fuels, including coal, natural gas, and petrol are required for
which are used in electrode materials [8]. 2,5-Furandicarboxylic acid
various operations in chemical industries, transportation, and energy
(FDCA) was produced from HMF with 95% yield over Au-Pd alloy
sector [1]. Day-by-day these non-renewable sources are depleting and
nanoparticle encapsulated in metal organic framework (MOF) [9].
creating serious environmental concerns such as pollution and high
Majorly, 5‑Hydroxymethylfurfural (HMF) is considered to be a platform
carbon emission. Hence, a bio-based safe alternative is needed to fulfill
molecule, which is a renewable and sustainable building block for a
the demand of the fuel sector, which can be sustainable, low cost, and
variety of fuel compounds. For instance, 2,5-bis-(hydroxymethyl)furan
environmentally friendly alternative. 2,5-Dimethylfuran (DMF) is an
(BHMF), [10] 2,5-dimethylfuran (DMF), [11,12] 5-methylfurfural
excellent choice of fuel as it has a very high octane number (RON 119),
(MFF),[13] 5-methylfurfuryl alcohol (MFA), 2,5-dimethyl tetrahydro­
high energy content (30 kJ cm− 3) than ethanol (23 kJ cm− 3) [2,3], it can
furan (DMTHF), 2,5-dihydroxy methyl tetrahydrofuran (DHMTHF),
be easily blended with gasoline [4], and has unique properties including
1,2,6-hexanetriol (HTO), 1,6-hexanediol (HDO), and 1-hydroxyhexane-
high boiling point and water insolubility [5]. Most importantly, DMF is
2,5-dione (HHD) [13] all are value-added products of HMF. Considering
easily prepared by hydrogenation of bio-based feedstock, 5-hydroxyme­
all these value added products and their extensive importance as a
thylfurfural (HMF) [6].
feedstock, HMF hydrogenation has been studied extensively in recent
In recent years, several groups have reported synthesis of value-
years by various research groups. Catalytic transfer hydrogenation was
added products using lignocellulosic biomass as feedstock; for

Abbreviations: HMF, 5‑Hydroxymethylfurfural; MFF, 5-Methylfurfural; MFA, 5-Methylfurfuryl alcohol; DMF, 2,5-Dimethylfuran.
* Corresponding author.
E-mail address: gd.yadav@ictmumbai.edu.in (G.D. Yadav).

https://doi.org/10.1016/j.fuel.2020.119947
Received 16 May 2020; Received in revised form 4 November 2020; Accepted 2 December 2020
Available online 9 January 2021
0016-2361/© 2020 Elsevier Ltd. All rights reserved.
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

accessibility of the catalyst due to the use of such support [27]. Whereas
Nomenclature mixed oxide catalysts have many advantages such as activity and sta­
bility over a wide range of temperatures, being reusable and inexpen­
A HMF sive, etc. [28]. Mixed oxide catalysts containing two or more metals
B MFF were prepared by various techniques such as sol–gel [29,30] co-
D MFA precipitation method [28], impregnation [31], physical mixing [32],
E DMF solution combustion method [33], etc. Different methods result in
W Water variation in structure, morphology, and hence the activity of the cata­
CA Concentration of A (mol/cm3) lyst. Among reported oxides, Al2O3-ZrO2 mixed oxide possesses tunable
CD Concentration of D (mol/cm3) acidity and basicity, and because of this characteristic property, it was
CE Concentration of E (mol/cm3) explored to use as a catalyst or as support in various reactions such as
pH2 Partial pressure of hydrogen (atm) oxidation [34], hydrotreatment, hydrogenation [35], dehydrogenation
K Adsorption constant of different species (cm3/mol) [36], and transfer hydrogenation, etc. [37].
k Rate constant (cm3 g-1mol− 1 min− 1) Reports are available on the selective hydrogenation of –C=O group
KH Adsorption constant for hydrogen (cm3/mol) using a combination of bimetallic catalysts such as Pt-Pd and Fe-Pd
r Rate of reaction (mol cm− 3 min− 1) [38,21]. The metal catalysts containing a combination of noble metal
Sa Acid sites and non-noble metal are favored due to economic as well as engineering
Sm Metal sites aspects, and it also such a catalyst effectively carries out selective hy­
CTa Total concentration of acid site (mol/cm3) drogenation. Co-based and Pd-based catalysts were also reported for
CTm Total concentration of metal site (mol/cm3) –C=O hydrogenation reactions, which resulted in higher yield [39,40].
w Catalyst loading (g/cm3) Therefore, design and synthesis of a suitable catalyst with eco-friendly
methodologies has become a prime requirement in HMF hydrogena­
tion. Mixed oxide support Al-Zr with higher acidity and surface area in
combination with bimetallic catalysts was chosen in this work for a
studied using NiCo/C in combination with formic acid as a hydrogen single-step hydrogenation reaction at mild conditions.
donor to obtain 90% yield of DMF at 210 ◦ C [14]. Various mono- and bi- In the present investigation, we have prepared a series of Al-Zr mixed
metallic combinations were used in previous work; for instance, Zu et al. oxides (AZMO) by co-precipitation and hydrothermal methods. All the
used Ru/Co3O4 catalyst and reported 93.4% yield of DMF at 130 ◦ C and prepared catalysts were characterized in detail for comprehending the
7 atm H2 pressure in 24 h [6]. Pd/Zr-based metal–organic framework relation for choosing the best catalyst for hydrogenation of HMF. A
(MOF)-sulfonated graphene oxide catalyst was used at 10 atm H2 pres­ novel catalyst is thus proposed for the selective hydrogenation of HMF in
sure and 160 ◦ C, which resulted in DMF yield of 70.5% [15]. Some re­ one-pot, to get the liquid biofuel DMF at mild reaction conditions.
actions involved use of either harsh reactants or reaction conditions; for
instance, bimetallic Pd-Au catalyst was used in the presence of corrosive 2. Experimental
hydrochloric acid to convert HMF to DMF [16]. About 98.7% yield of
DMF was reported using Ni-Cu3NC catalyst at 33 atm and 180 ◦ C [17]. 2.1. Reagents
Raney Ni catalyst was used for HMF hydrogenation, which gave com­
plete conversion of HMF with 88.5% yield of DMF at 180 ◦ C and 15 atm Following chemicals were procured from reputed vendors: HMF (SRL
H2 in 15 h [18]. Similarly, operating conditions of 180 ◦ C, 15 atm H2 Pvt. Ltd, Mumbai, India); Ni(NO3)2⋅6H2O, Co(NO3)2⋅6H2O, Al
pressure and 15 h reaction time were reported by Kong et al. to obtain (NO3)3⋅9H2O, and urea (S. D. Fine Chemicals, Mumbai); Cu(NO3)2⋅3H2O
93.6% yield of DMF using NiZnAl hydrotalcite [19]. Use of Pd improved (AVRA synthesis Pvt. Ltd., Hyderabad); NaOH (Thomas Baker, Mum­
the yield of DMF in some cases; for example, hydrogenation in contin­ bai); zirconyl nitrate (Loba Chemie Pvt. Ltd, Mumbai); methanol, Pd
uous flow reactor was done using 10Cu-1Pd/RGO (reduced graphene (NO3)⋅3H2O, acetonitrile, and tetrahydrofuran (THF) (Merck, Mumbai).
oxide) resulted in 96% conversion of HMF and 95% yield of DMF [20].
Recently, a magnetically separable bimetallic Fe-Pd/C catalyst was 2.2. Catalyst synthesis
developed for HMF hydrogenation, which reported complete conversion
with 85% yield of DMF [21]. Cu-Pd bimetallic nanoparticles in carbon 2.2.1. Co-precipitation method
matrix catalyst reported 96.5% yield using THF at 15 atm H2 and 120 ◦ C At first, zirconyl nitrate hydrate and aluminum nitrate nonahydrate
in 7 h [22]. Low cost Co-based catalyst was also reported largely for its (Al/Zr mole ratio of 1:1) were dissolved separately in two separate
activity and stability; for example, a Co-based catalysts, 11.8%Co-(ZnO- beakers containing 50 mL distilled water (DW). Next, a solution of 35
ZnAl2O4) [23] was used to report 74.2% yield of DMF in 24 h at 130 ◦ C. mmol of sodium hydroxide was prepared and added dropwise along
Similarly, Cu-Co/Al2O3 catalyst, used at high temperature and pressure with aluminum nitrate solution into the zirconium nitrate solution. The
of 220 ◦ C and 30 atm, respectively, resulted in 76% yield of DMF in 8 h addition was done simultaneously with constant stirring at 40 ◦ C. The
[24]. DMF yield of 99% was reported at 20 atm and 170 ◦ C in 4 h over rate of addition was adjusted in such a way that the pH of the solution
3%Pd/C [25]. It was also reported that a catalyst with higher acidity was maintained between 9 and 10. After completion of the addition, the
resulted in high yield of DMF; for instance, a highly acidic Pd based slurry was stirred for 3 h, followed by aging at 70 ◦ C for 12 h. The ob­
catalyst resulted in 98% conversion of HMF with 81% yield at 10 atm tained precipitate was filtered, washed several times with DW to achieve
[26]. Therefore, it apprears that most of the processes discussed above, neutral pH, dried at 120 ◦ C for 12 h, and calcined at 600 ◦ C for 5 h. This
used harsh reaction conditions and were time consuming. Hence, to was labeled as AZMOCP. Similarly, AZMOCP2, AZMOCP3 (Al/Zr mole
obtain high yield of DMF at mild reaction condition remained a chal­ ratio of 2:1 and 3:1, respectively), Al2O3, and ZrO2 were prepared.
lenging task and was highly desirable. There is scope for studying Various metals (M = Co, Ni, Cu, and Pd, etc.) were impregnated, dried at
combination of bimetallic Pd and Co catalysts for their activity in HMF 120 ◦ C for 12 h, and calcined at 550 ◦ C for 3 h to obtain respective x wt.
hydrogenation to DMF. %M/AZMOCP catalyst (where x = 1, 2, 3, and 5).
One of the most essential and extensively used categories of het­
erogeneous catalysts is mixed metal oxides. To increase the surface area 2.2.2. Hydrothermal method
and acidity, usually, heteropolyacids are supported on activated carbon Zirconyl nitrate hydrate and aluminum nitrate nonahydrate (Al/Zr
matrix or silica support; conversely, this limits the efficiency and mole ratio of 1:1) solutions were prepared separately in 50 mL DW and

2
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Table 1
BET surface area, BJH adsorption total pore volume, and BJH adsorption
average pore diameter of different catalysts.
# Catalyst Al/Zr BET BJH adsorption BJH adsorption
Mole Surface total pore Avg. pore
ratio area (m2/ volume (cm3/g) diameter (nm)
g)

1) Al2O3 – 207 0.48 7


2) ZrO2 – 117.2 0.18 4.1
3) AZMOCP 1:1 193 0.30 6.2
4) AZMOCP2 2:1 226 0.35 6.9
5) AZMOCP3 3:1 257 0.52 8.1
6) AZMOHT 1:1 90 0.12 3
7) fresh 2%Pd- 1:1 168 0.26 6.2
5%Co/
AZMOCP
8) reused 2% 1:1 163 0.25 5.7
Pd-5%Co/
AZMOCP

Table 2
NH3-TPD of various synthesized catalysts.

Fig. 1. N2 adsorption–desorption isotherm of a) AZMOHT, b) Al2O3, c) ZrO2, d) # Catalyst Al/Zr mole NH3-TPD Acidity (mmol/gcat)
ratio
AZMOCP, e) fresh 2%Pd-5%Co/AZMOCP, f) reused 2%Pd-5%Co/AZMOCP, g) Weak Moderate/ Total
AZMOCP2, and h) AZMOCP3. Strong

1 Al2O3 – 0.49 0.24 0.73


mixed with stirring for 30 min. The precipitating agent urea (0.09 mol) 2 ZrO2 – 0.38 0.2 0.58
was dissolved in 20 mL DW and added dropwise followed by stirring for 3 AZMOCP 1:1 1.8 1.81 3.61
4 AZMOHT 1:1 0.59 0.012 0.602
another 30 min. This solution was put into a Teflon lined bomb reactor
5 AZMOCP2 2:1 0.31 0.29 0.60
and kept at 150 ◦ C for 24 h. The obtained mass was filtered, washed 6 AZMOCP3 3:1 0.63 0.012 0.642
several times until pH 7, and dried at 120 ◦ C for 12 h. The calcination 7 2%Pd-5%Co/AZMOCP 1:1 0.44 0.8 1.24
was done at 600 ◦ C for 5 h. The obtained catalyst was labeled as 8 reused 2%Pd-5%Co/ 1:1 0.44 0.79 1.23
AZMOCP
AZMOHT.

2.3. Hydrogenation reaction setup and products analysis

All HMF hydrogenation reactions were investigated in a 50 mL high-


pressure standard batch reactor (Autoclave Engineers, USA). At first, a
specific amount of catalyst was charged in the autoclave and was
reduced by H2 pressure of 30 atm, at 200 ◦ C for 2 h. After cooling the
reactor, a known amount of HMF and solvent THF were taken into the
reactor. The reactor was flushed three times with hydrogen, and after
reaching the desired temperature, it was pressurized with hydrogen to
perform the reactions. The collected samples were analyzed by GC-FID
(TRACE 1110, Thermo Scientific) equipped with TG-5MS capillary col­
umn (0.25 mm diameter and 30 m length) and FID, as well as by using
HPLC (Agilent 1260 infinity) equipped with CHEMSIL C18 column (4.6
mm I.D × 250 mm length × 5 µm) and UV detector (260 nm). Aceto­
nitrile and DI water (60:40) were employed with 1 mL/min flow rate as
the mobile phase. The products were identified through HR-MS (SI,
Figs. S1–S3) and NMR study (SI, Figs. S4 and S5).

3. Result and discussion

3.1. Characterization of catalyst


Fig. 2. NH3-TPD of a) ZrO2, b) AZMOHT, c) Al2O3, d) AZMOCP2, e) AZMOCP, f)
AZMOCP3, g) fresh 2%Pd-5%Co/AZMOCP2, and h) reused 2%Pd-5%
3.1.1. BET
Co/AZMOCP2.
According to IUPAC classification, the N2 adsorption–desorption
isotherm obtained with a typical hysteresis loop belongs to type IV being
size of zirconia. AZMOHT has the lowest surface area, 90 m2/g, which
mesoporous in nature [41]. Fig. 1 shows N2 adsorption–desorption
was in direct proportion to the narrowing of the hysteresis loop (Fig. 1a).
isotherms of different catalysts, all of them are mesoporous, and relative
As the Al content increased (Table 1, entries 3–5), the surface area
pressure (P/P0) are in the range of 0 to 1. The relative pressure (P/P0) of
increased significantly. The surface area of different catalysts was found
enunciation points was related to the pore diameter of the mesoporous
in the order of: AZMOCP3 (max) > AZMOCP2 > Al2O3 > AZMOCP > fresh
material and sharpness at the end of the loops is attributed to the uni­
2%Pd-5%Co/AZMOCP > reused 2%Pd-5%Co/AZMOCP > ZrO2 >
formity in pore size distribution (Fig. 1c-h). ZrO2 showed surface area
AZMOHT.
(117.2) comparatively lesser than Al2O3 (207) probably because of large

3
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 3. XRD of a) ZrO2, b) AZMOHT, c) AZMOCP, d) fresh 2%Pd-5%Co/AZMOCP,


Fig. 4. ATR-FTIR of a) AZMOCP, b) AZMOHT, c) fresh 2%Pd-5%Co/AZMOCP,
e) reused 2%Pd-5%Co/AZMOCP, and f) Al2O3.
and d) reused 2%Pd-5%Co/AZMOCP.

The trivial change in the surface area and pore volume of reused 2%
Pd-5%Co/AZMOCP when compared with that of fresh 2%Pd-5%Co/
AZMOCP, represents structural integrity of the catalyst.

3.1.2. NH3-TPD
The acidity of all the catalysts was measured by using the NH3-TPD
technique. For comparison, the total acidity of Al2O3 and ZrO2 was
measured (Table 2 and Fig. 2). The measured peak intensities for various
catalysts, which were divided based on the peak obtained at a particular
temperature and represent corresponding acidic sites; namely, weak,
moderate, and strong, respectively. As compared to ZrO2, Al2O3 showed
a wide and intense peak showing the presence of a higher amount of
weak, moderate as well as strong acidic sites. The total acidity of ZrO2
was found to be 0.56 mmol/gcat and that of Al2O3 as 0.73 mmol/gcat.
However, both Al and Zr gave a synergistic effect to boost the total
acidity of AZMOCP to be the highest among all other catalysts, i.e., 3.61
mmol/gcat, which was also confirmed by sharp signals at 150 and 350 ◦ C
(Fig. 2e). There was no noteworthy difference in the total acidity of
AZMOCP2 and AZMOCP3; however, AZMOHT showed a much lesser value Fig. 5. TPR of a) fresh 2%Pd-5%Co/AZMOCP and b) reused 2%Pd-5%
of total acidity than that of AZMOCP. In the case of 2%Pd-5%Co/ Co/AZMOCP.
AZMOCP, after doping of Co and Pd, the total acidity was reduced to
0.52 mmol/gcat, and a marginal change in the total acidity of reused 2% e) was found to be overlapped with that of Al2O3 peak at ~ 46.1◦
Pd-5%Co/AZMOCP showing that the active sites were intact even after (Fig. 3c), which might have resulted due to wide dispersion of PdO over
reuse, and catalyst remained active. the surface of AZMO. The crystallite size of Al2O3, ZrO2, AZMOCP and
2%Pd-5%Co/AZMOCP catalysts was calculated using Scherrer equation
3.1.3. XRD as 9.2, 8.8, 7.5, and 5.5 nm, respectively. It is worth noting that AZMOCP
The XRD patterns of Al2O3, ZrO2, AZMOCP, AZMOHT, fresh 2%Pd-5% catalyst showed crystallite size smaller (7.5 nm) than that of individual
Co/AZMOCP, and reused 2%Pd-5%Co/AZMOCP are shown in Fig. 3. All oxides Al2O3 and ZrO2. Moreover, the crystallite size of 2%Pd-5%Co/
AZMO catalysts except AZMOHT showed the presence of both Al2O3 and AZMOCP catalyst further decreased to 5.5 nm because of the impreg­
ZrO2 oxides. In the case of AZMOHT, due to the hydrothermal techniques nation of Pd and Co and filling of the pores of AZMOCP.
of preparation, alumina and zirconia oxide existed in amorphous form,
and hence, only two broad peaks were observed at 2θ of 32.6◦ and 40- 3.1.4. ATR-FTIR
65◦ . Both Al2O3 and ZrO2 prepared by co-precipitation techniques were The comprehensive study of stretching and bending vibrations of
as per JCPDS # 01–086-1410 (I) and 01–072-2742 (A), respectively. The various functional groups present in the catalysts was done by ATR-
peaks of ZrO2 were found at 2θ value of 30.21 (111), 35.46 (200), 50.55 FTIR. All samples showed the presence of Al-O and Zr-O bands. The
(220), and 60.3 (311). Similarly, for Al2O3 values of 2θ were at 14.7 broad absorption band obtained between 700 and 950 cm− 1 corre­
(001), 20.32 (20-1), 38.5(111), 40.09 (31–1), and 66.7 (512). The sponds to stretching of Zr–O–Zr bond [31] and Al–O tetrahedral co-
palladium oxide and cobalt oxide peaks were observed for both fresh ordination [42]. The vibration band due to Al–O was seen at 648
and reused 2%Pd-5%Co/AZMOCP catalysts, which were a perfect match cm− 1. The stretching and bending vibration due to the –OH group [31]
for JCPDS # 01–075-0200 (I) and 01–074-2120 (A), respectively. The attached to ZrO2 were present at 1631 cm− 1. The broadband around
PdO peaks correspond to 2θ of 31.1 (002), 33.2 (101), and 54.2 (112). 3373 cm− 1 was most likely due to the stretching vibration of –OH group
Similarly, CoO peaks were found at 37.6 (10–1), 58.2 (2–1-1), and 64.75 of physically adsorbed water molecules [43], which was present in all
(20–2). It is worth noting that due to the formation of Al-Zr mixed oxide, the catalysts (Fig. 4).
all the peaks in Fig. 3 (c-f) were found either overlapped or had smaller
shifts in the 2θ values. For instance, PdO peak found at 46◦ (Fig. 3d and

4
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 6. FESEM of AZMOCP (A and B), 2%Pd-5%Co/AZMOCP (C and D).

3.1.5. TPR doping of Co and Pd leads to the uniform dispersion of these metals on
Fig. 5 depicts the metal-support interaction of various mixed oxide the surface of AZMO (Fig. 6C and D). The focused surface of 2%Pd-5%
catalysts performed using H2-TPR techniques. For this analysis, samples Co/AZMOCP was further analyzed for elemental mapping, and the im­
were freshly calcined and used immediately. The peak for chemisorbed ages for the various elements were discussed in Fig. 7. As per the
H2 was seen in both samples. Impregnation of Pd and Co gave short expectation, the concentration of Al, Zr, and O was very high than that of
peaks at ~ 100 and 200 ◦ C representing a reduction of these metals, and Co and Pd. The values for the percentage concentration of various metals
the broad peak at 545 ◦ C attributed to the strong interaction of Pd-Co were shown in Table S1 (SI). The results of the HRTEM analysis were
with that of AZMO (Fig. 5a and b). The small peak at 73 ◦ C resulted discussed in Fig. 8. Consistency to that of FESEM analysis, the aggre­
because of metallic Pd, which adsorbs H2 at room temperature and forms gated particles of mixed oxide with impregnated Pd and Co metals were
β-PdHx species, and decomposition of this releases H2 on elevating the spread uniformly throughout the sample as seen in dark spots in Fig. 8A
temperature [44]. and B. The magnified area of Fig. 8C reveals the co-existence of aggre­
gates of spherical shaped Al-Zr mixed oxides with that of Pd and Co, and
3.1.6. FESEM, EDS, elemental mapping, and HRTEM it appears to be a composite material than a single species. Furthermore,
The detailed knowledge of the surface morphology was obtained the presence of crystal fringes and SAED patterns (Fig. 8D) reveals that
through FESEM and HRTEM techniques. As indicated in Fig. 6, AZMOCP the sample is highly crystalline.
and 2%Pd-5%Co/AZMOCP both represented highly porous and globule-
type surface morphology due to the densely formed aggregates. The

5
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 7. Elemental mapping of 2%Pd-5%Co/AZMOCP catalyst A) scanning image B) Pd, C) Co, D) Al, E) Zr, and F) O.

3.1.7. TGA-DSC 3.2. Comprehensive study of HMF hydrogenation


The catalyst thermal stability is an important parameter as it reveals
the changes in the catalyst with temperature and therefore, exposes the 3.2.1. Reaction scheme
robustness. Hence, TGA was performed (Fig. 9). There was a slight Scheme 1 illustrates the hydrogenation pathways of HMF, which
weight loss of 2%Pd-5%Co/AZMOCP catalyst (~5%) up to 120 ◦ C, which were reported by various groups [2,25]. As shown in Scheme 1, there are
was because of adsorbed water molecules. The catalyst remained stable mainly two pathways reported by various groups for the production of
from 120 until 600 ◦ C, indicating stability at a wide temperature range, DMF from sequential hydrogenation of HMF: (a) Path 1: Hydrogenation
which was confirmed through a straight line of TGA. of HMF via MFF (preferentially hydrogenation of –CH2OH group of HMF
to –CH3) followed by hydrogenation to MFA (a stable product) and then
3.1.8. XPS finally to DMF, and (b) Path 2: Hydrogenation of HMF (preferential
The composition and oxidation state of different metals in 2%Pd-5% hydrogenation of –C=O group of HMF to –CH2OH) to BHMF followed by
Co/AZMOCP were explored by XPS studies. XPS spectra of Pd 3d, Co 2p, subsequent hydrogenation to MFA and finally to DMF. It was reported
Zr 3d, and O1s are shown (Fig. 10). Due to the strong interaction of Zr 3d by various groups that, when noble metals such as Pd or Ru based cat­
with Pd 3d, there was a shift in the binding energy to a lower value than alysts were used for HMF hydrogenation, it was found to follow path 1
normal (~0.1 to 0.3 eV). The clear doublet at 333.9 and 339.4 eV was (via MFF) predominantly and not path 2 [51,52]. This alteration might
observed in the spectra (Fig. 10A) corresponding to Pd 3d5/2 and 3d3/2, have occurred due to the different properties of the different catalysts;
respectively, which were in accordance with the literature [45,46]. The however, specific factors are not clear yet [1]. It is well-known that it is a
binding energy (B.E) of 335.7 eV was assigned to Pd0 species of metallic sequential hydrogenation reaction and hence, hydrogenation of HMF in
Pd [47]. A smaller shoulder peak was observed at 338.6 eV associated present work proceeds via MFF path, where HMF was first hydrogenated
with Pd4+ form [48]. Al 2p peak at 71.5 eV attributed to oxide form over 2%Pd-5%Co/AZMOCP catalyst (discussed in section 3.2.2) to give
[45], and it was shifted to a lower B.E, which might be due to the co- short-lived species MFF which was not detected in GC; however, seen in
existence with ZrO2. The peak at 181.3 eV corresponds to the 3d5/2 HRMS spectra (SI, Fig. S1) which was immediately hydrogenated to
and ZrO2 [49]. The binding energy of the lattice oxygen was located in MFA. In the next hydrogenation step MFA was hydrogenated to DMF.
the region of 527–530 eV as confirmed from the literature [45], and the Based on these experimental findings, we propose the hydrogenation of
peak observed at 528.4 eV might be due to the Al2O3-ZrO2 mixed oxide HMF via MFF path (Scheme 1). The detail depiction is presented in
(Fig. 10C). Two peaks of Co 2p1/2 and Co metallic form were seen at 792 Scheme 2. HRMS was done for the confirmation of the products and
and 777 eV, respectively [50]. It was worth noting that the intensity of included in the SI (Figs. S1–S3).
Co was suppressed, which might be due to the mobility of oxygen or a
composite formation with Pd, Al, and Zr. 3.2.2. Catalytic performance
In order to choose the best suitable catalyst for the selective hydro­
genation of HMF, screening of different catalysts was done (Fig. 11). As

6
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 8. HRTEM of 2%Pd-5%Co/AZMOCP at different magnification (A and B), (C) Magnified area of (B), and (D) SAED of (C).

per the literature both Co and Pd were used for –CHO group hydroge­ selectivity of DMF was only ~ 70%. Any additional increase in Pd
nation [15,39] and hence, 3%Co/AZMOCP, 5%Co/AZMOCP, and 1% Pd/ loading beyond 2% did not significantly improve selectivity (entry 3,
AZMOCP catalysts were screened for their activity. It was found that both Table S2, SI). The catalytic activity was found in the order of 2%Pd-5%
the conversion of HMF and selectivity of DMF were improved with an Co/AZMOCP > 1%Pd-5%Co/AZMOCP > 2%Pd-3%Co/AZMOCP > 1%Pd-
increase in Co loading. In the case of 1%Pd/AZMOCP, both values of 3%Co/AZMOCP > 5%Ni-5%Co/AZMOCP > 1%Pd/AZMOCP > 5%Co/
conversion of HMF and selectivity of DMF were further improved to 86 AZMOCP > 5%Cu-5%Co/AZMOCP > 3%Co/AZMOCP. As explained in
and 89%, respectively. Bimetallic 2%Pd-5%Co/AZMOCP resulted in a Table 3, the results of the present work were compared with the prior
complete conversion of HMF, giving 97% yield of DMF. It is worth literature for comprehensive understanding. Many efforts were made by
noting that 5%Ni-5%Co/AZMOCP and 5%Cu-5%Co/AZMOCP were used various groups using different catalysts for producing high yield of DMF
to substitute Pd; however, it did not improve the yield of DMF. Although (Table 3). Almost all the reported work (Table 3, entry 1–8) showed
5%Ni-5%Co/AZMOCP successfully gave complete conversion, the complete conversion of HMF; however, the process parameters and

7
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 9. TGA-DSC of 2%Pd-5%Co/AZMOCP catalyst.

Fig. 10. XPS of 2%Pd-5%Co/AZMOCP catalyst. A) Pd 3d, B) Al 2p, C) O 1 s, D) Zr 3d, and E) Co 2p.

selectivity of DMF were different. The selectivity of DMF was strongly 3.2.3. Effect of speed of agitation
dependent on the choice of the catalyst. Mostly, bimetallic combination For studying the absence of external mass transfer resistance, the
gave higher yield of DMF. Moreover, use of Pd (Table 3, entry 3, 4, 7, reaction was carried out varying the speed of agitation at 800, 900, and
and 8) resulted in high selectivity of DMF. The present work (Table 3, 1000 rpm (Fig. 12). It was observed that there was only a marginal
entry 9) showed that our work on developing bimetallic 2%Pd-5%Co/ change in the conversion until 900 rpm, after that any additional in­
AZMOCP catalyst and optimizing the reaction parameters gives 97% crease in the speed of agitation did not increase the conversion of HMF.
yield of DMF, which ultimately should make the process the economical. Therefore, all the experiments were performed at 900 rpm.

8
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Scheme 1. Illustration of hydrogenation of HMF to DMF and the likely paths for production of 2,5-DMF.

Scheme 2. Representation of reaction mechanism for HMF hydrogenation.

3.2.4. Influence of catalyst loading conversion, indicating that the number of active site was just enough to
The effect of catalyst loading on the progress of HMF hydrogenation give maximum conversion, and any additional increase in the catalyst
was studied over 2%Pd-5%Co/AZMOCP catalyst (Fig. 13). As the catalyst loading was not required. Hence 0.008 g/cm3 loading was finalized for
amount altered from 0.005 to 0.01 g/cm3 there was a proportional in­ further study.
crease in the conversion of HMF from 70 to 99.5%. The higher con­
centration of catalyst, therefore, gave higher conversion of HMF. 3.2.5. Effect of the initial concentration of HMF
However, at 0.008 and 0.01 g/cm3, there was no further rise in the The initial concentration of HMF was varied to understand its effect

9
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 11. Catalyst screening. HMF 2 mmol, catalyst loading 0.008 g/cm3, solvent THF, total volume 30 cm3, agitation speed 900 rpm, reaction time 2 h, temperature
100 ◦ C, and H2 pressure 10 atm.

Table 3
HMF hydrogenation over different catalysts.
# Catalyst H2 (atm) Solvent T (◦ C) t (h) Conversion(%) Yield of DMF (%) Ref

1 Raney Ni 15 dioxane 180 15 100 88.5 [18]


2 NiZnAl 15 dioxane 180 15 100 93.6 [19]
3 Cu-Pd@C-B 15 THF 120 7 100 96.5 [22]
4 Fe-Pd/C 20 THF 150 3 100 85 [21]
5 11.8%Co–(ZnO-ZnAl2O4) 7 THF 130 24 99.9 74.2 [23]
6 Cu-Co/Al2O3 30 THF 220 8 100 78 [24]
7 3% Pd/C 20 IPA 170 4 100 99 [25]
8 Pd-Cs2.5H0.5PW12O40/K-10 10 THF 90 2 98 81 [26]
9 2%Pd-5%Co/AZMOCP 10 THF 100 2 >99 97 This work

Fig. 12. Effect of speed of agitation. HMF 2 mmol, 2%Pd-5%/AZMOCP, catalyst Fig. 13. Influence of catalyst loading. HMF 2 mmol, catalyst 2%Pd-5%/
loading 0.008 g/cm3, solvent THF, total volume 30 cm3, reaction time 2 h, AZMOCP, solvent THF, total volume 30 cm3, agitation speed 900 rpm, reaction
temperature 100 ◦ C, and H2 pressure 10 atm. time 2 h, temperature 100 ◦ C, and H2 pressure 10 atm.

10
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 14. Effect of the initial concentration of HMF. 2%Pd-5%/AZMOCP catalyst.


catalyst loading 0.008 g/cm3, solvent THF, total volume 30 cm3, agitation
speed 900 rpm, reaction time 2 h, temperature 100 ◦ C, H2 pressure 10 atm.

Fig. 16. Effect of temperature on conversion of HMF. HMF 2 mmol, 2%Pd-5%/


AZMOCP. catalyst loading 0.008 g/cm3, solvent THF, total volume 30 cm3,
agitation speed 900 rpm, reaction time 2 h, H2 pressure 10 atm.

Both conversion and selectivity were dependent on the temperature.


With an increase in temperature from 80 to 110 ◦ C, the conversion of
HMF was increased from 60 to 99.5%, correspondingly. At temperature
100 ◦ C, all the HMF was consumed and hydrogenated to DMF. Further
increase in temperature did not change this value of conversion, and
hence 100 ◦ C was chosen as the optimum value. Fig. 17 demonstrates
the change in the temperature on the concentration of reactants and
products. At lower temperature (Fig. 17 A and B), MFA and DMF were
formed in equal amounts. Furthermore, the increase in the temperature
significantly improved the DMF yield as more MFA was converted to
DMF because of increase in the rate of reaction (Fig. 17C and D). Hence,
for the high yield of DMF, the temperature was optimized based on the
concentration profile.

Fig. 15. Effect of H2 pressure on hydrogenation of HMF. HMF 2 mmol, 2%Pd- 3.3. Possible reaction mechanism and kinetic model
5%/AZMOCP. catalyst loading 0.008 g/cm3, solvent THF, total volume 30 cm3,
agitation speed 900 rpm, reaction time 2 h, and temperature 100 ◦ C. The mechanism of hydrogenation of HMF follows dual-site Lang­
muir–Hinshelwood–Hougen–Watson (LHHW) model, which was
on both conversion and selectivity of the products. Fig. 14 depicts the consistent with the experimental findings. Initially, HMF (A) was
change in the behavior of the selectivity of DMF and MFA. There was a adsorbed on the acidic site (Sa), and hydrogen was dissociatively
subsequent decrease in both conversion and selectivity of DMF with the adsorbed on metallic site (Sm). In the case of surface reaction, –CH2OH
increase in concentration of HMF. At 1 mmol concentration of HMF, group of (A) hydrogenates to –CH3 group through a short-lived MFF (B)
there was a complete conversion of HMF to DMF. Furthermore, an in­ followed to generate stable species MFA (D) through hydrogenation of
crease in the HMF concentration to 3 mmol reduced the rate of hydro­ –CHO group of MFF (B). Lastly, the second hydrogenation stage occurs
genation. Hence, both conversion of HMF and selectivity of DMF were where –CH2OH group of D gets converted to –CH3 group to form DMF
found to decrease to 88 and 85%, respectively. Therefore, 2 mmol was (E). At desorption stage product E is desorbed and both sites Sa and Sm
chosen to be the optimized concentration of HMF. become vacant for the next catalytic cycle (Scheme 2). Detail elucidation
is discussed in SI, since a similar kinetic study is already performed in
3.2.6. Effect of H2 pressure on hydrogenation of HMF our lab [53].
For studying the effect of H2 pressure, experiments were carried out The rate of consumption of HMF (A):
between 5 and 15 atm (Fig. 15). It was observed that with the increase in √̅̅̅̅̅̅̅̅̅̅̅̅̅
dCA k1 KA CA KH pH2 .w
the H2 pressure the conversion was also increased. At 5 atm H2 pressure, r1 = − = [ √̅̅̅̅̅̅̅̅̅̅̅̅̅ ] (1)
dt
the conversion was 63%, and as the pressure was increased to 10 atm, [1 + KA CA + KD CD + KE CE ] 1 + KH pH2
the conversion of HMF reached 97%. Any additional increase in H2
Rate of formation of intermediate MFA (D):
pressure neither changed conversion nor selectivity (Fig S6, SI). Hence,
√̅̅̅̅̅̅̅̅̅̅̅̅̅
10 atm was chosen to be the optimum for further experiments. dCD (k2 − k1 )KD CD KH pH2 .w
r2 = = [ √̅̅̅̅̅̅̅̅̅̅̅̅̅ ] (2)
dt [1 + KA CA + KD CD + KE CE ] 1 + KH pH2
3.2.7. Effect of temperature
Further study was performed by varying temperature between 80 Rate of formation of DMF (E):
and 110 ◦ C to comprehend its effect on the rate of hydrogenation. Fig. 16
illustrates the proportional increase in the rate of conversion of HMF.

11
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 17. Concentration profile for different temperatures. (A) 80 ◦ C, (B) 90 ◦ C, (C) 100 ◦ C, and (D) 110 ◦ C Reaction condition: HMF 2 mmol, 2%Pd-5%/AZMOCP,
catalyst loading 0.008 g/cm3, solvent THF, total volume 30 cm3, agitation speed 900 rpm, reaction time 2 h, and H2 pressure 10 atm.

reported for two steps hydrogenation of HMF to DMF, where it is stated


Table 4 that lower activation energy indicates the lower temperature required
Adsorption constants and rate constants of different species at different for the reaction [26]. Similarly, apparent activation energy of 48.5 kJ/
temperatures. mol was reported for DMF production indicating a kinetically controlled
#. T Rate constants (cm3 g-1 KA KH KD KE reaction [21]. The kinetics is strongly dependent on the reaction con­
(K) mol− 1 min− 1) x103 (cm3/ x102 x103 ditions, reaction mechanism and catalyst activity. It is worth noting that
(cm3/ mol) (cm3/ (cm3/ above reported reactions followed BHMF path (Path 2) and present work
Formation Formation
mol) mol) mol)
of MFA (k1) of DMF (k2) followed MFF path (Path 1) (see Scheme 1). The activation energies
1 353 0.23 0.3603 0.36 6.9 1.3 2.5 obtained in the present work also suggest that it is a kinetically
2 363 0.16 0.291 0.22 5.3 0.92 1.8 controlled reaction.
3 373 0.102 0.221 0.13 3.9 0.65 1.5
4 383 0.089 0.162 0.07 2.1 0.42 1.1
3.4. Reusability tests and catalyst stability
√̅̅̅̅̅̅̅̅̅̅̅̅̅
dCE k2 KD CD KH pH2 .w The catalyst stability was studied by means of reusability tests
r3 = = [ √̅̅̅̅̅̅̅̅̅̅̅̅̅ ] (3)
dt [1 + KA CA + KD CD + KE CE ] 1 + KH pH2 (Fig. 19). For this, the catalyst was recovered after each cycle and
treated with methanol for three times to remove the adsorbed species,
where, w is catalyst loading (g/cm3). followed by drying. The lost quantity in this process (~3–5%) was
The various terms in equations 1–3 are defined in the nomenclature adjusted by adding fresh catalyst. With each reusability cycle, there was
section. By solving Eqs. (1), (2) and (3) the different adsorption con­ a minor decrease in the conversion; overall, there was about ~ 7%
stants (KA, KD, KE and KH) were calculated using Polymath 6.0 (Table 4). decrease in conversion after fourth reuse. It is probably because of the
The Arrhenius plot was used to calculate activation energies for both loss of active sites from the catalyst surface. All the catalyst samples used
steps, i.e., formation of MFA and DMF as 37.2 and 29.9 kJ/mol, in the experiments were used after they were freshly calcined. The
respectively (Fig. 18). Reports are available for the kinetics of hydro­ catalyst stability was analyzed by characterization technique TGA-DSC
genation of HMF. Activation energies of 61.5 and 36.8 kJ/mol were (section 3.1.7, Fig. 9). In the case of 2%Pd-5%Co/AZMOCP, there was
only 5% weight loss due to loss of moisture and the catalyst remained

12
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

Fig. 18. Arrhenius plot for a) Formation of MFA b) Formation of DMF.

benefit upgrading some other value-added chemicals in biomass-based


fuels and has potential for commercialization.

Declaration of Competing Interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence
the work reported in this paper.

Acknowledgements

D.S.P. acknowledges the University Grants Commission (UGC) for


the award of a BSR Senior Research Fellowship under its Green tech­
nology program. G.D.Y. acknowledges support as the R.T. Mody
Distinguished Professor, Tata Chemicals Darbari Seth Distinguished
Professor of Leadership and Innovation, and J.C. Bose National Fellow of
Fig. 19. Reusability study. HMF 2 mmol, 2%Pd-5%/AZMOCP, catalyst loading the Department of Science and Technology, Government of India.
0.008 g/cm3, solvent THF, total volume 30 cm3, agitation speed 900 rpm, re­
action time 2 h, H2 pressure 10 atm, temperature 100 ◦ C.
Appendix A. Supplementary data
stable from 120 until 600 ◦ C, indicating its stability over a wide tem­
Supplementary data to this article can be found online at https://doi.
perature range. The reusability study of the catalyst demonstrated only
org/10.1016/j.fuel.2020.119947.
about ~ 7% decrease in conversion after fourth reuse, which is good
considering the loss during filtration. NH3-TPD and BET surface area
studies after the first reuse cycle denoted a marginal change when References
compared with the fresh catalyst (Tables 1 and 2). Hence, it confirmed
[1] Wang X, Liang X, Li J, Li Q. Catalytic hydrogenolysis of biomass-derived 5-
that the catalyst is robust supporting its integrity of morphology and hydroxymethylfurfural to biofuel 2, 5-dimethylfuran. Appl Catal A Gen 2019;576:
active sites. 85–95. https://doi.org/10.1016/j.apcata.2019.03.005.
[2] Shi J, Wang Y, Yu X, Du W, Hou Z. Production of 2,5-dimethylfuran from 5-
hydroxymethylfurfural over reduced graphene oxides supported Pt catalyst under
4. Conclusion
mild conditions. Fuel 2016;163:74–9. https://doi.org/10.1016/j.fuel.2015.09.047.
[3] Huang Y-B, Chen M-Y, Yan L, Guo Q-X, Fu Y. Nickel-tungsten carbide catalysts for
Biofules are in great demand in the view of their excellent properties the production of 2,5-dimethylfuran from biomass-derived molecules.
ChemSusChem 2014;7(4):1068–72. https://doi.org/10.1002/cssc.201301356.
among which 2,5-dimethylfuran (DMF) stands out due to its desirable
[4] Guerrero Peña GDJ, Hammid YA, Raj A, Stephen S, Anjana T, Balasubramanian V.
characteristics, and it can be synthesized by selective hydrogenation of On the characteristics and reactivity of soot particles from ethanol-gasoline and
5-hydroxymethylfurfural (HMF), a platform molecule. We evaluated a 2,5-dimethylfuran-gasoline blends. Fuel 2018;222:42–55. https://doi.org/
series of Al-Zr mixed oxide catalysts (AZMO), prepared by co- 10.1016/j.fuel.2018.02.147.
[5] Román-Leshkov Y, Barrett CJ, Liu ZY, Dumesic JA. Production of dimethylfuran for
precipitation and hydrothermal techniques. AZMO prepared by the co- liquid fuels from biomass-derived carbohydrates. Nature 2007;447(7147):982–5.
precipitation method gave higher acidity and surface area for support­ https://doi.org/10.1038/nature05923.
ing both mono- and bimetals. 2%Pd5%Co/AZMOCP showed higher [6] Zu Y, Yang P, Wang J, Liu X, Ren J, Lu G, Wang Y. Efficient production of the liquid
fuel 2,5-dimethylfuran from 5-hydroxymethylfurfural over Ru/Co3O4 catalyst.
surface area and acidity, which was used successfully for hydrogenation Appl Catal B 2014;146:244–8. https://doi.org/10.1016/j.apcatb.2013.04.026.
of bio-based HMF to get liquid biofuel DMF. At 100 ◦ C and 10 atm H2, [7] Liao Y-T, Matsagar BM, Wu K-W. Metal–Organic Framework (MOF)-derived
there was 97% yield of DMF with > 99.5% conversion of HMF. The effective solid catalysts for valorization of lignocellulosic biomass. ACS Sustainable
Chem Eng 2018;6(11):13628–43. https://doi.org/10.1021/
catalyst showed high stability and reusability upto the fourth cycle. The acssuschemeng.8b03683.
activation energies were found to be 37.2 and 29.9 kJ/mol, respectively. [8] Dutta S, Kim J, Ide Y, Ho Kim J, Hossain MSA, Bando Y, et al. 3D network of
Additionally, the effects of various promoters and reaction parameters, cellulose-based energy storage devices and related emerging applications. Mater
Horiz 2017;4(4):522–45. https://doi.org/10.1039/C6MH00500D.
including full characterization of the catalysts were exclusively inves­
[9] Liao Y-T, Nguyen VC, Ishiguro N, Young AP, Tsung C-K, Wu K-W. Engineering a
tigated. 2%Pd5%Co/AZMOCP has proven stable activity as well as homogeneous alloy-oxide interface derived from metal-organic frameworks for
selectivity upto five cycles including the fresh use. This work may selective oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid. Appl
Catal B 2020;270:118805. https://doi.org/10.1016/j.apcatb.2020.118805.

13
D.S. Pisal and G.D. Yadav Fuel 290 (2021) 119947

[10] Cao Q, Liang W, Guan J, Wang L, Qu Q, Zhang X, et al. Catalytic synthesis of 2,5- [31] Mallakpour S, Shafiee E. The synthesis of poly(vinyl chloride) nanocomposite films
bis-methoxymethylfuran: A promising cetane number improver for diesel. Appl containing ZrO2 nanoparticles modified with vitamin B 1 with the aim of
Catal A 2014;481:49–53. https://doi.org/10.1016/j.apcata.2014.05.003. improving the mechanical, thermal and optical properties. Des Monomers Polym
[11] Hu L, Lin Lu, Liu S. Chemoselective hydrogenation of biomass-derived 5-hydrox­ 2017;20(1):378–88. https://doi.org/10.1080/15685551.2016.1273436.
ymethylfurfural into the liquid biofuel 2,5-dimethylfuran. Ind Eng Chem Res 2014; [32] Khaodee W, Tangchupong N, Jongsomjit B, Praserthdam P, Assabumrungrat S.
53(24):9969–78. https://doi.org/10.1021/ie5013807. A study on isosynthesis via CO hydrogenation over ZrO2–CeO2 mixed oxide
[12] Hu L, Tang X, Xu J, Wu Z, Lin L, Liu S. Selective transformation of 5-hydroxyme­ catalysts. Catal Commun 2009;10(5):494–501. https://doi.org/10.1016/j.
thylfurfural into the liquid fuel 2,5-dimethylfuran over carbon-supported catcom.2008.10.017.
ruthenium. Ind Eng Chem Res 2014;53(8):3056–64. https://doi.org/10.1021/ [33] Thimmaraju N, Shamshuddin SZM. Synthesis of 2,4,5-trisubstituted imidazoles,
ie404441a. quinoxalines and 1,5-benzodiazepines over an eco-friendly and highly efficient
[13] Fulignati S, Antonetti C, Licursi D, Pieraccioni M, Wilbers E, Heeres HJ, et al. ZrO2 –Al2O3 catalyst. RSC Adv 2016;6(65):60231–43. https://doi.org/10.1039/
Insight into the hydrogenation of pure and crude HMF to furan diols using Ru/C as C6RA13956F.
catalyst. Appl Catal A 2019;578:122–33. https://doi.org/10.1016/j. [34] Núñez F, Del Angel G, Tzompantzi F, Navarrete J. Catalytic wet-air oxidation of p-
apcata.2019.04.007. cresol on Ag/Al2O3 − ZrO2 Catalysts. Ind Eng Chem Res 2011;50(5):2495–500.
[14] Yang P, Xia Q, Liu X, Wang Y. Catalytic transfer hydrogenation/hydrogenolysis of https://doi.org/10.1021/ie100694g.
5-hydroxymethylfurfural to 2,5-dimethylfuran over Ni-Co/C catalyst. Fuel 2017; [35] Aberuagba F, Kumar M, Muralidhar G, Datt Sharma L. Characterization of
187:159–66. https://doi.org/10.1016/j.fuel.2016.09.026. Al2O3–ZrO2 mixed oxide supported Mo hydrotreating catalyst. Pet Sci Technol
[15] Insyani R, Verma D, Kim SM, Kim J. Direct one-pot conversion of monosaccharides 2004;22(9–10):1287–98. https://doi.org/10.1081/LFT-200034082.
into high-yield 2,5-dimethylfuran over a multifunctional Pd/Zr-based [36] Larese C, Campos-Martin JM, Fierro JLG. Alumina- and zirconia− alumina-loaded
metal–organic framework@sulfonated graphene oxide catalyst. Green Chem 2017; tin− platinum. surface features and performance for butane dehydrogenation.
19(11):2482–90. https://doi.org/10.1039/C7GC00269F. Langmuir 2000;16(26):10294–300. https://doi.org/10.1021/la0009644.
[16] Nishimura S, Ikeda N, Ebitani K. Selective hydrogenation of biomass-derived 5- [37] He J, Li Hu, Riisager A, Yang S. Catalytic transfer hydrogenation of furfural to
hydroxymethylfurfural (HMF) to 2,5-dimethylfuran (DMF) under atmospheric furfuryl alcohol with recyclable Al-Zr@Fe mixed oxides. ChemCatChem 2018;10
hydrogen pressure over carbon supported PdAu bimetallic catalyst. Catal Today (2):430–8. https://doi.org/10.1002/cctc.201701266.
2014;232:89–98. https://doi.org/10.1016/j.cattod.2013.10.012. [38] Liu L, Lou H, Chen M. Selective hydrogenation of furfural over Pt based and Pd
[17] Luo J, Monai M, Wang C, Lee JD, Duchoň T, Dvořák F, et al. Unraveling the surface based bimetallic catalysts supported on modified multiwalled carbon nanotubes
state and composition of highly selective nanocrystalline Ni–Cu alloy catalysts for (MWNT). Appl Catal A 2018;550:1–10. https://doi.org/10.1016/j.
hydrodeoxygenation of HMF. Catal Sci Technol 2017;7(8):1735–43. https://doi. apcata.2017.10.003.
org/10.1039/C6CY02647H. [39] Esen M, Akmaz S, Koç SN, Gürkaynak MA. The hydrogenation of 5-hydroxyme­
[18] Kong X, Zhu Y, Zheng H, Dong F, Zhu Y, Li Y-W. Switchable synthesis of 2,5- thylfurfural (HMF) to 2,5-dimethylfuran (DMF) with sol–gel Ru-Co/SiO2 catalyst.
dimethylfuran and 2,5-dihydroxymethyltetrahydrofuran from 5-hydroxymethyl­ J Sol-Gel Sci Technol 2019;91(3):664–72. https://doi.org/10.1007/s10971-019-
furfural over Raney Ni catalyst. RSC Adv 2014;4(105):60467–72. https://doi.org/ 05047-7.
10.1039/C4RA09550B. [40] Ando C, Ikumoto A, Kurokawa H, Sugiyama K, Miura H. Selective hydrogenation of
[19] Kong X, Zhu Y, Zheng H, Zhu Y, Fang Z. Inclusion of Zn into metallic Ni enables (E)-2-hexenal to (E)-2-hexen-1-ol over Co-based bimetallic catalysts. Catal
selective and effective synthesis of 2,5-dimethylfuran from bioderived 5- Commun 2001;2(10):323–7. https://doi.org/10.1016/S1566-7367(01)00052-8.
hydroxymethylfurfural. ACS Sustainable Chem Eng 2017;5(12):11280–9. https:// [41] Mallick S, Rana S, Parida K. Facile method for the synthesis of phosphomolybdic
doi.org/10.1021/acssuschemeng.7b01813.s001. acid supported on zirconia–ceria mixed oxide and its catalytic evaluation in the
[20] Mhadmhan S, Franco A, Pineda A, Reubroycharoen P, Luque R. Continuous flow solvent-free oxidation of benzyl alcohol. Ind Eng Chem Res 2012;51(23):7859–66.
selective hydrogenation of 5-hydroxymethylfurfural to 2,5-dimethylfuran using https://doi.org/10.1021/ie2022724.
highly active and stable Cu–Pd/reduced graphene oxide. ACS Sustainable Chem [42] Adamczyk A, Długoń E. The FTIR studies of gels and thin films of Al2O3–TiO2 and
Eng 2019;7(16):14210–6. https://doi.org/10.1021/acssuschemeng.9b03017.s001. Al2O3–TiO2–SiO2 systems. Spectrochim Acta Part A Mol Biomol Spectrosc 2012;89:
[21] Talpade AD, Tiwari MS, Yadav GD. Selective hydrogenation of bio-based 5- 11–7. https://doi.org/10.1016/j.saa.2011.12.018.
hydroxymethyl furfural to 2,5-dimethylfuran over magnetically separable Fe-Pd/C [43] Rabee AIM, Le SD, Nishimura S. MgO-ZrO2 mixed oxides as effective and reusable
bimetallic nanocatalyst. Molecular Catalysis 2019;465:1–15. https://doi.org/ base catalysts for glucose isomerization into fructose in aqueous media. Chem
10.1016/j.mcat.2018.12.009. Asian J 2020;15(2):294–300. https://doi.org/10.1002/asia.201901534.
[22] Sarkar C, Koley P, Shown I, Lee J, Liao Y-F, An K, et al. Integration of interfacial [44] Cui W, Li S, Wang D, Deng Y, Chen Y. High reactivity and sintering resistance of
and alloy effects to modulate catalytic performance of metal–organic-framework- CH4 oxidation over modified Pd/Al2O3. Catal Commun 2019;119:86–90. https://
derived Cu–Pd nanocrystals toward hydrogenolysis of 5-hydroxymethylfurfural. doi.org/10.1016/j.catcom.2018.10.028.
ACS Sustainable Chem Eng 2019;7(12):10349–62. https://doi.org/10.1021/ [45] Li G, Hu W, Huang F, Chen J, Gong M, Yuan S, et al. Pd catalyst supported on ZrO2
acssuschemeng.9b00350.s001. -Al2O3 by double-solvent method for methane oxidation under lean conditions : Pd
[23] An Z, Wang W, Dong S, He J. Well-distributed cobalt-based catalysts derived from Catalyst Supported on ZrO2-Al2O3 by Double-Solvent. Can J Chem Eng 2017;95(6):
layered double hydroxides for efficient selective hydrogenation of 5-hydroxyme­ 1117–23. https://doi.org/10.1002/cjce:22750.
thyfurfural to 2,5-methylfuran. Catal Today 2019;319:128–38. https://doi.org/ [46] Yasuda K, Masui T, Miyamoto T, Imanaka N. Catalytic combustion of methane over
10.1016/j.cattod.2018.03.052. Pt and PdO-supported CeO2–ZrO2–Bi2O3/γ-Al2O3 catalysts. J Mater Sci 2011;46
[24] Srivastava S, Jadeja GC, Parikh J. Influence of supports for selective production of (11):4046–52. https://doi.org/10.1007/s10853-011-5333-y.
2,5-dimethylfuran via bimetallic copper-cobalt catalyzed 5-hydroxymethylfurfural [47] Stefanov P, Todorova S, Naydenov A, Tzaneva B, Kolev H, Atanasova G, et al. On
hydrogenolysis. Chin J Catal 2017;38(4):699–709. https://doi.org/10.1016/ the development of active and stable Pd–Co/γ-Al2O3 catalyst for complete
S1872-2067(17)62789-X. oxidation of methane. Chem Eng J 2015;266:329–38. https://doi.org/10.1016/j.
[25] Solanki BS, Rode CV. Selective hydrogenolysis of 5-(hydroxymethyl)furfural over cej.2014.12.099.
Pd/C catalyst to 2,5-dimethylfuran. J Saudi Chem Soc 2019;23(4):439–51. https:// [48] de la Peña O’Shea VA, Alvarez-Galvan MC, Requies J, Barrio VL, Arias PL,
doi.org/10.1016/j.jscs.2018.08.009. Cambra JF, et al. Synergistic effect of Pd in methane combustion PdMnOx/Al2O3
[26] Gawade AB, Tiwari MS, Yadav GD. Biobased green process: selective catalysts. Catal Commun 2007;8:1287–92. https://doi.org/10.1016/j.
hydrogenation of 5-hydroxymethylfurfural to 2,5-dimethyl furan under mild catcom.2006.11.010.
conditions using Pd-Cs2.5H0.5PW12O40/K-10 clay. ACS Sustainable Chem Eng 2016; [49] Guo X, Zhi G, Yan X, Jin G, Guo X, Brault P. Methane combustion over Pd/ZrO2/
4(8):4113–23. https://doi.org/10.1021/acssuschemeng.6b00426.s001. SiC, Pd/CeO2/SiC, and Pd/Zr0.5Ce0.5O2/SiC catalysts. Catal Commun 2011;12(10):
[27] Reddy BM, Patil MK, Reddy BT. An efficient and ecofriendly WOx–ZrO2 solid acid 870–4. https://doi.org/10.1016/j.catcom.2011.02.007.
catalyst for classical mannich reaction. Catal Lett 2008;125(1–2):97–103. https:// [50] Kanakkillam SS, Shaji S, Krishnan B, Vazquez-Rodriguez S, Martinez JAA,
doi.org/10.1007/s10562-008-9518-1. Palma MIM, et al. Nanoflakes of zinc oxide:cobalt oxide composites by pulsed laser
[28] Reddy PS, Sudarsanam P, Raju G, Reddy BM. Selective acetylation of glycerol over fragmentation for visible light photocatalysis. Appl Surf Sci 2020;501:144223.
CeO2–M and SO2− 4 /CeO2–M (M=ZrO2 and Al2O3) catalysts for synthesis of https://doi.org/10.1016/j.apsusc.2019.144223.
bioadditives. J Ind Eng Chem 2012;18(2):648–54. https://doi.org/10.1016/j. [51] Chidambaram M, Bell AT. A two-step approach for the catalytic conversion of
jiec.2011.11.063. glucose to 2,5-dimethylfuran in ionic liquids. Green Chem 2010;12(7):1253.
[29] Barrera A, Montoya JA, Viniegra M, Navarrete J, Espinosa G, Vargas A, et al. https://doi.org/10.1039/c004343e.
Isomerization of n-hexane over mono- and bimetallic Pd–Pt catalysts supported on [52] Wei Z, Lou J, Li Z, Liu Y. One-pot production of 2,5-dimethylfuran from fructose
ZrO2–Al2O3–WOx prepared by sol–gel. Appl Catal A 2005;290(1–2):97–109. over Ru/C and a Lewis–Brønsted acid mixture in N,N-dimethylformamide. Catal Sci
https://doi.org/10.1016/j.apcata.2005.05.011. Technol 2016;6(16):6217–25. https://doi.org/10.1039/C6CY00275G.
[30] Salinas D, Escalona N, Pecchi G, Fierro JLG. Lanthanum oxide behavior in La2O3- [53] Pisal DS, Yadav GD. Single-step hydrogenolysis of furfural to 1,2-pentanediol using
Al2O3 and La2O3-ZrO2 catalysts with application in FAME production. Fuel 2019; a bifunctional Rh/OMS-2 catalyst. ACS Omega 2019;4(1):1201–14. https://doi.
253:400–8. https://doi.org/10.1016/j.fuel.2019.05.015. org/10.1021/acsomega.8b01595.

14

You might also like