Gas-phase hydrodeoxygenation of guaiacol over Fe/SiO2 catalyst

https://doi.org/10.1016/j.apcatb.2011.12.005Get rights and content

Abstract

Lignin could be an important green source for aromatic hydrocarbon production (benzene, toluene and xylenes, BTX). Catalytic hydrodeoxygenation (HDO) of guaiacol was studied over Fe/SiO2 as a model reaction of lignin pyrolysis vapours hydrotreatment. The catalytic conditions were chosen to match with the temperature of never-condensed lignin pyrolysis vapours. The catalyst was characterised by XRD, Mössbauer spectroscopy, N2 sorption and temperature programmed oxidation. A comparison is made with a commercial cobalt-based catalyst. Cobalt-based catalyst shows a too high production of methane. Fe/SiO2 exhibits a good selectivity for BT production. It does not catalyse the aromatic ring hydrogenation. Temperature (623–723 K) and space time (0.1–1.5 gcat h/gGUA) influence the aromatic carbon–oxygen bond hydrogenolysis reaction whereas H2 partial pressure (0.2–0.9 bar) has a minor influence. 38% of BT yield was achieved under the best investigated conditions. Reaction mechanisms for guaiacol conversion over Fe/SiO2 are discussed.

Highlights

► Guaiacol is a model molecule of lignin pyrolysis vapours. ► Guaiacol was hydrotreated using Fe/SiO2 catalyst at 623–723 K. ► These temperatures match with the temperature of lignin pyrolysis vapours. ► Fe/SiO2 is a versatile and selective catalyst for benzene and toluene (BT) production. ► 38% of BT yield was achieved.

Introduction

Lignin is a natural polymer of methoxylated phenylpropane units and it is the second most important compound in lignocellulosic biomass (about 25 wt.%) [1]. It is industrially available at low prices from cellulose pulping processes (fuel value of about $0.04 per pound on a dry basis depending on the technical–economical scenario) [2]. Lignin is today mainly converted to heat and/or power by combustion. Lignin valorisation into aromatic hydrocarbons (benzene, toluene and xylene, BTX) represents a key route for lignocellulosic biorefinery (see Fig. 1) [1], [2], [3], [4].

BTX are widely used as fuel additives and solvents and are starting blocks for several molecules [1], [2], [5]. Conventional technologies for BTX production are: naphtha reforming (C6–C12; over Pt/Al2O3), C2–C6 paraffins conversion on Gallium promoted ZSM-5, coal pyrolysis and methanol-to-gasoline Mobil process on ZSM-5.

Catalytic pyrolysis was proposed as a single-step lignin conversion process [6], [7], [8], [9], [10], [11]. The catalyst is mixed with lignin before pyrolysis. Different lignins were mixed and pyrolysed with NH3-treated H-ZMS-5 or CoO-MoO3/Al2O3 in a pyroprobe [8]. With the zeolite catalyst lignin is first cracked and deoxygenated to C2–C6 olefins which then are converted into BTX. With CoO-MoO3/Al2O3 catalyst, lignin is thought to depolymerise to phenols, and then phenols are deoxygenated on CoO-MoO3 sites to aromatic hydrocarbons [8].

One of the main drawbacks of the direct catalytic pyrolysis is the difficulty to recover and regenerate the catalyst (mixed with the char) [12]. Moreover, catalytic pyrolysis involves many physical–chemical phenomena that are difficult to be monitored in the same reactor.

Lignin can also be valorised in two-step processes consisting in a depolymerisation of lignin and a treatment of depolymerised lignin (lignin bio-oil). Several techniques were proposed for lignin depolymerisation: pyrolysis, liquefaction (with a huge variety of solvents and catalysts), oxidation, electrochemical treatments, etc. [1]. Pyrolysis and liquid phase hydrogenation were the most studied [1]. Liquid phase hydrogenation consists on the high pressure (50–100 bar, 523–723 K) treatment of lignin in a solvent like an aliphatic alcohol, water [13], [14] or phenol [15], [16] under H2 and/or a hydrogen donor (like tetralin). The main disadvantages of liquid treatments are: the consumption of solvent, the cost of high pressure equipment and the difficulties to separate catalyst, char, products, solvent and non-reacted lignin from the reaction media. Lignin pyrolysis produces permanent gases (H2, CO, CO2, CH4), a fluffy char and a complex mixture of condensable compounds rich in OMACs (oxygenated monoaromatic compounds) [12], [17], [18], [19]. Faix et al. [20] studied the products of the non-catalytic pyrolysis of lignin at 600 °C in a fixed bed reactor. 15–30 wt.% of OMACs was obtained using different lignins.

The catalytic conversion of lignin pyrolysis oils (or model compounds like guaiacol) was extensively investigated in the liquid phase under H2 high pressure [12], [13], [21], [22], [23]; and in the gas phase with H2 [24], [25], [26], [27], [28], [29], [30] or without H2 [31].

OMACs are generated in the gas phase during lignin pyrolysis in a reactor set at a temperature of about 773 K [17], [18], [19], [32]. Once condensed, the re-evaporation of bio-oils is difficult since bio-oils are very reactive and cause fouling of pipes. In addition, working on the liquid phase requires high pressures (that means costly equipment) and the condensation step after pyrolysis is detrimental for the heat recovery of the entire process [17]. For these reasons, the vapour phase hydrotreatment of OMACs before condensation seems a better route.

This paper is a contribution on the hydrodeoxygenation (HDO) of OMACs, using guaiacol as a model compound. Guaiacol is a minor but significant component in a very complex mixture in lignin bio-oils [32]. It has been chosen as a model compound because its elemental composition is close to the overall elemental composition of lignin bio-oils (see H/O/C ratios in Fig. 2) and its hydroxyl and methoxyl functions on the aromatic ring are significant and key functions for aromatics species in lignin bio-oil.

The hydrogenation should be soft in order to favour BTX production at the expense of the ring hydrogenation (Fig. 2). The catalyst should act selectively towards the hydrogenolysis of the oxygen–aromatic bonds of hydroxyl or methoxyl functions, i.e. it should be able to activate molecular hydrogen and the C(aromatic)single bondO bond. Moreover, it should be resistant to lignin pyrolysis products and have the desired aspects of any catalyst: low price, low deactivation, easy regeneration and environmentally friendly disposal.

Phenol could also be an interesting product, but allowing some oxygen atoms linked to aromatic ring could produce an endless list of isomers difficult to be separated from the different compounds in bio-oils. Consequently, in our opinion, a complete hydrodeoxygenation to BTX seems a better route instead of a partial HDO to phenols.

Whilst undertaking our work, few papers appeared on the subject using similar reaction conditions but different catalysts (Ni2P/SiO2, Fe2P/SiO2, MoP/SiO2, Co2P/SiO2 and WP/SiO2 [24]; CoMoS over alumina, zirconia or titania [25], [26]; Pt-Sn/CNF [29]). Aims and practical results of these papers are discussed in this manuscript. From the academic point of view, they indicated that the catalysts activity and selectivity strongly depend on the nature of the active phase and that of the support. These factors particularly influence the two main competitive reactions, namely the HDO and the aromatic ring hydrogenation.

Transition (Fe, Co, Ni) and precious metals (Pt, Pd, Ru, Rh, Ir, etc.) are known to catalyse hydrogenation [34], but unfortunately most of them also catalyse benzene hydrogenation in the temperature range 473–673 K [35]. Emmett and Skau [34] detected no activity for benzene hydrogenation at 673 K working with Fe. Lately, Yoon and Vannice [35] measured a relatively low activity of Fe for benzene hydrogenation compared to other transition or precious metals (Ni, Co). Fe is a trade-off between activity and selectivity. This is the reason why we chose iron as the active phase in the guaiacol HDO, expecting minimum aromatic ring hydrogenation.

Metal–support interactions and support acidity play a crucial role in complex chemistry of transition metal supported catalysts. It was studied for hydrogenation of aromatics hydrocarbons [36], [37], [38], [39], [40] as well as for aromatic alcohols [25], [26], [41], [42], [43], [44], [45]. Benzene is hydrogenated with higher rates on Ni catalysts when silica was used as a support instead of alumina [45]. For HDO reaction on supported Mo catalysts, alumina gave higher activity but rapid deactivation due to coke deposit [41], [42], [45]. In contrast, the use of the less acidic active carbon led to lower activity but a better selectivity in HDO products [41], [43], [44], [45]. We concluded that silica would be a good iron carrier due to its low acidity.

Cobalt was chosen for comparison. Cobalt is an inexpensive active phase compared to precious metals. It is a component of typical CoMoS/Al2O3 hydrotreating catalyst and it is active for HDO [46]. Our group performed a screening of catalyst at 623 K [47]. It was shown that cobalt is able to catalyse the production of BT from guaiacol.

For those reasons, our work focuses on the gas phase HDO of guaiacol to BTX as a model compound of lignin pyrolysis vapours using Fe/SiO2 and Co/Kieselguhr catalysts, at the temperature of interest (623–723 K) for coupling the HDO catalytic reactor to the pyrolysis reactor. The catalysts were characterised by BET, XRD, Mössbauer spectroscopy and temperature programmed oxidation (TPO).

Section snippets

Thermodynamic analysis

The chemical equilibrium was studied with GASEQ [48]. 31 compounds were defined: phenols (phenol, cresols, catechols, and guaiacol), ring hydrogenated (cyclohexane, cyclohexene, cyclohexanol, etc.), HDO products (BTX), PAHs, carbon, methanol, water, carbon dioxide and monoxide and alkanes (from methane to C6). Many compounds were not pre-defined on GASEQ library. Their properties (ΔHf°, ΔGf° and Cp(T)) were calculated using THERGAS [49] (based on tabulated data or Benson's method). Starting

Thermodynamic analysis of the reactive system

Some key reactions were analysed in THERGAS [49] to determine the variation of the free Gibbs energy as a function of the temperature. The key reactions (R1–R6) are given in Table 2 and their free Gibbs energy in Table 3.

Table 3 shows that the goal reactions (R1 and especially R2) are possible on the entire range of temperature. The selectivity of BTX formation from phenols HDO should be consequently optimised by a suitable kinetic approach. The formation of saturated rings (R3 or R5) is not

Discussion

It is difficult to compare our result with literature data since publications on guaiacol HDO in the gas phase are scarce and sometimes made from a different purpose than the gas-phase HDO of lignin vapours into BTX. Literature data are presented in Table 4.

Zhao et al. [24] obtained remarkable results: high BT yield (60%) at high HDO conversion (64%) using nickel phosphide. However, the optimal reaction temperature was 573 K, much lower than lignin vapours temperature. It would be interesting to

Conclusion

Fe/SiO2 was shown to be an active and selective catalyst for the conversion of guaiacol into aromatic hydrocarbons at temperatures as high as 673 K matching with the temperature of never-condensed lignin pyrolysis vapours. Hydrogen partial pressure (0.2–0.9 bar) affects product distribution only slightly. Temperature (623–723 K) accelerates reaction rate without big changes in the selectivity. The lower activity of Fe/SiO2 compared with Co-based catalyst is offset by a higher HDO selectivity on a

Acknowledgements

The CNRS (Centre National de la Recherche Scientifique, France)-Programme Interdisciplinaire Energie (« CRAKIN » project) and the MESR (Ministère de l’Enseignement Supérieur et de la Recherche, France) are acknowledged for financial supports. The workshop of LRGP-CNRS, Olivier Herbinet (LRGP-ENSIC), and Michel Mercy (SRSMC-UHP) are kindly acknowledged for technical supports. P.A. Glaude (LRGP-CNRS) is also acknowledged for his advice on thermodynamic calculations.

References (68)

  • M. Jackson et al.

    J. Anal. Appl. Pyrolysis

    (2009)
  • C. Mullen et al.

    Fuel Proc. Technol.

    (2010)
  • R.W. Thring et al.

    Fuel Proc. Technol.

    (2000)
  • R. French et al.

    Fuel Proc. Technol.

    (2010)
  • C. Amen-Chen et al.

    Bioresour. Technol.

    (2001)
  • D. Shen et al.

    Bioresour. Technol.

    (2010)
  • O. Faix et al.

    J. Anal. Appl. Pyrolysis

    (1987)
  • E. Laurent et al.

    Appl. Catal. A

    (1994)
  • H. Zhao et al.

    Appl. Catal. A

    (2011)
  • V. Bui et al.

    Catal. Today

    (2009)
  • V. Bui et al.

    Appl. Catal. B

    (2011)
  • S. Velu et al.

    Appl. Catal. A

    (2003)
  • X. Zhu et al.

    J. Catal.

    (2011)
  • J. Adjaye et al.

    Fuel Proc. Technol.

    (1995)
  • D. Nowakowski et al.

    J. Anal. Appl. Pyrolysis

    (2010)
  • T. Faravelli et al.

    Biomass Bioenergy

    (2010)
  • K. Yoon et al.

    J. Catal.

    (1983)
  • S.D. Lin et al.

    J. Catal.

    (1993)
  • S.D. Lin et al.

    J. Catal.

    (1993)
  • C.H. Bartholomew et al.

    J. Catal.

    (1976)
  • A. Centeno et al.

    J. Catal.

    (1995)
  • M. Huuska et al.

    J. Catal.

    (1985)
  • M. Ferrari et al.

    Catal. Today

    (2001)
  • M. Ferrari et al.

    Microporous Mesoporous Mater.

    (2002)
  • A. Jasik et al.

    J. Mol. Catal. A

    (2005)
  • V. Yakovlev et al.

    Catal. Today

    (2009)
  • A. Dufour et al.

    J. Chromatogr. A

    (2007)
  • J. Rodriguez-Carvajal et al.

    Phys. Condens. Matter

    (1993)
  • A. Dufour et al.

    Appl. Catal. A

    (2009)
  • R. Cypres

    Fuel Proc. Technol.

    (1987)
  • P. Decyk et al.

    J. Catal.

    (2003)
  • E. Shin et al.

    J. Catal.

    (1998)
  • S. Shore et al.

    Catal. Commun.

    (2002)
  • C. Perego et al.

    Microporous Mesoporous Mater.

    (1999)
  • Cited by (0)

    View full text