You are on page 1of 4

Fuel 259 (2020) 116208

Contents lists available at ScienceDirect

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

Short communication

Metal regulating the highly selective synthesis of gamma-valerolactone and T


valeric biofuels from biomass-derived levulinic acid

Zixiao Yia,1, Di Hua,b,1, Hong Xua, Zuotong Wua, Man Zhanga, Kai Yana,b,
a
School of Environmental Science and Engineering, Sun Yat-sen University, 135 Xingang Xi Road, Guangzhou 510275, PR China
b
Guangdong Provincial Key Laboratory of Environmental Pollution Control and Remediation Technology, Guangzhou 510275, PR China

A R T I C LE I N FO A B S T R A C T

Keywords: Highly selective upgrade of biomass-derived source to biofuels is crucial for the utilization of biomass. Herein,
Levulinic acid we report a strategy for the highly selective conversion of biomass-derived levulinic acid (LA) to gamma-va-
Ni/HZSM-5 lerolactone (GVL) or valeric biofuels simply by tuning the used metal Ni or Ru on HZSM-5 support. The catalytic
Ru/HZSM-5 experimental results show that 3 wt% Ni/HZSM-5 catalyst achieves high yield (93.1%) of GVL with negligible
Gamma-valerolactone
formation of pentanoic acid (< 1% yield) or pentanoic esters (< 1% yield), while 3 wt% Ru/HZSM-5 catalyst
Valeric biofuels
exhibits a high yield (85.7%) of pentanoic esters (PE) and pentanoic acid (PA) under the identical conditions. It
Metal effect
is found that the introduction of Ru into HZSM-5 would increase the strong acidic sites and enhance the ring-
opening of GVL intermediate, promoting the formation of PE and PA. In comparison, 3 wt% Ni/HZSM-5 catalyst
shows relative lower acidic sites and negligible ring-opening ability of GVL. The metal effect is further confirmed
using GVL and 1, 4-dioxane as substrate, respectively. After that, reaction kinetics are investigated in details and
a rational pathway is proposed for the metal effect in deciding the product distribution of LA hydrogenation. This
work provides a useful guideline to design proper catalyst for the valorization of biomass resources.

With the consumption of non-renewable fuels, crisis of energy is was formed through hydrogenation of keto-group and dehydration of
arousing extensive concerns around the world [1,2]. Selective up- intramolecular [31,35–38], subsequent hydrogenation of GVL and
grading of biomass-derived source to value-added chemicals and bio- further dehydration under acidic condition leading to the formation of
fuels have attracted extensive attentions for decades [3,4]. Levulinic PA at more severe conditions [11,28]. In this process, PA serves as an
acid (LA), derived from the only sustainable source biomass or biomass- intermediate in the subsequent esterification reaction to generate bio-
based derivatives [5,6], and serves as building block for many biomass- fuels PE. For the conversion of LA to GVL, Ru or Ni based catalysts have
based commodities, can be converted into fuels or fuel additives and been frequently utilized, whereas the high yield of GVL (Over 80%) was
value-added chemicals like gamma-valerolactone (GVL) [7,8], penta- obtained [31,36,37], while PA or PE was mainly produced from GVL
noic acid (PA) [9–11], 2-methyltetrahydrofuran (2-MTHF) [12–14], catalyzed by Cu-based catalysts [15,39,40]. Luo et al. reported the bi-
pentanoic esters (PE) [15–18], and 1,4-pentanediol [12,19,20]. Due to metallic Au-Pd and Ru-Pd/TiO2 for the selective hydrogenation of LA to
its unique property, a comparative research of GVL as fuel additives in a GVL with perfect selectivity. When Ru/HZSM-5 as catalyst, the main
mixture (10 v/v% GVL or ethanol and 90 v/v% 95-octane gasoline) had product PA with over 90% yield was produced [10,29]. Recently, we
been performed and displayed very similar properties, suggesting the [11] reported a facile synthesized Pd nanoparticles confined in meso-
potential of GVL as an attractive liquid fuel [21,22]. Around the line, PE porous MCM-41-derived supports for one-pot conversion of LA to PA,
has been frequently researched as fuel additives and road trials in ve- the optimal yield reached 45.1%. Besides, 60.6% yield of pentyl vale-
hicles [18,23]. Along with this strategy, highly selective conversion of rate (PV) was produced when Pd/HY catalyst was utilized [28].
LA to GVL or valerate biofuels have attracted numerous attention To the best of our knowledge, there is no report on metal effect in
[24–26]. the selective hydrogenation of biomass-derived LA. In this work, we
Hydrogenation of LA into GVL has been well studied over several reported that the used metal Ni or Ru could tune the hydrogenation of
types of catalysts (e.g., metal nanoparticles [11,27,28], nano-alloys LA into GVL or valeric biofuels with perfect selectivity. The catalytic
[29,30], metal oxides [31,32], hydrotalcite [33,34]). Classically, GVL experiment results showed that Ni/HZSM-5 catalyst mainly catalyzed


Corresponding author at: School of Environmental Science and Engineering, Sun Yat-sen University, 135 Xingang Xi Road, Guangzhou 510275, PR China.
E-mail address: yank9@mail.sysu.edu.cn (K. Yan).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.fuel.2019.116208
Received 11 July 2019; Received in revised form 3 September 2019; Accepted 12 September 2019
Available online 18 September 2019
0016-2361/ © 2019 Elsevier Ltd. All rights reserved.
Z. Yi, et al. Fuel 259 (2020) 116208

the formation of GVL, while Ru/HZSM-5 catalyst dominantly produce


PA and PE under the identical conditions. The physical properties of the
fresh and spent catalysts were comparatively studied. Reaction kinetics
and mechanism of metal effect were further investigated using different
intermediates, the rational pathway was proposed for the metal effect
in deciding products distribution of LA hydrogenation.
In this work, we firstly fabricated HZSM-5 according to previous
literatures [41,42] and the supported metal Ni or Ru on HZSM-5 cat-
alysts to achieve multiple effects [43]. X-Ray diffraction (XRD) pattern
was then used to confirm the crystal phase of the successful synthesis
(Fig. S1). The support HZSM-5 exhibited typical XRD patterns, which
corresponded to the single crystalline phase of MFI structure [44,45].
3 wt% Ni/HZSM-5 (denoted as 3NiHZSM) and 3 wt% Ru/HZSM-5 (de-
noted as 3RuHZSM) exhibited typical peaks of Ni indexed as 2θ at 44.5°
(1 1 1) and Ru indexed as 2θ at 44.0° (1 0 1), which was identified as
cubic phase of Ni (JCPDS No. 04-0850) and hexagonal phase of Ru
(JCPDS No. 06-0663) respectively. Energy dispersive spectrometer
(EDS) (Fig. S2) further confirmed the existence of Ru or Ni on HZSM-5.
Fig. 1. Catalytic conversion of LA to GVL, PA and PE over supports and sup-
ICP analysis showed the ratio which was close to the initial submission ported Ru or Ni catalystsa. a: Reaction conditions: 4 mmol LA in 10 ml 1, 4-
value. The morphology of HZSM-5 and resulting catalysts are shown in dioxane, 0.1 g catalyst, 220 °C for 10 h in 3 MPa initial H2 atmosphere. The
Fig. S3. The spherical particle of pristine HZSM-5 support kept intact dosage of catalyst was constant. b: 4 mmol GVL as feedstock.
after loading of metal nanoparticles, showing the supporting of metal
(Ru or Ni) didn’t change the morphology structure of HZSM-5 support.
acidic support HZSM-5 produced a small amount of PA (1.2% yield).
TEM images of the resulting 3RuHZSM and 3NiHZSM (Fig. S4)
NH3-TPD of TiMCM-41, ZrMCM-41 and HZSM-5 were also tested as
displayed that Ru and Ni particles were uniformly dispersed over the
shown in Fig. S7, the modified mesoporous supports by Ti or Zr dis-
HZSM-5 support with little agglomeration. Inset images of Fig. S4a and
played relatively weak acidity. After introduction of metal Ru into the
S4b exhibited the mean sizes of metal particles were 3.6 nm and 4.2 nm
mesoporous supports, 3RuZrMCM and 3RuTiMCM catalysts exhibited
for Ru and Ni nanoparticles, respectively. As shown in Fig. S4c and S4d,
virtually perfect conversion, and yields of GVL reached around 90% as
the lattice fringe was clearly presented and the d spacing of 0.206 nm
well as a small amount of PA (< 5% yield). When the support was
and 0.203 were measured, which well corresponded to the (1 0 1) lat-
switched to strong acidic support HZSM-5, 3NiHZSM catalyst per-
tice fringe of Ru and the (1 1 1) lattice fringe of Ni, respectively. X-ray
formed analogously to 3NiTiMCM catalyst, whereas GVL was formed as
photoelectron spectroscopy (XPS) was further conducted to study the
main products (yield 93.1%). Surprisingly, for RuHZSM catalyst, under
surface states of the resulting catalysts (Fig. S5). Ru (3p3/2) peak at
the identical conditions, 3RuHZSM exhibited the yield of PA and PE
461.9 eV was observed, which indicated the fresh 3RuHZSM catalyst
(yield 85.7%) whereby very little of GVL (1.2% yield) was formed. The
was reduced successfully. Additionally, three peaks were observed from
produced PE biofuels mainly contained methyl valerate and ethyl va-
3NiHZSM sample at Ni 2p3/2 region, which included a satellite peak,
lerate. Additionally, the catalytic effects on different reaction tem-
and the results indicated that not just metal Ni0 existed in the catalyst,
perature, time, source and hydrogen pressure were depicted in Fig. S12.
but also the oxidation state of Ni was present [9,46]. Ammonia tem-
To elucidate the reaction pathway, GVL was further chosen as the
perature programmed desorption (NH3-TPD) was further employed to
substrate. As expected, GVL got nearly full conversion to PA and PE
test acidity of HZSM-5 and the as-prepared catalysts as well as modified
catalyzed by 3RuHZSM catalyst while nearly no conversion (< 1.5%)
supports (Fig. S6). It can be observed that HZSM-5 and the two catalysts
was detected by 3NiHZSM catalyst. These data confirms that the
first kept a desorption peak at around 200 °C, indicating the existence of
RuHSZM catalyst has strong ability of ring-opening, while NiHZSM
weak acid sites. The desorption peak at around 410 °C for all three
catalyst doesn’t have. Furthermore, the ring-opening reaction was fur-
samples represented moderate acid sites. The amount of different acidic
ther performed using 1, 4-dioxane solvent as the substrate, it was very
sites was quantified and tabulated as shown in Table 1. TPD results
interesting to find methanol and ethanol were produced. However,
confirmed that the parent HZSM-5 held weak acid and moderate acid
NiHZSM catalyst did not catalyze the ring-opening of 1, 4-dioxane at
sites, and the total acid amount was 2.09 mmol/g while 3RuHZSM and
all. It further confirms the strong ring-opening ability of RuHZSM cat-
3NiHZSM catalysts both generated strong acid sites at around 535 °C.
alyst. Based on the above data, the rational pathway was thus proposed
The 3RuHZSM catalyst displayed relatively stronger acid sites than that
in Scheme 1. GVL could be formed at the first stage catalyzed by Ni or
of 3NiHZSM, which could contribute to the ring-opening ability of GVL
Ru nanoparticles, while RuHZSM catalyst hold strong acidic sites, and
leading to the formation of PA or PE.
very strong ring-opening ability of GVL and 1, 4-dioxane. The produced
The different supports and its supported Ru or Ni nanoparticle
PA from GVL would react with methanol and ethanol from 1, 4-dioxane
catalysts were evaluated in the hydrogenation of LA and the catalytic
through esterification to produce PE.
performance was shown in Fig. 1. Apparently, the neutral support
To explore the catalytic stability, we concentrated on the recycle
MCM-41 had scarce effects on the hydrogenation reaction, while the
capacity of 3RuHZSM catalyst and the catalytic performance was shown
in Fig. S8. Yield of PA and PE slightly decreased as the number of cycles
Table 1 increase. This was possible ascribe to the carbon deposits and the un-
Acidity properties of the parent HZSM-5 and its supported metal catalysts. avoidable catalyst loss during the recycle procedure. Thermogravi-
Catalysts Weak acidity Moderate acidity Strong acidity metric analysis (TGA) experiment indicated ~17.5% carbon deposits on
the spent catalyst after 4 recycle runs (Fig. 2). Fortunately, simply by
Td (°C) Amount Td (°C) Amount Td (°C) Amount
calcination and reduction, the activity could reach the same level as
(mmol/g) (mmol/g) (mmol/g)
fresh catalyst. XRD was further utilized to compare the fresh, the spent
HZSM-5 201 0.98 413 1.11 – – and regenerated catalysts (Fig. S9), no clear change of crystal phase
3RuHZSM 196 1.01 419 1.05 533 0.26 occurred, indicating the stable structure. TEM images of the spent
3NiHZSM 198 0.96 402 1.15 536 0.17 catalyst after 2nd and 3rd run were also shown in Fig. S10, clear

2
Z. Yi, et al. Fuel 259 (2020) 116208

Scheme 1. Proposed reaction pathway in the hydrogenation of LA over Ni/HZSM-5 or Ru/HZSM-5 catalysts in 1, 4-dioxane solvent.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://


doi.org/10.1016/j.fuel.2019.116208.

References

[1] Yan L, Yao Q, Fu Y. Conversion of levulinic acid and alkyl levulinates into biofuels
and high-value chemicals. Green Chem 2017;19(23):5527–47.
[2] Climent MJ, Corma A, Iborra S. Conversion of biomass platform molecules into fuel
additives and liquid hydrocarbon fuels. Green Chem 2014;16(2):516.
[3] Yan K, Yang Y, Chai J, Lu Y. Catalytic reactions of gamma-valerolactone: a feedstock
for fuels and chemicals. Appl Catal B: Environ 2015;179:292–304.
[4] Lipinsky ES. Chemicals from biomass: petrochemical substitution options. Science
1981;212(4502):1465–71.
[5] Zhang J, Wu S, Li B, Zhang H. Advances in the catalytic production of valuable
levulinic acid derivatives. ChemCatChem 2012;4(9):1230–7.
[6] Wei W, Wu S. Experimental and kinetic study of glucose conversion to levulinic acid
in aqueous medium over Cr/HZSM-5 catalyst. Fuel 2018;225:311–21.
Fig. 2. TGA and DTG curves of the fresh and spent 3RuHZSM after 4th run.
[7] Hu D, Xu H, Yi Z, Chen Z, Ye C, Wu Z, et al. Green CO2-assisted synthesis of mono-
and bimetallic Pd/Pt nanoparticles on porous carbon fabricated from sorghum for
highly selective hydrogenation of furfural. ACS Sustainable Chem Eng 2019.
agglomeration occurred during the recycle experiments. The spent
https://doi.org/10.1021/acssuschemeng.9b02665.
catalyst was further characterized by XPS shown in Fig. S11. Ru 3p3/2 [8] Sudhakar M, Kumar VV, Naresh G, Kantam ML, Bhargava SK, Venugopal A. Vapor
orbital was observed no change during the reaction process, further phase hydrogenation of aqueous levulinic acid over hydroxyapatite supported metal
(M = Pd, Pt, Ru, Cu, Ni) catalysts. Appl Catal B 2016;180:113–20.
explained the great performance of 3RuHZSM catalyst ascribe to the
[9] Velisoju VK, Gutta N, Tardio J, Bhargava SK, Vankudoth K, Chatla A.
stable existing Ru0. Besides, the slight peak of RuO2 was observed, Hydrodeoxygenation activity of W modified Ni/H-ZSM-5 catalyst for single step
which was possibly due to the partial oxidation or interaction between conversion of levulinic acid to pentanoic acid: an insight on the reaction mechanism
Ru and OH group from the support surface. and structure activity relationship. Appl Catal A 2018;550:142–50.
[10] Luo W, Bruijnincx PCA, Weckhuysen BM. Selective one-pot catalytic conversion of
In summary, the highly selective synthesis of GVL (93.1% yield) or levulinic acid to pentanoic acid over Ru/H-ZSM5. J Catal 2014;320:33–41.
valeric biofuels (85.7% yield) has been successfully achieved using the [11] Wang A, Lu Y, Yi Z, Ejaz A, Hu K, Zhang L. Selective production of γ-valerolactone
facilely fabricated NiHZSM-5 or RuHZSM-5 catalyst, respectively. It is and valeric acid in one-pot bifunctional metal catalysts. ChemistrySelect
2018;3(4):1097–101.
found that RuHZSM-5 catalyst has strong capacity of ring-opening of [12] Patankar SC, Yadav GD. Cascade Engineered Synthesis of γ-Valerolactone, 1,4-
GVL intermediate and 1, 4-dioxane solvent, prompting the synthesis of Pentanediol, and 2-Methyltetrahydrofuran from Levulinic Acid Using Pd-Cu/ZrO2
PE. In comparison, NiHZSM-5 catalyst has very weak ring-opening Catalyst in Water as Solvent. ACS Sustainable Chem Eng 2015;3(11):2619–30.
[13] Mizugaki T, Togo K, Maeno Z, Mitsudome T, Jitsukawa K, Kaneda K. One-Pot
ability of GVL. Overall, the metal effect reported here may be useful for transformation of levulinic acid to 2-methyltetrahydrofuran catalyzed by Pt-Mo/H-
the design of earth-abundant metal catalyst in the selective processing β in water. ACS Sustainable Chem Eng 2016;4(3):682–5.
of biomass. [14] Xie ZB, Chen BF, Wu HR, Liu MY, Liu HZ, Zhang JL, et al. Highly efficient hydro-
genation of levulinic acid into 2-methyltetrahydrofuran over Ni-Cu/Al2O3-ZrO2
bifunctional catalysts. Green Chem 2019;21(3):606–13.
[15] Li W, Li Y, Fan G, Yang L, Li F. Role of surface cooperative effect in copper catalysts
Acknowledgements toward highly selective synthesis of valeric biofuels. ACS Sustainable Chem Eng
2017;5(3):2282–91.
[16] Sun P, Gao G, Zhao Z, Xia C, Li F. Stabilization of cobalt catalysts by embedment for
The authors sincerely appreciate professor Ruifeng Li from Taiyuan efficient production of valeric biofuel. ACS Catal 2014;4(11):4136–42.
University of Technology for providing partial HZSM-5 support, and the [17] Sun P, Gao G, Zhao Z, Xia C, Li F. Acidity-regulation for enhancing the stability of
work was supported by National Key R&D Program of China Ni/HZSM-5 catalyst for valeric biofuel production. Appl Catal B 2016;189:19–25.
[18] Lange JP, Price R, Ayoub PM, Louis J, Petrus L, Clarke L, et al. Valeric biofuels: a
(2018YFD0800700), National Ten Thousand Talent Plan, National platform of cellulosic transportation fuels. Angew Chem Int Ed
Natural Science Foundation of China (21776324), the Key Research & 2010;49(26):4479–83.
Development Program of Guangdong Province, China [19] Cui J, Tan J, Zhu Y, Cheng F. Aqueous hydrogenation of levulinic acid to 1,4-
pentanediol over Mo-modified Ru/activated carbon catalyst. ChemSusChem
(2019B110209003) and the “Hundred Talent Plan” from Sun Yat-sen 2018;11(8):1316–20.
University. [20] Mizugaki T, Nagatsu Y, Togo K, Maeno Z, Mitsudome T, Jitsukawa K, et al. Selective
hydrogenation of levulinic acid to 1,4-pentanediol in water using a hydroxyapatite-
supported Pt-Mo bimetallic catalyst. Green Chem 2015;17(12):5136–9.

3
Z. Yi, et al. Fuel 259 (2020) 116208

[21] Horváth IT, Mehdi H, Fábos V, Boda L, Mika LT. γ-Valerolactone—a sustainable Technol 2017;7:1622–45.
liquid for energy and carbon-based chemicals. Green Chem 2008;10(2):238–42. [35] Yan K, Chen A. Selective hydrogenation of furfural and levulinic acid to biofuels on
[22] Savage N. Fuel options: the ideal biofuel. Nature 2011;474:S9. the ecofriendly Cu-Fe catalyst. Fuel 2014;115:101–8.
[23] Contino F, Dagaut P, Dayma G, Halter F, Foucher F, Mounaïm-Rousselle C. [36] Feng J, Gu X, Xue Y, Han Y, Lu X. Production of gamma-valerolactone from levu-
Combustion and emissions characteristics of valeric biofuels in a compression ig- linic acid over a Ru/C catalyst using formic acid as the sole hydrogen source. Sci
nition engine. J Energy Eng 2014;140(3):A4014013. Total Environ 2018;633:426–32.
[24] Luo W, Deka U, Beale AM, van Eck ERH, Bruijnincx PCA, Weckhuysen BM. [37] Yang Y, Gao G, Zhang X, Li F. Facile fabrication of composition-tuned Ru-Ni bi-
Ruthenium-catalyzed hydrogenation of levulinic acid: influence of the support and metallics in ordered mesoporous carbon for levulinic acid hydrogenation. ACS Catal
solvent on catalyst selectivity and stability. J Catal 2013;301:175–86. 2014;4(5):1419–25.
[25] Bozell JJ. Connecting biomass and petroleum processing with a chemical bridge. [38] Moustani C, Anagnostopoulou E, Krommyda K, Panopoulou C, Koukoulakis KG,
Science 2010;329(5991):522–3. Bakeas EB, et al. Novel aqueous-phase hydrogenation reaction of the key bior-
[26] Pan T, Deng J, Xu Q, Xu Y, Guo QX, Fu Y. Catalytic conversion of biomass-derived efinery platform chemical levulinic acid into γ-valerolactone employing highly
levulinic acid to valerate esters as oxygenated fuels using supported ruthenium active, selective and stable water-soluble ruthenium catalysts modified with ni-
catalysts. Green Chem 2013;15(10):2967–74. trogen-containing ligands. Appl Catal B 2018;238:82–92.
[27] Yang Y, Sun C, Ren Y, Hao S, Jiang D. New route toward building active ruthenium [39] Yan K, Jarvis C, Gu J, Yan Y. Production and catalytic transformation of levulinic
nanoparticles on ordered mesoporous carbons with extremely high stability. Sci Rep acid: a platform for fuels and commodity chemicals. Renew Sustain Energy Rev
2014;4:4540. 2015;51:986–97.
[28] Yan K, Lafleur T, Wu X, Chai J, Wu G, Xie X. Cascade upgrading of gamma-valer- [40] Liu S, Fan G, Yang L, Li F. Highly efficient transformation of γ-valerolactone to
olactone to biofuels. Chem Commun 2015;51(32):6984–7. valerate esters over structure-controlled copper/zirconia catalysts prepared via a
[29] Luo W, Sankar M, Beale AM, He Q, Kiely CJ, Bruijnincx PCA, et al. High performing reduction-oxidation route. Appl Catal A 2017;543:180–8.
and stable supported nano-alloys for the catalytic hydrogenation of levulinic acid to [41] Liu Z, Fan W, Ma J, Li R. Adsorption, diffusion and catalysis of mesostructured
gamma-valerolactone. Nat Commun 2015;6:6540. zeolite HZSM-5. Adsorption-J Int Adsorption Soc 2012;18(5–6):493–501.
[30] Zhang L, Mao J, Li S, Yin J, Sun X, Guo X, et al. Hydrogenation of levulinic acid into [42] Xu D, Ma J, Zhao H, Liu Z, Li R. Adsorption and diffusion of n-heptane and toluene
gamma-valerolactone over in situ reduced CuAg bimetallic catalyst: strategy and over mesostructured ZSM-5 zeolitic materials with acidic sites. Fluid Phase Equilib
mechanism of preventing Cu leaching. Appl Catal B 2018;232:1–10. 2016;423:8–16.
[31] Song S, Yao S, Cao J, Di L, Wu G, Guan N, et al. Heterostructured Ni/NiO composite [43] Jiang H, Bongard H, Schmidt W, Schüth F. One-pot synthesis of mesoporous Cu-γ-
as a robust catalyst for the hydrogenation of levulinic acid to γ-valerolactone. Appl Al2O3 as bifunctional catalyst for direct dimethyl ether synthesis. Microporous
Catal B 2017;217:115–24. Mesoporous Mater 2012;164:3–8.
[32] Kuwahara Y, Kaburagi W, Osada Y, Fujitani T, Yamashita H. Catalytic transfer [44] Popova M, Djinovic P, Ristic M, Lazarova H, Drazic G, Pintar A, et al. Vapor-phase
hydrogenation of biomass-derived levulinic acid and its esters to γ-valerolactone hydrogenation of levulinic acid to gamma-valerolactone over Bi-functional Ni/
over ZrO2 catalyst supported on SBA-15 silica. Catal Today 2017;281:418–28. HZSM-5 catalyst. Front Chem 2018;6:285.
[33] Long X, Sun P, Li Z, Lang R, Xia C, Li F. Magnetic Co/Al2O3 catalyst derived from [45] Zhang Y, Jiang H. A novel route to improve methane aromatization by using a
hydrotalcite for hydrogenation of levulinic acid to γ-valerolactone. Chin J Catal simple composite catalyst. Chem Commun 2018;54(73):10343–6.
2015;36(9):1512–8. [46] Li X, Xiang M, Wu D. Hydrogenolysis of glycerol over bimetallic Cu-Ni catalysts
[34] Yan K, Liu Y, Lu Y, Chai J, Sun L. Catalytic application of layered double hydro- supported on hierarchically porous SAPO-11 zeolite. Catal Commun
xides-derived catalysts for the conversion of biomass-derived molecules. Catal Sci 2019;119:170–5.

You might also like