Anda di halaman 1dari 7

View Article Online / Journal Homepage / Table of Contents for this issue

Catalysis
Science &Technology
www.rsc.org/catalysis Volume 2 | Number 1 | January 2012 | Pages 1–228
Published on 07 September 2011. Downloaded on 10/09/2014 18:20:29.

ISSN 2044-4753

COVER ARTICLE
Gates et al.
Catalytic conversion of compounds representative of lignin-derived bio-oils: a reaction network for guaiacol, anisole,
4-methylanisole, and cyclohexanone conversion catalysed by Pt/-Al2O3
View Article Online
Catalysis Dynamic Article Links

Science & Technology


Cite this: Catal. Sci. Technol., 2012, 2, 113–118

www.rsc.org/catalysis PAPER
Catalytic conversion of compounds representative of lignin-derived
bio-oils: a reaction network for guaiacol, anisole, 4-methylanisole,
and cyclohexanone conversion catalysed by Pt/c-Al2O3w
Ron C. Runnebaum,a Tarit Nimmanwudipong,a David E. Blockab and
Bruce C. Gates*a
Published on 07 September 2011. Downloaded on 10/09/2014 18:20:29.

Received 13th May 2011, Accepted 7th August 2011


DOI: 10.1039/c1cy00169h

The conversion of compounds representative of lignin and lignin-derived bio-oils (guaiacol, anisole,
4-methylanisole, and cyclohexanone), catalysed by Pt/Al2O3 in the presence of H2 at 573 K is
described by a reaction network indicating a high selectivity for platinum-catalysed aromatic
carbon–oxygen bond cleavage accompanied by acid-catalysed methyl group transfer reactions.

1. Introduction linkages—the compounds are guaiacol, anisole, 4-methylanisole,


and cyclohexanone. These reactants were converted in the
Among the potential routes for production of fuels and presence of a solid catalyst, platinum on g-Al2O3 (Pt/g-Al2O3).
chemicals from lignocellulosic biomass, fast pyrolysis accom-
panied by or followed by catalytic upgrading offers excellent
potential because the number of conversion steps is small and 2. Experimental
the processing may be cost effective.1 Catalytic upgrading is 2.1 Catalytic reaction experiments
required because the pyrolysis products are characterised
by thermal and chemical instability, low heating values, The Pt/g-Al2O3 catalyst (1 wt% Pt, Sigma-Aldrich) had a BET
corrosiveness, and immiscibility with petroleum-derived fuels.2 surface area of 206  1 m2 g 1 and a platinum dispersion
Research on catalytic upgrading of bio-oils has focused on determined by H2 chemisorption of 0.25.10 The catalyst was
cellulose-derived fractions,3 with less attention paid to used in powder form.
lignin-derived fractions—which offer attractive prospects for In the reaction experiments, carried out with a once-through
production of aromatic chemicals such as phenolics.4 Lignin- flow reactor with reactant vapors flowing at steady state, the
derived bio-oils can also be converted into fuels, with a key catalyst (0.001–0.100 g, mixed with particles of inert nonporous
processing challenge being the removal of oxygen—because a-Al2O3) was pre-treated in flowing gases at 573 K. The pre-
bio-oils derived from lignocellulose typically contain 35–40 wt% treatment gas, flowing at 100 mL min 1, was 30% H2/70% N2.
oxygen.2,5 The literature of bio-oils conversion is largely lacking In a typical experiment, liquid reactant (guaiacol, flowing at
in fundamental chemistry.6 The complexity of the feedstocks 0.015 mL min 1, or anisole, 4-methylanisole, or cyclohexanone,
limits the usefulness of the available data for predicting catalyst each flowing at 0.030 mL min 1) was vaporised into a gas stream
performance. Even data characterising the reactions of indivi- (30% H2/70% N2) flowing at 100 mL min 1 into the packed-bed
dual biomass-derived compounds7 are scarce, and the catalytic reactor. The reactor temperature was 573 K, and the pressure
reaction networks are largely limited to primary products or was 140 kPa. The effluent was condensed at 283–288 K.
products lumped into classes of similar compounds.8,9 Analysis of the products in the liquid samples was
Our goal was to determine a reaction network to account for performed with a gas chromatograph (GC, Agilent 7890)
the reactions of a group of compounds prototypical of lignin and equipped with a mass selective detector (MSD, Agilent 5795)
compounds derived from it, incorporating the representative and a flame ionisation detector. Analysis of the gas-phase
aromatic rings, hydroxyl groups, keto groups, and ether products was performed with a GC refinery gas analyser
(Agilent 7980). Details of the analysis are given in ESI.
a
Department of Chemical Engineering and Materials Science, Mass balance closures were typically greater than 95%.
University of California, Davis, Davis CA, USA.
E-mail: bcgates@ucdavis.edu; Fax: (+ 1) 530-752-1031
b
Department of Viticulture and Enology, University of California, 3. Results
Davis, Davis CA, USA
w Electronic supplementary information (ESI) available: The details of 3.1 Oxygen removal in reactions catalysed by Pt/c-Al2O3
materials and methods, table showing product list from the catalytic
reactions of each of the prototypical compounds. See DOI: 10.1039/ The products formed in the reactions catalysed by Pt/g-Al2O3
c1cy00169h in the presence of H2 are listed in Tables 1 and 2. In the

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 113–118 113
View Article Online

Table 1 Productsa formed with oxygen removal in the conversion of Table 2 Productsa formed without oxygen removal in the conversion
compounds representative of lignin and lignin-derived bio-oils of compounds representative of lignin and lignin-derived bio-oils
catalysed by Pt/g-Al2O3 in the presence of H2 catalysed by Pt/g-Al2O3 in the presence of H2

Products formed in conversion Product Products formed in conversion Product


Reactant of prototypical reactant structure Reactant of prototypical reactant structure

Guaiacol Anisole (methoxybenzene) Guaiacol Catechol (benzene-1,2-diol)


(2-methoxyphenol)

3-Methylcatechol
Cyclohexanone (3-methyl-1,2-benzenediol)

3-Methylguaiacol
Phenol (2-methoxy-3-methylphenol)
Published on 07 September 2011. Downloaded on 10/09/2014 18:20:29.

o-Cresol (2-methylphenol) 6-Methylguaiacol


(2-methoxy-6-methylphenol)

Veratrole (1,2-dimethoxybenzene)
Anisole Benzene
(methoxybenzene)
Cyclohexene
Anisole Phenol
Cyclohexane

Cyclohexanone
4-Methylanisole Toluene
(1-methoxy-4- (methylbenzene)
methylbenzene)
2-Methylphenol

Cyclohexanone Benzene
4-Methylphenol
Cyclohexane

2,6-Dimethylphenol
Cyclohexene

a
A list of the trace products is given in the ESI.
4-Methylanisole 4-Methylphenol

conversion of each of the reactants individually with H2, we


observed dozens of products. The data determine conversions Methylcyclohexanone
and selectivities of the products that were formed in relatively
high yields at conversions o 10%. Conversion of the four
individual reactants as a function of time on stream, shown in 2,4-Dimethyphenol
Fig. 1, demonstrates similar rates of catalyst deactivation.
Plots of selectivity as a function of conversion (e.g., Fig. 2)
demonstrate which of the major products are primary and
2,4,6-Trimethylphenol
which are not; the methodology for these determinations is
summarised in the papers by Bhore et al.11
The data show that one of the dominant classes of reactions
Tetramethylphenol
observed with guaiacol, anisole, and 4-methylanisole was
transalkylation. Reported results indicate that with HY zeolite
catalyst, transalkylation was the only kinetically significant
Cyclohexanone Cyclohexanol
reaction class.10,12 This observation implies that the alumina
support in Pt/g-Al2O3 catalysed the transalkylation reactions.
The other important reaction classes observed with Pt/g-
Cyclohexenone
Al2O3 (but only when H2 was a coreactant) were hydrogenation
(e.g., the conversion of cyclohexanone to cyclohexanol and of
benzene to cyclohexene), dehydrogenation (e.g., the conversion Phenol
of cyclohexanone to phenol and hydrogen), and hydrogenolysis a
A list of the trace products is given in the ESI.
(e.g., the reaction of guaiacol to give catechol and methane).

114 Catal. Sci. Technol., 2012, 2, 113–118 This journal is c The Royal Society of Chemistry 2012
Published on 07 September 2011. Downloaded on 10/09/2014 18:20:29. View Article Online

Fig. 1 Change in conversion of anisole (K), 4-methylanisole (m),


cyclohexanone (n), and guaiacol (‘) during operation of a flow
reactor. Reactants were fed individually and all conversions were
catalysed by Pt/g-Al2O3 in the presence of H2 at 573 K. The WHSV,
in units of (g of reactant)/(g of catalyst  h), of the feed streams was
chosen to achieve similar initial conversions: anisole, 18; 4-methyl-
anisole, 17; cyclohexanone, 160; and guaiacol, 20.

There was no evidence of hydroxylation reactions under


our experimental conditions; for example, anisole was not
observed to produce guaiacol.
C–O bond cleavage reactions that removed oxygen from the Fig. 2 (a) Selectivity for the formation of phenol (K), anisole (J),
organic reactant were observed, as were C–O bond cleavage and benzene (.) in the conversion of guaiacol. (b) Selectivity for the
reactions that did not remove oxygen from the organic formation of 4-methylphenol (B) and toluene (E) in the conversion of
4-methylanisole. Each conversion was catalysed by Pt/g-Al2O3 in the
reactant. The former are categorised as hydrodeoxygenation
presence of H2 at 573 K. Data for each product were fitted with a
(HDO) and the latter as hydrogenolysis. In the HDO
straight line and extrapolated to zero conversion; intercepts of regres-
reactions, oxygen was removed as methanol or water sion lines significantly different from zero selectivity at zero conversion
(exemplified by the reaction of 4-methylanisole to give toluene (determined with 95% confidence limits) indicate primary products, in
and methanol). In contrast, in hydrogenolysis, one of the this case phenol, anisole, 4-methylphenol, and toluene.
products was an alkane (as in the reaction of anisole to give
phenol and methane). In the conversion of cyclohexanone, conversions were kept low in almost all of our experiments
HDO was observed, but, because of the lack of substitutents to simplify the analyses, we lack the data to determine the
such as methyl, hydrogenolysis was not. reaction orders precisely and use pseudo-first-order kinetics
The observation of these three reaction classes when the as a matter of convenience. The rate constants are summarised
catalyst was the supported metal and not when it was the in Table 3.
zeolite demonstrates that they were catalysed by the platinum.

3.2 Identification of primary reactions and determination of


kinetics: first step towards elucidation of reaction networks
The data determine quantitative conversions and selectivities
of the products that were formed in relatively high yields at
conversions o 10%.
In the conversion of guaiacol with H2, phenol and anisole
are primary products, and benzene was a non-primary
product, as shown by the plot of Fig. 2a. The plot of Fig. 2b
shows that 4-methylphenol and toluene are primary products
in the reactions of 4-methylanisole and H2. (We emphasise
that the designations of primary and non-primary products
are empirical and do not provide details about intermediates
that were too reactive to be detected.) Fig. 3 Conversion of guaiacol to give phenol (n) and conversion of
The data also determine approximate kinetics of the anisole to give benzene (K); both conversions were catalysed by
primary reactions. As illustrated by the data of Fig. 3, which Pt/g-Al2O3 in the presence of H2 at 573 K. The term Xi represents
are typical, the reactions are satisfactorily represented as the conversion to product i. WHSV is weight hourly space velocity.
first-order in the organic reactant. However, because the The vertical scale is logarithmic.

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 113–118 115
View Article Online

Table 3 Pseudo-first-order rate constants characterising reactions in The data provide a comparison of the reactivities of the
the network of Fig. 4 various reactants. For example, the rate constants representing
Reaction the reactions of guaiacol and anisole to remove the methoxy
number group are 4.4 and 0.86 L (g of catalyst) 1 h 1, respectively.
(keyed to Rate This comparison indicates that the methoxy group becomes
Fig. 4) Reactant constanta Reaction class easier to remove when a hydroxy group is present in the ortho
1 Guaiacol 0.11 Hydrodeoxygenation position. The rate constant for removal of the methyl group
2 Guaiacol 4.4 Hydrodeoxygenation from the methoxy substituent is greater for anisole, 12 L
3 Guaiacol 6.5 Hydrogenolysis
4 Guaiacol 1.8 Transalkylation (g of catalyst) 1 h 1, than for guaiacol, 6.5 L (g of catalyst) 1 h 1.
5 Guaiacol 0.50 Bimolecular transalkylation
6 Guaiacol 0.21 Bimolecular transalkylation 4.2 Evaluation of the reaction network and comparison with
7 Guaiacol 0.26 Bimolecular transalkylation reported data
8 Anisole 12 Hydrogenolysis
9 Anisole 0.86 Hydrodeoxygenation A statement of the reaction network including the minor and
10 Anisole 2.8 Transalkylation trace products is presented in Fig. 5; we inferred this network
Published on 07 September 2011. Downloaded on 10/09/2014 18:20:29.

11 Anisole 0.039 Transalkylation


12 Anisole 0.14 Bimolecular transalkylation
by recognising which compounds were primary products and
13 4-Methylanisole 0.76 Hydrodeoxygenation by using our chemical judgment of the most likely pathways
14 4-Methylanisole 4.2 Hydrogenolysis for formation of the minor and trace compounds by presuming
15 4-Methylanisole 2.2 Transalkylation that the important reaction classes are methyl group transfer,
16 Cyclohexanone 77 Dehydrogenation
17 Cyclohexanone 5.5 Hydrogenation hydrodeoxygenation, hydrogenolysis, and hydrogenation.
18 Cyclohexanone 0.05 Hydrodeoxygenation The network of Fig. 5 provides a more complete description
a
Rate constants unit in L (g catalyst) 1
h 1. than that given in Fig. 4, but it is not quantitative. The
inference that the most important reaction classes are
hydrodeoxygenation, transalkylation, hydrogenolysis, and
4. Discussion hydrogenation is in good agreement with earlier reports,8
but the results provide detailed new evidence of the reactions.
4.1 Combined reaction network
This is a more detailed representation of the catalytic reactions
The data form the basis for determining a reaction network to of lignin-derived compounds than any reported.
summarise the quantitative results, as shown in Fig. 4. This As represented in Fig. 5, for example, the data show that
network shows all the primary products formed from each some of the reactions are reversible (e.g., cyclohexanone
reactant. produces phenol), but others appeared to be irreversible at

Fig. 4 Reaction network accounting for formation of primary products determined from analysis of selectivity-conversion plots for the
conversion of individual reactants (shown in red), guaiacol, anisole, 4-methylanisole, and cyclohexanone, catalysed by Pt/g-Al2O3 in the presence
of H2 at 573 K. H2, as a reactant, is omitted for simplicity. Pseudo-first-order rate constants for the individual primary products formed in the
conversion of the individual reactants catalysed by Pt/g-Al2O3 are shown in Table 3; the numbers next to the arrows in this figure are keyed to the
list in Table 3.

116 Catal. Sci. Technol., 2012, 2, 113–118 This journal is c The Royal Society of Chemistry 2012
Published on 07 September 2011. Downloaded on 10/09/2014 18:20:29. View Article Online

Fig. 5 Reaction network for the conversion of lignin-derived compounds (each compound shown in red was used as a reactant) with H2 catalysed
by Pt/g-Al2O3 at 573 K and 140 kPa. HDO, hydrogenolysis, and hydrogenation (or dehydrogenation) reactions are represented by dashed green,
blue, and black arrows, respectively. Transalkylation reactions are represented by solid black arrows. H2 as a reactant is omitted for simplicity. The
representation in this network is simplified: for example, in transalkylation reactions in which two guaiacol molecules are involved (e.g., 2-guaiacol
- catechol + veratrole), the stoichiometry is not represented here.

the relatively low conversions observed in our experiments The data characterising the conversion of anisole and of
(e.g., 4-methylanisole gave negligible yields of anisole). The 4-methylanisole show that the conversions of reactants in
data are too few to identify reversible reactions involving trace homologous series give homologous products (e.g., benzene
products. and toluene, respectively). These results suggest that the data
The results indicate that methyl group transfer reactions may be extrapolated with some confidence to increasingly
were negligible when the methyl group was bonded to the higher homologues including those found in pyrolysis oils.13
aromatic ring; for example, only a trace of 2-methylanisole The results highlight the requirement of higher partial
was observed to form from 4-methylanisole. We contrast this pressures of H2 than we used when selective deoxygenation
observation with the relatively rapid methyl group transfer is a goal; for example, deoxygenated compounds such as
reactions occurring when the methyl group migrated from toluene and xylenes are valuable high-octane-number fuel
the methoxy group to the aromatic ring, as, for example, in the components. The data also point to potentially useful routes
formation of 2,4-dimethylphenol from 4-methylanisole. The for production of chemicals. For example (Fig. 5), the
data also imply the simultaneous occurrence of intermolecular products phenol and cyclohexanone could be useful as chemical
and intramolecular methyl group transfers; for instance, both intermediates.
2-methylanisole and o-cresol were identified as primary products The results also clearly demonstrate the value of metals as
in the reactions of anisole. Similarly, methylguaiacols (formed catalysts for HDO reactions. We suggest further that the
by intermolecular transalkylation) and 3-methylcatechol results presented here may be useful in guiding the conversion
(formed by intramolecular transalkylation) were identified as of lignin fractions produced in new processes for biomass
primary products in the conversion of guaiacol. conversion, such as those using molten salt catalysts.4,14

This journal is c The Royal Society of Chemistry 2012 Catal. Sci. Technol., 2012, 2, 113–118 117
View Article Online

Conclusions 3 For examples see: G. W. Huber, J. N. Chheda, C. J. Barrett and


J. A. Dumesic, Science, 2005, 308, 1446–1450; G. W. Huber,
In summary, the reaction network of Fig. 5 represents our R. D. Cortright and J. A. Dumesic, Angew. Chem., 2004, 116,
data well, accounting quantitatively for the kinetically impor- 1575–1577 (Angew. Chem., Int. Ed., 2004, 43, 1549); P. Gallezot,
Catal. Today, 2007, 121, 76–91; Y. Román-Leshkov, C. J. Barrett,
tant reaction classes. This reaction network is much more Z. Y. Liu and J. A. Dumesic, Nature, 2007, 447, 982–985.
extensive than any reported for the catalytic conversion of 4 J. Zakzeski, P. C. A. Bruijnincx, A. L. Jongerius and B. M.
lignin-derived compounds. The data show that the hydro- Weckhuysen, Chem. Rev., 2010, 110, 3552–3599.
5 D. C. Elliott and T. R. Hart, Energy Fuels, 2009, 23, 631–637.
genolysis reactions (including HDO) are kinetically significant 6 For examples see: R. K. Sharma and N. N. Bakhshi, Fuel Process.
under our operating conditions and require a metal function Technol., 1991, 27, 113–130; R. K. Sharma and N. N. Bakhshi,
and H2. Transalkylation reactions occur in the presence of an Bioresour. Technol., 1991, 35, 57–66; R. K. Sharma and
acid such as our catalyst support (g-Al2O3), which is not active N. N. Bakhshi, Energy Fuels, 1993, 7, 306–314; R. K. Sharma
and N. N. Bakhshi, Biomass Bioenergy, 1993, 5, 445–455;
for oxygen removal reactions. Thus, the identification of the J. D. Adjaye and N. N. Bakhshi, Fuel Process. Technol., 1995,
roles of the acid and metal functions in the bifunctional 45, 161–183; D. C. Elliot, T. R. Hart, G. G. Neuenschwander,
catalyst implies an opportunity to optimize the catalyst for a L. J. Rotness and A. H. Zacher, Environ. Prog. Sustainable Energy,
particular application by balancing these functions and opti- 2009, 28, 441–449.
Published on 07 September 2011. Downloaded on 10/09/2014 18:20:29.

7 For examples see: C. Zhao, Y. Kou, A. A. Lemonidou, X. Li and


mizing processes for the conversion of lignin-derived bio-oils J. A. Lercher, Angew. Chem., Int. Ed., 2009, 48, 3987–3990;
to fuels and chemicals. D. Y. Hong, S. J. Miller, P. K. Agrawal and C. W. Jones, Chem.
Commun., 2010, 46, 1038–1040; A. Gutierrez, R. K. Kaila,
M. L. Honkela, R. Siloor and A. O. I. Krause, Catal. Today,
Acknowledgements 2009, 147, 239–246.
8 P. D. Chantal, S. Kaliaguine and J. L. Grandmaison, Appl. Catal.,
We thank Jennifer Heelan of the University of California, 1985, 18, 133–145; X. Zhu, R. G. Mallinson and D. E. Resasco,
Davis, for help with the analytical instrumentation. Support of Appl. Catal., A, 2010, 379, 172–181.
this work was provided by Chevron Technology Ventures, a 9 A. G. Gayubo, A. T. Aguayo, A. Atutxa, R. Aguado and J. Bilbao,
Ind. Eng. Chem. Res., 2004, 43, 2610–2618; A. G. Gayubo,
division of Chevron USA, Inc. An Agilent Technologies A. T. Aguayo, A. Atutxa, R. Aguado, M. Olazar and J. Bilbao,
Foundation Research Project Gift provided a GC7890 Ind. Eng. Chem. Res., 2004, 43, 2619–2626.
Refinery Gas Analyzer. Tarit Nimmanwudipong thanks the 10 T. Nimmanwudipong, R. C. Runnebaum, D. E. Block and
Fulbright Foundation for an Open Competition scholarship. B. C. Gates, Catal. Lett., 2011, 41, 779–783.
11 N. A. Bhore, M. T. Klein and K. B. Bischoff, Ind. Eng. Chem. Res.,
1990, 29, 313–316; N. A. Bhore, M. T. Klein and K. B. Bischoff,
Notes and references Chem. Eng. Sci., 1990, 45, 2109–2116.
12 R. C. Runnebaum, T. Nimmanwudipong, D. E. Block and
1 T. P. Vispute, H. Zhang, A. Sanna and G. W. Huber, Science, B. C. Gates, Catal. Lett., 2011, 41, 817–820.
2010, 330, 1222–1227; M. Stöcker, Angew. Chem., Int. Ed., 2008, 13 G. M. Loudon, Organic Chemistry, Benjamin/Cummings, Menlo
47, 9200–9211. Park, 1988, p. 57.
2 G. W. Huber, S. Iborra and A. Corma, Chem. Rev., 2006, 106, 14 B. A. Simmons, D. Loque and J. Ralph, Curr. Opin. Plant Biol.,
4044–4098; S. Czernik and A. V. Bridgwater, Energy Fuels, 2004, 2010, 13, 313–320; S. H. Lee, T. V. Doherty, R. J. Linhardt and
18, 590–598. J. S. Dordick, Biotechnol. Bioeng., 2009, 102, 1368–1376.

118 Catal. Sci. Technol., 2012, 2, 113–118 This journal is c The Royal Society of Chemistry 2012

Anda mungkin juga menyukai