You are on page 1of 11

Chapter 6

Catalytic Hydrodeoxygenation
of Lignin Model Compounds

Basudeb Saha, Ian Klein, Trenton Parsell, and Mahdi M. Abu-Omar

Abstract Constituting about 20–35 % by weight of dry wood biomass, lignin


accounts for ca. 40 % of the energy in lignocellulosic biomass and is the only renew-
able feedstock composed of aromatic building block. Therefore, efficient lignin uti-
lization is important in the viability of future biorefineries. Herein we review recent
work highlighting selective catalytic depolymerization of intact lignin into phenolic
monomers. Furthermore, we explore advances in catalysis allowing for upgrading
of lignin-derived aromatic functionalized monomers through hydrodeoxygenation,
yielding both intact saturated and unsaturated aromatic products.

Keywords Lignin depolymerization • Hydrodeoxygenation • Phenolic aromatic


building block • Lignocellulosic biomass

6.1 Introduction

High volatility in crude oil price due to demand–supply imbalance and awareness of
climate change necessitate the utilization of biorenewable resources on a large scale
for manufacturing chemicals and biofuels [1]. The existing biorefinery industries
(biofuels and bio-based chemicals) prefer to use feedstock with high sugar or starch
content, commonly known as the first-generation feedstock, such as grains, sugar
beet, sugarcane, etc. However, due to additional land requirements and adverse
effect on food and fiber crops, it was realized that the first-generation biomass feed-
stock based on edible crops could not achieve the desired target for petrochemical
substitution and economic growth [2]. Therefore, the utilization of nonfood

B. Saha • I. Klein • M.M. Abu-Omar (*)


Department of Chemistry and the Center for Catalytic Conversion of Biomass
to Biofuels (C3Bio), Purdue University, West Lafayette, IN 47907, USA
Spero Energy, Inc., 1281 Win Hentschel Blvd., West Lafayette, IN 47906, USA
e-mail: mabuomar@purdue.edu
T. Parsell
Spero Energy, Inc., 1281 Win Hentschel Blvd., West Lafayette, IN 47906, USA

© Springer Science+Business Media Singapore 2016 119


M. Schlaf, Z.C. Zhang (eds.), Reaction Pathways and Mechanisms
in Thermocatalytic Biomass Conversion II, Green Chemistry and Sustainable
Technology, DOI 10.1007/978-981-287-769-7_6
120 B. Saha et al.

Table 6.1 Cellulose, Component Wood (%) Non-wood (%)


hemicellulose, and lignin
Cellulose 40–45 30–45
composition in wood and
non-wood lignocelluloses [3] Hemicellulose 23–35 20–35
Lignin 20–35 3–25

lignocellulosic biomass, also known as second-generation feedstock, such as waste


wood, forest wood residue, sugarcane bagasse, corn stover, and grasses including
sorghum and miscanthus, has received attention as sustainable feedstock in recent
years [3]. However, as compared to first-generation biomass, lignocellulose has
slightly lower sugar content, but more critically, the sugars are more difficult to
access because of cross-linking of sugar monosaccharides in hemicellulose and cel-
lulose with phenolic units of recalcitrant lignin [4, 5]. Cellulose, hemicellulose, and
lignin composition in lignocellulose varies in the range of 40–45, 23–35, and
20–35 %, respectively, by weight of dry biomass as shown in Table 6.1.
Due to the complex structure of plant cell walls, lignocellulose is commonly
pretreated to make plant cell walls more amenable to digestion by enzymatic and
chemical catalysis. In the pretreatment processes, lignocellulose is treated with
acid, alkali, ammonia, or hot water/steam under harsh reaction conditions, and the
lignin fraction is eventually separated [6].

6.2 Lignin Potential

Representing about 20–35 % by weight of dry woody biomass, lignin biopolymer is


structurally much more complex than cellulose and hemicellulose. The complex
structure of lignin contains numerous ether linkages, OH groups, and methoxy
groups. This cross-linked macromolecule is composed of three types of monoli-
gnols, also known as phenolic monomers, including p-coumaryl alcohol, coniferyl
alcohol, and sinapyl alcohol [7]. Figure 6.1 shows that these units are intercon-
nected through various cross-linkages (C–O–C: β–O–4, α–O–4, 4–O–5) and C–C
interunit linkages (β–1, β–5, β–β, 5–5) in macromolecular lignin. The most abun-
dant is the β–O–4 ether linkage [8], representing about 50–65 % of all inter-subunit
bonds, as shown in Table 6.2.
While chemical conversions of cellulose and hemicellulose to bio-based chemi-
cals and biofuels have been extensively studied, lignin is neglected as a biorenew-
able feedstock despite the fact that lignin is the only source for renewable aromatic
building block chemicals and contains about 40 % of the total energy content of
lignocellulosic biomass [11, 12]. Thus, utilization of lignin as a feedstock for the
production of hydrocarbons and chemicals offers a significant opportunity for
enhancing the overall operational efficiency, carbon conversion rate, economic via-
bility, and sustainability of biorefinery processes [13]. However, this requires selec-
tive methods to break apart the 3D network of lignin while preserving the aromaticity
of the building block units.
6 Catalytic Hydrodeoxygenation of Lignin Model Compounds 121

Fig. 6.1 Schematic depiction of lignin showing various linkages and lignin model compounds.
Model (A) phenol and methoxy functionalities, (B) a β–O–4 linkage, (C) a 5–5′ linkage, (D) a
propyl side chain, and (E) a benzylic group (Ref. [9] Reproduced by permission of The Royal
Society of Chemistry)

Table 6.2 Types and frequencies (%) of inter-subunit linkages in softwood and hardwood lignin
[9, 10]
Linkage Softwood lignin Hardwood lignin
β-O-4 49–51 65
O α-O-4 6–8 –
β β-5 9–15 6
α
γ β-1 2 15
1
6 2 5–5 10 2
3 4-O-5 3 2
5
4 OMe β-β 2 6
O

6.3 Challenges to Selective Lignin Depolymerization

The challenge, however, is selective upgrading of lignin biopolymer into a manage-


able stream of aromatic products. Given the intrinsic heterogeneity of lignin as a
polymer in lignocellulosic biomass, most depolymerization processes including
thermal, catalytic, and biological yield a heterogeneous mixture of oxygenated aro-
matics [11, 14]. Catalytic depolymerization of lignin has been extensively studied
over the past several decades [15–17], but with limited success in terms of yield and
selectivity of phenolic products. For example, depolymerization of wood lignin
with Raney Ni, Pd/C, Rh/C, Rh/Al2O3, and Ru/Al2O3 catalysts has been reported by
122 B. Saha et al.

Table 6.3 Main applications for lignin


Group Volume Value Application type
Power/heat High Low Carbon source for energy production
Macromolecules Low Medium High-molecular-mass applications such as
wood adhesives (binders)
Carbon fibers
Filler for polymers
Aromatics Low Medium to Polymer building block chemicals
high Aromatic monomers (benzene, toluene,
xylene)
Phenol, methoxyphenols
Vanillin

Pepper et al. more than 40 years ago [18–20]. Although these reactions using spruce
wood have achieved maximum 34 % conversion of total available wood lignin by
Rh/C catalyst, the resultant product stream contained six different phenols. Similarly,
Pd/C- and HZSM-5-catalyzed cleavage of C–O bonds in monomeric and dimeric
phenol substrates produced a mixture of cyclohexane products [17]. Recent analysis
by liquid chromatography/high-resolution tandem mass spectrometry demonstrated
the formation of hundreds of molecule from organosolv switchgrass lignin, which
makes product separation difficult and expensive [21].
Because of the aforementioned challenges, lignin is currently underutilized. The
bulk of produced lignin, representing about 70 million metric tons per year, is only
employed as a combustible material for its heat value [13]. Table 6.3 summarizes
the current state of lignin usages and the processes that are being undertaken for
development.

6.4 Lignin Depolymerization

As the development of efficient processes for selective upgrading of lignin into


high-value aromatic chemicals, e.g., methoxyphenols and vanillin, has received sig-
nificant attention in recent years to valorize lignin as well as to improve the overall
economy of lignocellulosic biorefinery processes, several journal and patent articles
have been published within the past few years [9, 22–26]. Among these articles, two
recent publications by the groups of Abu-Omar and Xu stand out in terms of selec-
tive upgrading of wood lignin into high-value methoxyphenols: 2-methoxy-4-
propylphenol (dihydroeugenol) and 2,6-dimethoxy-4-proylphenol (4-propylsyringol)
[22, 23]. While the process reported by Xu group, demonstrating maximum 54 %
6 Catalytic Hydrodeoxygenation of Lignin Model Compounds 123

Table 6.4 Conversion of lignin starting with intact lignocellulosic biomass over Pd/C/Zn2+
catalyst
Biomass type Total lignin (wt %)a Selectivity (%) Yield (wt%)b
Dihydroeugenol 4-propylsyringol
Poplar 18 28 72 44
White birch 16 31 69 52
Eucalyptus 24 30 70 49
Lodgepole pine 31 100 – 19
a
Lignin content as determined by ABSL (acetyl bromide-soluble lignin) lignin analysis
b
Yield is calculated using the initial mass of lignin and the mass of the products, factoring in the
loss of two atoms of oxygen for each mole of product produced

lignin conversion of birch wood with Ni/C catalyst and high selectivity (89 %) of
the resultant methoxyphenols, showed promise for selective upgrading of wood,
this process did not evaluate the potential of the carbohydrate fractions of birch
wood. Additionally, this process has some limitations on commercial-scale biorefin-
ery processing, including the possibility of catalyst deactivation by the process-
generated water and the limitation of the strategy requiring a specific variety of birch
wood sawdust.
In this context, Pd/C/Zn2+-catalyzed depolymerization of a wide variety of wood
lignin, e.g., poplar wood containing variable ratio of S and G lignin mutant, euca-
lyptus, white birch and pine, and selective upgrading of the corresponding monoli-
gnols into dihydroeugenol and 4-propylsyringol, has not only achieved high yield
(maximum 54 %) of the two methoxyphenols with nearly 100 % selectivity, but this
process also demonstrated the potential of different biomass usages, thus improving
the suitability for commercial biorefinery processing. Table 6.4 shows yield and
selectivity of the two methoxyphenol products from different biomass feedstock.
Depending on the nature of woods, as measured by the ratio of G and S mono-
mers in the lignin, the selectivity of the two products varied. For example, pine
wood containing only G lignin exclusively produced dihydroeugenol. After cataly-
sis and extraction of methoxyphenols, leftover solid carbohydrate residue was sub-
jected to cellulase enzyme digestion under standard enzymatic hydrolysis conditions
[27]. The authors reported that the carbohydrate residue produced 95 % of the theo-
retical glucose yield. In comparison, intact poplar wood released only 11 % of theo-
retical glucose from cellulase enzyme digestion over the same time period. These
results are consistent with those reported in the literature showing that cellulose can
be hydrolyzed by enzymes once recalcitrant lignin is removed, as lignin inhibits
cellulose hydrolysis [28]. Further analysis of soluble sugars, mostly xylose, in the
liquid phase suggested nearly complete mass balance of the catalytic products after
catalysis as illustrated in Fig. 6.2.
124 B. Saha et al.

Fig. 6.2 Mass balance after catalytic cleavage and HDO of WT poplar lignin over Zn2+/Pd/C cata-
lyst. *Mass of phenolic products includes quantified dihydroeugenol and 4-propylsyringol and
also accounts for the loss of O into H2O during hydrodeoxygenation. Liquid-phase sugars were
quantified by HPLC analysis. The solid-phase residue was hydrolyzed with acid. Then glucose,
arabinose, and xylose were quantified by HPLC analysis (Ref. [22] Reproduced by permission of
the Royal Society of Chemistry)

6.5 Upgrading Lignin Model Compounds

Several research articles have been published in recent years illustrating catalytic
hydrodeoxygenation (HDO) of lignin model compounds to produce high-value
chemicals and hydrocarbons. Catalytic HDO to remove covalently bonded oxygen
from lignin offers opportunities to not only break apart the complex structure of
lignin but also to increase the overall value of the resulting products. The major
complexity is that C–O and C–C bond strengths are comparable, meaning that
selective HDO versus aromatic hydrogenation is challenging. The recent publica-
tions have addressed such critical issues. For example, a bifunctional catalyst based
on Pd metal nanoparticles on carbon support and ZnCl2 can selectively cleave C–O
bonds in a variety of lignin models under relatively mild conditions (150 °C and
2–20 bar H2) [29]. In the presence of both Pd/C and Zn2+, the benzyl alcohol and
aldehyde groups were selectively deoxygenated in good yields without hydrogena-
tion of the phenyl ring. Most importantly, monomeric lignin surrogates can be deox-
ygenated depending on the synergy between the Pd/C and Zn2+. The β–O–4 linkage
of the lignin macromolecule is the most abundant repeating subunits, and selective
cleavage of such ether linkages by carrying out HDO using the Pd/C–Zn2+ system
provides a way of unfolding the complex polymeric structure of lignin into small
molecules of high value. In this model, guaiacyglycerol-β-guaiacyl ether was hydro-
genated and deoxygenated at 150 °C under 20 bar of hydrogen using Zn2+-Pd/C in
MeOH yielding 85 % dihydroeugenol as a major product along with a small amount
of 3-(3-methoxy-4-hydroxyphenyl)-1-propanol as shown in equation 1 of Scheme
1. The catalyst performance was tested for the β–O–4 synthetic lignin polymer
(Mn = 3390 and DPI = 12), which underwent complete conversion with excellent
yields (80 %) of two major products, dihydroeugenol (56 %) and 3-(3-methoxy-4-
hydroxyphenyl)-1-propanol (44 %) (Reaction 2 in Fig. 6.3).
6 Catalytic Hydrodeoxygenation of Lignin Model Compounds 125

Fig. 6.3 β-O-4 dimeric and polymeric lignin cleavage and HDO to remove ether oxygen using
Zn2+/Pd/C system in methanol

A MoO3 catalyst has also been reported as an efficient HDO catalyst for the con-
version of lignin-derived oxygenates to aromatic hydrocarbons without ring-
saturated products [30]. The catalyst achieved highest selectivity in the C–O bond
cleaved products under low H2 pressures (1 bar) and temperatures ranging from 220
to 250 °C. Analysis of bond dissociation energy for relevant phenolic C–O bonds
indicated that the bond strengths followed an order of Ph–OH > Ph–OMe > Ph–O–
Ph > PhO–Me, although for all model compounds investigated, the MoO3 catalyst
preferentially cleaved phenolic Ph–OMe bonds over weaker aliphatic PhO–Me
bonds. Control experiments for understanding catalyst deactivation revealed that a
partial carbon deposition on the catalyst surface occurred, which might play a role
in stabilizing a Mo5+ state on the MoO3 surface and in slowing down the over-
reduction of MoO3 into an inactive MoO2 phase.
In contrary to selective cleavage of C–O bonds without aromatic ring saturation,
catalytic systems have been developed for the production of ring-saturated aromatic
hydrocarbons in which hydrogenation of aromatic ring or side-chain propenyl group
occurs predominantly over the subsequent HDO of C–O bonds. Catalytic HDO of
anisole, a methoxy-rich lignin model compound, is an example in which anisole has
been deoxygenated over a series of Ni-containing (10 wt% loading) catalysts sup-
ported on activated carbon, SBA-15, SiO2, and Al2O3, to understand the effect of
supports on the cleavage of C–O bonds from anisole [31]. Under investigated reac-
tion conditions of 180–220 °C and 5–30 bar H2 pressure, the authors demonstrated
that the hydrogenation of aromatic ring took place predominantly over the subse-
quent demethylation and deoxygenation steps. Among the four catalysts, Ni/SiO2
displayed the highest activity in HDO of anisole, possibly due to the high dispersion
of metal sites on the catalyst surface and the acidity of the catalyst support. Similar
effect of acidic support has been reported for catalytic HDO of 2-methoxyphenol
(guaiacol) and 2-methoxy-4-propenylphenol (eugenol) [32]. Evaluation of a combi-
nation of metals and supports under comparable reaction conditions showed that the
126 B. Saha et al.

Table 6.5 Hydrogenation of eugenol over Pt and Ru catalysts


Supports Metals Products

SiO2-Al2O3 Ru 44 10 15 0 3
Pt 75 7 16 0 0
γ-Al2O3 Ru 59 11 4 0 13
Pt 92 2 0 0 0
C Ru 77 6 0 0 2
Pt 0 0 0 95 0
Hydrotalcite Ru 60 5 1 0 12
Pt 80 8 0 0 2

yield and selectivity of the hydrogenation products are highly dependent on the
nature of the catalyst support as illustrated in Table 6.5. The results show that
γ-Al2O3 and SiO2–Al2O3 containing acidic sites achieved high degree of deoxygen-
ation followed by hydrogenation of the propenyl chain.
Bimetallic Re–Pt catalysts on different supports (ZrO2, CeO2, TiO2, AC, SiO2)
have been developed for catalytic HDO of n-propylphenol to the corresponding
n-propylbenzene [33]. Among all investigated supports, ZrO2 supported catalyst
achieved maximum yield of propylbenzene with about 80 % selectivity. The homo-
geneous Re-catalytic system (methyldioxorhenium (MDO)) was also found to be
effective for C–O bond cleavage of lignin β-O-4-model compound,
2-(2-methoxyphenoxy)-1-phenyl ethanol, under mild reaction conditions [34]. The
mechanistic studies of these reactions, giving phenolic and aldehyde compounds as
the main products, showed that the reduction of ReVII to ReV by the substrate itself
generated the catalytically active species, MDO, which was responsible for the C–O
bond activation. The reaction achieved quantitative cleavage of C–O bonds at
135 °C for 12 h in the presence of 5 mol% catalyst. Hartwig and coworkers have
tested the effectiveness of homogeneous nickel(II)-N-heterocyclic carbene catalyst
for cleavage of C–O bonds of model diaryl ether compounds. Although this method
of cleaving ether linkages of small model compounds is effective for quantitative
conversions of model ethers with high yields (up to 99 %) of the corresponding
aromatic products under mild conditions (120 °C and 1 bar H2), this hydrogenolysis
process required high concentration of nickel catalyst (~20 mol%) [35]. Most
recently mechanistic insight for the cleavage of β-O-4 ether bonds of lignin model
compound (PhCH(OH)CH2-O-Ph) has been elucidated using Pd/C catalyst. Using
solvent and benzylic hydroxyl group (internal hydrogen) of the model compounds
as hydrogen source, the authors demonstrated excellent yields (>95 %) of the cor-
responding C–O bond cleaved and hydrogenolysis products with aromatic ring
intact. Control experiments using different substituents in model compounds, solvents,
6 Catalytic Hydrodeoxygenation of Lignin Model Compounds 127

Fig. 6.4 Proposed β-O-4


bond cleavage mechanism
(Reproduced from Ref.
[36] by permission of John
Wiley & Sons Ltd)

and deuterated analogues suggested that β-benzylic-H of the substrate facilitate


cleavage of the ether bonds [36]. The authors proposed the formation of
β-phenoxyalkyl palladium(II) hydride as an intermediate adduct via oxidative addi-
tion of the benzylic-H bond by Pd0 which is subsequently cleaved to the observed
products (Fig. 6.4).

6.6 Conclusions

Lignin is a natural biopolymer of biomass constituting aromatic building block


units that can be utilized as a sustainable feedstock for producing aromatic specialty
and commodity chemicals. Among several strategies emerged to date on lignin val-
orization, depolymerization via hydrogenolysis of lignin into monomers has been
the key strategy. However, this approach requires efficient catalysts for selective
depolymerization of lignin polymer into manageable stream of products for separa-
tion and purification. In this account, we have reviewed recent work on selective
catalytic depolymerization of wood lignin to form high-value aromatic phenolic
chemicals. Among several catalytic systems published to date for lignin valoriza-
tion, the Pd/C/Zn2+- and Ni/C-catalyzed depolymerization approaches, producing
dihydroeugenol and 4-propylsyringol with high selectivity, are the most appealing.
While both processes are efficient for producing methoxyphenols without hydroge-
nating the aromatic ring, the Pd/C/Zn2+ catalytic system is strategically advanta-
geous for commercial biorefinery processing because of high carbon efficiency of
the catalytic products and suitability of a wide range of biomass feedstock feasibil-
ity. This book chapter also analyzed the scope of the deoxygenation techniques for
upgrading of lignin-derived aromatic functionalized monomers and its successful
recent outcomes for accessing high-value deoxygenation products containing an
intact saturated and unsaturated aromatic ring. This analysis suggests that the cata-
lyst supports and reaction conditions (pressure and temperature) play a significant
128 B. Saha et al.

role on HDO pathways, besides the active metal sites. The noble metal catalysts
containing acidic support achieved high degree of deoxygenation followed by
hydrogenation of aromatic ring as well as side-chain unsaturated bonds, whereas
neutral (activated carbon) or basic support (hydrotalcite) tends to keep integrity of
the aromatic ring.

Acknowledgment This work was supported as part of the Center for Direct Catalytic Conversion
of Biomass to Biofuels (C3Bio), an Energy Frontier Research Center (EFRC) funded by the US
Department of Energy, Office of Science, Office of Basic Energy Sciences, Award Number
DE-SC0000997.

References

1. Boisen A, Christensen TB, Fu W, Gorbanev YY, Hansen TS, Jensen JS, Klitgaard SK, Pedersen
S, Riisager A, Staahlberg T, Woodley JM (2009) Process integration for the conversion of
glucose to 2,5-furandicarboxylic acid. Chem Eng Res Des 87:1318–1327
2. Nigam PS, Singh S (2011) Production of liquid biofuels from renewable resources. Prog
Energy Combust Sci 37:52–68
3. High-Value Opportunities for Lignin: Ready for liftoff, Frost & Sullivan report, (2014)
4. Li X, Ximenes E, Kim Y, Slininger M, Meilan R, Ladisch M, Chapple C (2010) Lignin mono-
mer composition affects Arabidopsis cell-wall degradability after liquid hot water pretreat-
ment. Biotechnol Biofuels 27:1–7
5. Rubin EM (2008) Genomics of cellulosic biofuels. Nature (London) 454:841–845
6. Brodeur G, Yau E, Badal K, Collier J, Ramachandran KB, Ramakrishnan1 S (2011) Chemical
and physicochemical pretreatment of lignocellulosic biomass: a review. Enzym Res. doi:http://
dx.doi.org/10.4061/2011/787532
7. El Hage R, Brosse N, Chrusciel L, Sanchez C, Sannigrahi P, Ragauskas A (2009)
Characterization of milled wood lignin and ethanol organosolv lignin from Miscanthus. Polym
Degrad Stab 94:1632–1638
8. Nimz HH (1974) Beech lignin – proposal of a constitutional scheme. Angew Chem Int Ed Engl
13:313–321
9. Dutta S, Wu KCW, Saha B (2014) Emerging strategies for breaking the 3D amorphous net-
work of lignin. Catal Sci Technol 4:3785–3799
10. Chatel G, Rogers RD (2014) Review: oxidation of lignin using ionic liquids-an innovative
strategy to produce renewable chemicals. ACS Sustain Chem Eng 2:322–339
11. Zakzeski J, Bruijnincx PCA, Jongerius AL, Weckhuysen BM (2010) The catalytic valorization
of lignin for the production of renewable chemicals. Chem Rev 110:3552–3599
12. Holladay J, White JF, Bozell JJ, Johnson D (2007) Top value-added chemicals from biomass.
DOE Report, p 2, 2007
13. Laskar DD, Yang B, Wang H, Lee J (2013) Pathways for biomass-derived lignin to hydrocar-
bon fuels. Biofuels Bioprod Biorefin 7:602–626
14. Ragauskas AJ, Beckham GT, Biddy MJ, Chandra R, Chen F, Davis MF, Davison BH, Dixon
RA, Gilna P, Keller M, Langan P, Naskar AK, Saddler JN, Tschaplinski TJ, Tuskan GA,
Wyman CE (2014) Lignin valorization: improving lignin processing in the biorefinery. Science
(Washington, DC, US) 344:709
15. Song Q, Wang F, Xu J (2012) Hydrogenolysis of lignosulfonate into phenols over heteroge-
neous nickel catalysts. Chem Commun (Camb) 48:7019–7021
16. Olcese RN, Lardier G, Bettahar M, Ghanbaja J, Fontana S, Carre V, Aubriet F, Petitjean D,
Dufour A (2013) Aromatic chemicals by iron-catalyzed hydrotreatment of lignin pyrolysis
vapor. ChemSusChem 6:1490–1499
6 Catalytic Hydrodeoxygenation of Lignin Model Compounds 129

17. Zhao C, Lercher JA (2012) Selective hydrodeoxygenation of lignin-derived phenolic mono-


mers and dimers to cycloalkanes on Pd/C and HZSM-5 catalysts. ChemCatChem 4:64–68
18. Pepper JM, Lee YW (1969) Lignin and related compounds. I. Comparative study of catalysts
for lignin hydrogenolysis. Can J Chem 47:723–727
19. Pepper JM, Lee YW (1970) Lignin and related compounds. II. Studies using ruthenium and
Raney nickel as catalysts for lignin hydrogenolysis. Can J Chem 48:477–479
20. Pepper JM, Steck WF, Swoboda R, Karapally JC (1966) Hydrogenation of lignin using nickel
and palladium catalysts. In: Martin J (ed) Lignin structure and reactions. Advances in
Chemistry, American Chemical Society, Washington DC, pp 238–248
21. Jarrell TM, Marcum CL, Sheng H, Owen BC, O’Lenick CJ, Maraun H, Bozell JJ, Kenttaemaa
HI (2014) Characterization of organosolv switchgrass lignin by using high performance liquid
chromatography/high resolution tandem mass spectrometry using hydroxide-doped negative-
ion mode electrospray ionization. Green Chem 16:2713–2727
22. Parsell TH, Yohe S, Degenstein J, Jarrell T, Klein I, Gencer E, Hewetson B, Hurt M, Kim JI,
Choudhari H, Saha B, Meilan R, Mosier N, Ribeiro F, Delgass WN, Chapple C, Kenttamaa HI,
Agrawal R, Abu-Omar MM (2014) A synergistic biorefinery based on catalytic conversion of
lignin prior to cellulose starting from lignocellulosic biomass. Green Chem 17:1492–1499
23. Song Q, Wang F, Cai J, Wang Y, Zhang J, Yu W, Xu J (2013) Lignin depolymerization (LDP)
in alcohol over nickel-based catalysts via a fragmentation-hydrogenolysis process. Energy
Environ Sci 6:994–1007
24. Galkin MV, Samec JSM (2014) Selective route to 2-propenyl aryls directly from wood by a
tandem organosolv and palladium-catalysed transfer hydrogenolysis. ChemSusChem
7:2154–2158
25. Mu W, Ben H, Ragauskas A, Deng Y (2013) Lignin pyrolysis components and upgrading –
technology review. Bioenerg Res 6:1183–1204
26. Jongerius AL, Bruijnincx PCA, Weckhuysen BM (2013) Liquid-phase reforming and hydro-
deoxygenation as a two-step route to aromatics from lignin. Green Chem 15:3049–3056
27. Selig M, Weiss N, Ji Y (2008) Enzymatic saccharification of lignocellulosic biomass: labora-
tory analytical procedure. National Renewable Energy Laboratory, Golden, Colorado
28. Kim Y, Ximenes E, Mosier NS, Ladisch MR (2011) Soluble inhibitors/deactivators of cellulase
enzymes from lignocellulosic biomass. Enzym Microb Technol 48:408–415
29. Parsell TH, Owen BC, Klein I, Jarrell TM, Marcum CL, Haupert LJ, Amundson LM,
Kenttaemaa HI, Ribeiro F, Miller JT, Abu-Omar MM (2013) Cleavage and hydrodeoxygen-
ation (HDO) of C-O bonds relevant to lignin conversion using Pd/Zn synergistic catalysis.
Chem Sci 4:806–813
30. Prasomsri T, Shetty M, Murugappan K, Roman-Leshkov Y (2014) Insights into the catalytic
activity and surface modification of MoO3 during the hydrodeoxygenation of lignin-derived
model compounds into aromatic hydrocarbons under low hydrogen pressures. Energy Environ
Sci 7:2660–2669
31. Jina S, Xiaoa Z, Lia C, Chena X, Wanga L, Xinga J, Lib W, Liang C (2014) Catalytic hydro-
deoxygenation of anisole as lignin model compound over supported nickel catalysts. Catal
Today 234:125–132
32. Deepa AK, Dhepe PL (2014) Function of metals and supports on the hydrodeoxygenation of
phenolic compounds. ChemPlusChem 79:1573–1583
33. Ohta H, Feng B, Kobayashi H, Hara K, Fukuoka A (2014) Selective hydrodeoxygenation of
lignin-related 4-propylphenol into n-propylbenzene in water by Pt-Re/ZrO2 catalysts. Catal
Today 234:139–144
34. Harms RG, Markovits IIE, Drees M, Herrmann WA, Cokoja M, Kuehn FE (2014) Cleavage of
C-O bonds in lignin model compounds catalyzed by methyldioxorhenium in homogeneous
phase. ChemSusChem 7:429–434
35. Sergeev AG, Hartwig JF (2011) Selective, nickel-catalyzed hydrogenolysis of aryl ethers.
Science 332:439–442
36. Zhou X, Mitra J, Rauchfuss TB (2014) Lignol cleavage by Pd/C under mild conditions and
without hydrogen: a role for benzylic C-H activation? ChemSusChem 7:1623–1626

You might also like