You are on page 1of 25

polymers

Review
Catalytic Oxidation of Lignin in Solvent Systems for
Production of Renewable Chemicals: A Review
Chongbo Cheng 1 , Jinzhi Wang 1 , Dekui Shen 1, *, Jiangtao Xue 2 , Sipian Guan 2 , Sai Gu 3 and
Kai Hong Luo 4
1 Key Lab of Thermal Energy Conversion and Control of MoE, Southeast University, Nanjing 210096, China;
220130473@seu.edu.cn (C.C.); 220160516@seu.edu.cn (J.W.)
2 Jiangsu Frontier Electric Power Technology Co., Ltd., Nanjing 211102, China; ludong901122@sina.com (J.X.);
Parallel79@163.com (S.G.)
3 Department of Chemical and Process Engineering, Faculty of Engineering and Physical Sciences,
University of Surrey, Surrey GU2 7XH, UK; sai.gu@surrey.ac.uk
4 Department of Mechanical Engineering, University College London, London WC1E 7JE, UK; k.luo@ucl.ac.uk
* Correspondence: 101011398@seu.edu.cn; Tel.: +86-138-5170-6572

Academic Editor: Wolfgang Gindl-Altmutter


Received: 5 May 2017; Accepted: 16 June 2017; Published: 21 June 2017

Abstract: Lignin as the most abundant source of aromatic chemicals in nature has attracted a great
deal of attention in both academia and industry. Solvolysis is one of the promising methods to
convert lignin to a number of petroleum-based aromatic chemicals. The process involving the
depolymerization of the lignin macromolecule and repolymerization of fragments is complicated
influenced by heating methods, reaction conditions, presence of a catalyst and solvent systems.
Recently, numerous investigations attempted unveiling the inherent mechanism of this process in
order to promote the production of valuable aromatics. Oxidative solvolysis of lignin can produce
a number of the functionalized monomeric or oligomeric chemicals. A number of research groups
should be greatly appreciated with regard to their contributions on the following two concerns:
(1) the cracking mechanism of inter-unit linkages during the oxidative solvolysis of lignin; and (2) the
development of novel catalysts for oxidative solvolysis of lignin and their performance. Investigations
on lignin oxidative solvolysis are extensively overviewed in this work, concerning the above issues
and the way-forward for lignin refinery.

Keywords: biomass; lignin depolymerization; oxidative solvolysis; catalysis

1. Introduction
Biomass is an abundant renewable feedstock for the production of fuels, chemicals, and energy.
Nowadays, about 10% of the world’s primary energy is from biomass [1]. Biomass is primarily
used to generate heat and power, and it stands as the fourth largest source of energy in the word
(after oil, coal, and natural gas) [2]. Fuels produced from biomass-refinery processes are replacing the
fossil-based fuels in the daily life [3]. However, the use of biomass components and derivatives as the
raw feedstocks for the chemical industry is still in the fledging stages [4–8]. With the pressure of the
price of fossil fuels and the high cost of biorefinery processes of biomass, more and more attention has
been paid on the production of value-added chemicals from biomass.
Biomass consists of three valuable fractions: 38–50% cellulose composed of semi-crystalline
polysaccharide, 23–32% hemicellulose composed of amorphous multicomponent polysaccharide,
and 15–25% lignin composed of amorphous phenylpropanoid polymer [9,10]. Hemicellulose has
important applications for biofuel production and for the generation of valuable chemical intermediates
(e.g., furfural) [11]; cellulose can be decomposed into valuable products such as biofuels and platform

Polymers 2017, 9, 240; doi:10.3390/polym9060240 www.mdpi.com/journal/polymers


Polymers 2017, 9, 240 2 of 25

chemicals (including levulinic and formic acids, gammavalerolactone and derived products) [12];
and lignin is the most underutilized fraction, only approximately 2% of the lignin available from the
pulp and paper industry is commercially used while the remainder is burned as a low value fuel [13].
There is a general lack of efficient processes for the utilization of lignin regarded as the most important
renewable source of aromatics on the planet [14]. It becomes highly desirable to develop efficient
technologies for the conversion of lignin to aromatic chemicals [15–17].
In the past decades, several strategies have been employed for lignin valorization, such as
pyrolysis and gasification [18–23]. Lignin valorization in solvent systems for the production
of renewable aromatic chemicals has attracted ever-increasing attention during recent years.
The methodology for lignin depolymerization (LDP) in solvent systems can be categorized as
hydrolysis, hydrogenolysis, oxidation and a two-step LDP. Lange et al. [3] and Chatel et al. [24,25]
have published reviews on catalytic oxidation for lignin valorization into value-added chemicals.
Some relevant reviews on bond cleavage mechanisms during lignin depolymerization have been
published [26] and the separation and purification processes in lignin depolymerization have
been reviewed recently [27]. Catalysis is regarded as a key factor to fulfill the promise of lignin
depolymerization to specific products. Catalytic (homogeneous/heterogeneous) conversion of lignin
for production of aromatic chemicals has also been reviewed in recent years [1,28–31]. Lignin oxidative
solvolysis has been widely applied in the bleaching process for pulping industry due to its economic
feasibility [32,33]. Phenolic products with low molecular weight were detected as the primary products
from the oxidative solvolysis of lignin [33]. Organometallic catalyzed oxidation, biomimetic oxidation
and enzyme-based oxidation have been studied [3,34,35], aiming at the cleavage of the designated
inter-unit linkages [27]. Fundamental research on understanding and controlling the oxidative process
is essential for promoting the production of specific valuable compounds [33].
The chemical structure and morphological properties of lignin are briefly introduced in this work,
regarding the typical inter-unit linkages and lignin-related model compounds. Catalytic oxidation
of lignin and lignin model compounds in solvent systems are then comprehensively overviewed,
concentrating on the cracking of the inter-unit linkages and the effect of catalyst and solvent on the
production of renewable aromatic chemicals.

2. Chemical Structure of Lignin


The roles of lignin in plants are to afford structural strength, protect the plants from degradation
and enable the water conduction system connecting roots with leaves. Lignin is composed of
p-hydroxyphenyl (H), guaiacyl (G) and syringyl (S) units linked by C–O or C–C bonds, which exhibit a
complex three-dimensional amorphous structure. The content and chemical structure of lignin are
primarily influenced by its sources and isolation methods. The chemical structure for lignin from
different sources is different in terms of proportion of p-coumaryl-alcohol, coniferyl-alcohol, and
sinapyl-alcohol (C9 unit) (Figure 1a). More than 95% of guaiacyl units (that derived from coniferyl
alcohol) are estimated to be in softwood lignin, while the proportions of coniferyl-alcohols and
sinapyl-alcohols are approximately equal in hardwood lignin. Herbaceous biomass (e.g., straw) lignin
and woody lignin are of different chemical structures because of the extra content of p-hydroxyphenyl
units and significant amounts of hydroxycinnamic acid (mainly p-coumaric acid). The proportions of
p-coumaryl-, coniferyl- and sinapyl-alcohols in lignin depend on sources of lignin.
Lignin units are linked by either C–O or C–C bonds. More than two-thirds of the total linkages
are C–O bonds while the others are C–C bonds in native lignin [1,36]. The carbon atoms in aromatic
moieties are numbered as 1–6 and those in the aliphatic side chains of the units are labeled as α, β, and
γ to distinguish various types of linkages between two structural units. The β-position is generally
favored by coupling other units, leading to β-O-4, β-β, β-1 and β-5 becoming main linkages among
the units of lignin [28,37–39]. The most common linkage in lignin is the β-O-4 ether bond, accounting
for 45 to 48% in native lignin [40,41]. Other linkages include α-O-4, 4-O-5, 5-5, etc. The linkages in
Polymers 2017, 9, 240 3 of 25

Polymers 2017, 9, 240 3 of 25

lignin are illustrated in Figure 1b, while the proportion of these linkages and the functional groups
The linkages in lignin are illustrated in Figure 1b, while the proportion of these linkages and the
depends on the lignin
functional groupssources
depends and extraction
on the methods
lignin sources [28,42,43].
and extraction methods [28,42,43].

Figure 1. (a) Schematic of the structure of monomer C9 units; and (b) schematic of the proposed
Figure 1. (a) Schematic of the structure of monomer C9 units; and (b) schematic of the proposed lignin
lignin structure with several linkages (B1: β-O-4, B2: α-O-4, B3: β-5, B4: β-1, B5: 5-5, B6: β-β, B7:
structure with several linkages (B1: β-O-4, B2: α-O-4, B3: β-5, B4: β-1, B5: 5-5, B6: β-β, B7: 4-O-5) [44],
4-O-5) [44], reproduced with permission from Springer.
reproduced with permission from Springer.
Lignin-related model compounds are widely employed for better understanding of the
cracking
Lignin-related mechanism
modelofcompounds
the specific bonds in lignin.
are widely The mostfor
employed widely-reported model compounds
better understanding of the cracking
and linkages containing aromatic or aliphatic ethers are shown in Figure 2. The synthetic methods
mechanism of the specific bonds in lignin. The most widely-reported model compounds and linkages
for making β-O-4 model compounds [45], β-5 model compounds [46], β-1 model compounds [47,48]
containing aromatic or aliphatic ethers are shown in Figure 2. The synthetic methods for making β-O-4
and β-β model compounds [49] have been reported in the literature. However, due to the lack of
model complexity
compounds [45], models
of these β-5 model compounds
compared with lignin [46], β-1
itself, themodel compounds
application of model [47,48]
compound andinβ-β the model
compounds [49] have
development been reported
of selective in therestricted
LDP has been literature.
[46]. However,
Unlike a wide due to the
variety of lack
trimer,oftetramer,
complexity of
hexamercompared
these models and polymer lignin
with model
lignin compounds
itself, containingof
the application only β-O-4compound
model linkage, Ouyang in theet development
al. [50]
synthesized a lignin model compound composed of three guaiacyl
of selective LDP has been restricted [46]. Unlike a wide variety of trimer, tetramer, hexamer structural units linked by α-O-4 and
and β-O-4 ether bonds, via a three-step sequence method assisted with microwave irradiation.
polymer lignin model compounds containing only β-O-4 linkage, Ouyang et al. [50] synthesized a
Sheldrake et al. [46] also reported a flexible, efficient and convergent preparation method for
lignin model compound composed of three guaiacyl structural units linked by α-O-4 and β-O-4 ether
synthesizing model lignin hexamers and octamers linked by β-O-4, 5-5′ and β-5. These lignin model
bonds, compounds
via a three-step sequence
can reflect method
some natural assisted with
information microwave
of lignin, giving a irradiation. Sheldrake
better understanding et al. [46]
of the
also reported
inherent mechanism of lignin depolymerization processes. Other novel approaches for synthesizing lignin
a flexible, efficient and convergent preparation method for synthesizing model
hexamersthe model compounds
and octamers that are
linked bymore
β-O-4, 5-50 inand
accurate mimicking
β-5. Thesecharacteristics of lignin
lignin model are still required.
compounds can reflect
some natural information of lignin, giving a converting
Solvolysis is a promising method for lignin to various
better understanding valuable
of the aromatic
inherent chemicals.
mechanism of lignin
The lignin macromolecule depolymerization and fragments repolymerization are involved in this
depolymerization processes. Other novel approaches for synthesizing the model compounds that are
process, which is conducted by using different catalysts, solvent systems, experimental conditions
more accurate in mimicking characteristics of lignin are still required.
and even heating methods. The depolymerization of lignin to renewable aromatic chemicals is
Solvolysis
difficult due is a promising
to the thermally method for converting
stable, amorphous lignin to various
and complex valuable
structure aromatic
of lignin. chemicals.
Lignin
The lignin macromolecule
depolymerization often depolymerization and fragments
necessitates harsh conditions to break therepolymerization
stable structure ofare involved
lignin. Under in this
process,such
whichconditions, the undesired
is conducted by usingrepolymerization of fragments
different catalysts, solvent is systems,
favored, resulting in a decrease
experimental of
conditions and
productmethods.
even heating yield. The The target linkages of lignin might
depolymerization be hard
of lignin to access by
to renewable heterogeneous
aromatic chemicals catalysts
is difficult
due to because of the steric hindrance of the lignin amorphous backbone. Although the target linkages of
the thermally stable, amorphous and complex structure of lignin. Lignin depolymerization
lignin could access by homogeneous catalysts, it is difficult to separate homogeneous catalysts from
often necessitates harsh conditions to break the stable structure of lignin. Under such conditions, the
the liquid products. The variability of a large number of substructural units in lignin would result in
undesired repolymerization
different catalytic reaction of rates
fragments is favored,
at different resulting
sites. A number of in a decrease
lignin of product
sources from different yield. The target
isolation
linkages of lignin
methods might
should be be hard toconcerning
examined access bythe heterogeneous catalysts
selection of feedstock for because
LDP. of the steric hindrance of
the lignin amorphous backbone. Although the target linkages of lignin could access by homogeneous
catalysts, it is difficult to separate homogeneous catalysts from the liquid products. The variability
of a large number of substructural units in lignin would result in different catalytic reaction rates
at different sites. A number of lignin sources from different isolation methods should be examined
concerning the selection of feedstock for LDP.
Polymers 2017, 9, 240 4 of 25
Polymers 2017, 9, 240 4 of 25

Figure 2. Lignin-related model compounds with different inter-unit linkages and functional groups.
Figure 2. Lignin-related model compounds with different inter-unit linkages and functional groups.

2.1. Kraft Lignin


2.1. Kraft Lignin
Kraft pulping is the predominant chemical pulping process around the world. This process is
Kraft pulping
directed is the predominant
at temperatures in a range ofchemical
150–180 pulping process
°C for around two around
hours the
andworld.
at highThis process
pH with the is
directed at temperatures in a range of 150–180 ◦ C for around two hours and at high pH with the
presence of considerable amounts of aqueous sulfide, sulfhydryl and polysulfide ions. Thus, the
presence
native of considerable
lignin will undergoamounts of aqueous
dramatic chemical sulfide, sulfhydryl
and structural and polysulfide
changes due to these ions. Thus,
severe the native
conditions,
lignin
which makes it difficult to be depolymerized to fragments in the subsequent processes. which
will undergo dramatic chemical and structural changes due to these severe conditions, The
makes it difficult
percentage to bein
of lignin depolymerized
the black liquortogenerated
fragments byin thepulping
kraft subsequentrangesprocesses.
from 29 toThe
45%percentage
for cook ofof
lignin
paperin grade
the black
andliquor
8 to 16%generated
for cook by kraftgrade
of liner pulping [51]ranges
and thefrom 29 to 45%
percentage for cook
of kraft blackof paper
liquor grade
lignin
andextracted
8 to 16%fromfor cook of linerisgrade
hardwood [51]lower
generally and the percentage
than of kraft black liquor lignin extracted from
that of softwood.
hardwood is generally lower than that of softwood.
2.2. Sulfite Lignin
2.2. Sulfite Lignin
As kraft pulping becomes the dominant pulping process all around world, just about 10% of the
pulp is produced
As kraft pulping bybecomes
sulfite pulping processpulping
the dominant now [52]. Sulfiteall
process pulping
around is world,
conducted
just under
about pH
10%2–12,
of the
pulp is produced by sulfite pulping process now [52]. Sulfite pulping is conducted under pH are
depending on the cationic composition of the pulping liquor. Nearly all sulfite pulping processes 2–12,
acidic, and
depending on calcium or magnesium
the cationic composition (Caof2+/Mg2+) are used as the counterions, while, in higher pH
the pulping liquor. Nearly all sulfite pulping processes
areenvironments, sodium or
acidic, and calcium or ammonium
magnesiumare (Ca usually
2+ /Mgused instead
2+ ) are usedofasCa/Mg. Due to the sulfite
the counterions, while,presence,
in higher
lignin is rendered soluble throughout the full range of pH;
pH environments, sodium or ammonium are usually used instead of Ca/Mg. Due to the it cannot simply be isolated bysulfite
pH
Polymers 2017, 9, 240 5 of 25

presence, lignin is rendered soluble throughout the full range of pH; it cannot simply be isolated by
pH adjustment. As mixtures with only 70–75% actually lignin, commercial sulfite lignin shows wide
molecular weight profiles and variable grades of sulfonation, making it difficult to be applied in lignin
depolymerization. The average molecular weight of sulfite lignin from softwood is notably higher that
of the sulfite lignin from hardwood.

2.3. Organosolv Lignin


Organosolv lignin, derived from organosolv pulping process, is designed to dissolve lignin
from plant cell walls by using a high proportion of an organic solvent (e.g., dioxane) in the cooking
liquor [53]. Compared to kraft and sulfite pulping, organosolv pulping is a more environmentally
benign pathway. Organosolv lignin reserves a large extent of its original structure in the form of
original inter-unit linkages (e.g., β-O-4 linkage), which can be used as an appropriate lignin sample for
subsequent conversion and valorization.

2.4. Pyrolytic Lignin


Pyrolytic lignin has unusual structural properties featured by its C8 basic unit skeleton and lower
average molecular weight [54,55]. The average molecular weight of the pyrolytic lignin was reported
to be within a range of 650–1300 Da [55]. Pyrolytic lignin is composed of some thermally ejected
oligomers and the side chains in pyrolytic lignin are primarily in the form of inter-unit C–C linkages,
leading to the difficulty of its application in LDP.

2.5. Dilute Acid Lignin


Diluted acids (e.g., H2 SO4 and HCl) pulping have been applied in all kinds of lignocellulosic
materials. Diluted acid pulping is generally conducted at a temperature of >160 ◦ C with the presence
of mineral acids as a catalyst. The C–O linkages as well as ester linkages are partially cracked during
the diluted acid pulping process. This process may also undermine the three-dimensional amorphous
structure in lignin and produce small lignin fragments with the increased proportion of hydroxyl
groups [56,57]. Thus, lignin should be removed under benign conditions if it is considered as a
feedstock for further chemical production.

2.6. Steam Explosion Lignin


The steam explosion pulping process is a fast processing method for lignocellulose, which releases
single biomass components via steam impregnation under pressure and then the fast release of
pressure [51]. For example, the disposal of wood or bagasse is conducted under high-pressure steam
at a short contact time. More than 90% of lignin in hardwood can be recovered from the steam
explosion pulping process. The properties of steam explosion lignin are similar to that of organosolv
lignin, showing both a higher solubility and a lower molecular weight in an organic solvent. It is also
composed of phenolic oligomers with 3 to 12 benzene rings. As a result, steam explosion lignin has
potential for producing renewable chemicals through LDP.

3. Oxidative Solvolysis of Lignin


Oxidative solvolysis of lignin is a widely used and efficient strategy for producing multiple
aromatic monomers. The oxidative solvolysis of lignin reaction includes the cracking of C–O bonds,
C–C bonds in lignin. Aromatic aldehydes and carboxylic acids are the main products from oxidative
solvolysis of lignin with different kinds of homogeneous catalysts. O2 , H2 O2 , metal oxides and
nitrobenzene have been considered as effective oxidation agents. In this section, the effect of catalyst
and solvent on the production of valuable aromatic chemicals in the catalytic oxidation of lignin process
will be discussed. Research on oxidative solvolysis of raw lignin (Kraft lignin, Organosolv lignin, etc.)
to the aromatic products is summarized in Table 1.
Polymers 2017, 9, 240 6 of 25

Table 1. Summary of oxidative depolymerization of the raw lignin material.

Entry Feedstock Conditions Solvent Catalyst Products Yield Ref.


Acetonitrile-THF or ethyl Vanadium complexes bearing Schiff Monophenolic compounds (vanillin, 0.78 wt %, 0.67 wt %, 0.59 wt %
1 Organosolv lignin (from M. giganteus) 80 ◦ C, 24 h, air [58]
acetate-THF base ligands syringic acid, syringaldehyde) (Catalyst = Complex 3)
Vanadium complexes and other
2 Organosolv lignin 100 ◦ C, 18 h, 0.8 MPa synthetic air Ethyl acetate Bio-oil Mw = 575 Da (Catalyst = Complex 2) [59]
organometallic catalysts
3 Organosolv lignin and Kraft lignin 100 ◦ C, 8 h, H2 O2 DMSO and acetic acid {Fe-DABCO} Bio-oil / [60]
Co(salen) supported on
4 Organosolv lignin 80 ◦ C, 24 h, air Acetonitrile-THF Vanillin (main) 3067 g [61]
graphene oxide
V(acac)3 and Cu(NO3 )2 ·3H2 O or
5 Organosolv lignin and Kraft lignin 135 ◦ C, 40 h, 1.0 MPa O2 Pyridine Bio-oil Mw = 300 Da [62]
HTc-Cu-V
Hydrolytic sugar cane lignin and red MTO or poly(4-vinylpyridine)/MTO
6 25 ◦ C, 24 h, H2 O2 Acetic acid Bio-oil / [63]
spruce kraft lignin or polystyrene/MTO
H2 O, tert-butyl Nitrogen-containing graphene
7 Organosolv ligin (from birch wood) 140 ◦ C, 24 h, 0.1 MPa O2 Bio-oil 45.8 wt % [64]
hydroperoxide (TBHP) material (LCN)
Formic acid, acetic acid, succinic acid,
8 Alkali lignin 175–225 ◦ C, 0–1 h, 0.5–1.5 MPa O2 Water NaOH 44.0 wt % [65]
oxalic acid, glutaconic acid
Water/methanol/1,4-dioxane/
9 Wheat alkali lignin 150 ◦ C, 1 h, H2 O2 CuO, Fe2 (SO4 )3 and NaOH Monophenolic compounds 17.92 wt % (in methanol-water) [66]
tetrahydrofuran/ethanol
Copper-phenanthroline complex
10 Wood lignin from Loblolly pine 80 ◦ C, 24 h, 0.27/1.24 MPa O2 Methanol Vanillic acid, vanillin 3.5 wt %, 12.6 wt % [67]
and NaOH
11 Kraft lignin 170 ◦ C, 0.3 h, 0.5 MPa O2 Methanol-water and H2 SO4 H3 PMo12 O40 Vanillin, methyl vanillate 5.2 wt % [68]
12 Kraft lignin 170 ◦ C, 0.3 h, 1.0 MPa O2 Methanol-water and H2 SO4 H3 PMo12 O40 Vanillin, methyl vanillate 4.6 wt %, 4.2 wt % [69]
CuSO4 ; FeCl3 ;
13 Kraft lignin 170 ◦ C, 1 h, 1.0 MPa O2 Methanol-water and H2 SO4 Vanillin, methyl vanillate 6.3 wt % [70]
CuCl2 ; CoCl2
14 Kraft lignin 45 ◦ C, 1 h, H2 O2 Acetone-water Metal salt catalysts Vanillin-based monomers 0.51 wt % [71]
Vanillin, vanillin acid, 0.99 wt %, 2.91 wt %, 2.52 wt %,
15 Organosolv lignin 180 ◦ C, 2 h, 13.8 MPa air Acetic Acid-water Co/Mn/Zr/Br mixture [72]
syringaldehyde, syringic acid 4.51 wt %
16 Organosolv beech wood lignin 200W, 5–30 min, H2 O2 NaOH solution La/SBA-15 Vanillin, syringaldehyde 9.94 wt %, 15.66 wt %. [73]
Vanillin, guaiacol,
17 Organosolv lignin 185 ◦ C, 24 h, 0.1 MPa O2 Methanol Pd/CeO2 5.2 wt %, 0.87 wt %, 2.4 wt % [74]
4-hydroxybenzaldehyde

◦ C, Vanillin, p-hydroxybenzyl
18 Enzymatic hydrolysis lignin 120 0–3 h, 0.5 MPa O2 NaOH solution LaMnO3 4.32 wt %, 2.03 wt %, 9.33 wt % [75]
aldehyde, syringaldehyde

◦ C, Vanillin, p-hydroxybenzyl
19 Enzymatic hydrolysis lignin 120 0–3 h, 0.5 MPa O2 NaOH solution LaCoO3 4.55 wt %, 2.23 wt %, 9.99 wt % [76]
aldehyde, syringaldehyde
Polymers 2017, 9, x FOR PEER REVIEW 7 of 25
Polymers 2017, 9, 240 7 of 25

3.1. Effects of Catalysts


3.1. Effects of Catalysts
Catalysts are involved in most cases of the oxidative depolymerization of lignin to produce
Catalysts
particular are involved
products and even in most cases
perform of the oxidative
chemoselective oxidation depolymerization
of a particular linkage of lignin to produce
in lignin [77,78].
particular products
The catalysts widely andused
evenfor perform chemoselective
the oxidative oxidation
solvolysis of ligninof a can
particular linkage
be divided in lignin
into [77,78].
five groups:
The catalysts widely
organometallic used
catalysts, for the oxidativecatalysts,
metal-free-organic solvolysis of lignincatalysts,
acid/base can be divided
metal salts intocatalysts,
five groups: and
organometallic
zeolite catalystscatalysts, metal-free-organic
(heterogeneous catalyst). catalysts, acid/base catalysts, metal salts catalysts, and
zeolite catalysts (heterogeneous catalyst).
3.1.1. Organometallic Catalysts
3.1.1. Organometallic Catalysts
Several studies have reported on the oxidative solvolysis of raw lignin by using an
Several studies
organometallic haveChan
catalyst. reportedet al.on[58]
the demonstrated
oxidative solvolysis
that aofseries
raw lignin by usingcomplexes
of vanadium an organometallic
bearing
catalyst. Chan et al. [58] demonstrated that a series of vanadium
Schiff base ligands (e.g., Complex 3 in Figure 3) could be used to degrade organosolv complexes bearing Schifflignin
base
ligands (e.g.,
(extracted from Complex
Miscanthus 3 in Figure
giganteus 3) could
using be used toacetone,
dioxane, degrade andorganosolv
ethanol) intolignin (extracted
phenolic from
products
Miscanthus giganteus using dioxane, acetone, and ethanol) into phenolic products
in acetonitrile–THF or ethyl acetate–THF. After 24 h of reaction at 80 °C under air, vanillin, syringic in acetonitrile–THF
or ethyl
acid acetate–THF. After
and syringaldehyde were 24 thehmost of reaction
common at 80 ◦ C under
degradation air, vanillin,
products, resultingsyringic
from theacid and
guaiacyl
syringaldehyde were the most common degradation products, resulting
and syringyl units of lignin, respectively (Table 1, Entry 1). Pretreatment and isolation methods from the guaiacyl and
syringyl
could units of lignin,
remarkably alter respectively
the structure(Table 1, Entry
of lignin and1).thus
Pretreatment
affect the and isolation
reactivity of methods
lignin towardcould
remarkably
catalysts. altervanadium
Then, the structure of lignin
Complexes 1, 2and
andthus affect3)the
4 (Figure and reactivity of lignin toward
other organometallic catalysts.
catalysts such
Then, vanadium Complexes 1, 2 and 4 (Figure 3) and other organometallic
as 4-acetamido-TEMPO/acid (TEMPO = 2,2,6,6-tetramethylpiperidine-1-oxyl) and (salen)cobalt(II) catalysts such as
4-acetamido-TEMPO/acid (TEMPO = 2,2,6,6-tetramethylpiperidine-1-oxyl)
were also used in the catalytic oxidation of organosolv lignin [59]. Though most catalysts oxidized and (salen)cobalt(II) were
also lignin
the used in the catalytic
extracts, oxidation of organosolv
bis(8-oxy-quinoline) oxovanadium lignin
and[59]. Though most catalysts
bis(phenolate)pyridine oxidized
oxovanadium
the lignin extracts, bis(8-oxy-quinoline) oxovanadium and bis(phenolate)pyridine
catalyst, Complex 2 and Complex 5 (Figure 3), respectively, exhibited the best performance for oxovanadium
catalyst, Complex 2 and Complex
depolymerization/oxidation 5 (Figure
of organosolv 3), respectively,
lignin. After 18 h of exhibited
reactionthe at 100best°C performance
in 0.8 MPa for of
depolymerization/oxidation of organosolv lignin. After 18 h of reaction at 100 ◦ C in 0.8 MPa of
synthetic air (8% O2 in Ar), the best depolymerization performance was exhibited by Complex 2,
synthetic
which air (8%
gave the Olargest
2 in Ar),molecular
the best depolymerization
weight (MW) shift performance was exhibited
and a considerable by Complex
increase in the 2, which
lower
gave the largest
molecular weight molecular
products weight
(Table(M ) shift2).
1,WEntry and a considerable
Iron complexes could increase in be
also thesuitable
lower molecular
for oxidation weight of
products (Table 1, Entry 2). Iron complexes could also be suitable
lignin. {Fe-DABCO} (DABCO = 1,4-diazabicyclo[2.2.2]-octane) catalyst was presented for the for oxidation of lignin. {Fe-DABCO}
(DABCO
cleavage of= 1,4-diazabicyclo[2.2.2]-octane)
β-O-4 linkages in lignin in thecatalyst presence wasofpresented
H2O2 [60].for the cleavage
After of β-O-4
8 h of reaction linkages
at 100 °C, thein
ligninmaximum
in the presence ◦
mass in theof H2 Osample
lignin 2 [60]. After 8 h of reaction
had shifted from 3200 at 100
Da toC,around
the mass 2500maximum
Da (Tablein1,the lignin
Entry 3).
sample had shifted from 3200 Da to around 2500 Da (Table 1, Entry
GPC analysis showed that the mass maximum had shifted to lower masses in all lignin samples, 3). GPC analysis showed that the
mass maximum
providing had for
evidence shifted to lower masses
the formation of lowinmolecular
all lignin samples, providingInevidence
weight products. addition,for allthe
of formation
the β-O-4
of low molecular weight products. In addition, all of the β-O-4 linkages
linkages and resinol structures were degraded independent of the lignin pretreatment and isolation and resinol structures were
degraded
conditions. independent of the lignin pretreatment and isolation conditions.

Figure 3. Schematics of the catalyst: vanadium Complexes 1–5.


Figure 3. Schematics of the catalyst: vanadium Complexes 1–5.
Polymers 2017, 9, 240 8 of 25

Co(salen)/oxidant has been proven to be an efficient catalytic system for oxidation of phenolic
and nonphenolic β-O-4 aryl ether lignin model compounds [79,80]. To improve complexes’ stability,
selectivity, and recyclability, Co salen complexes were immobilized on graphene oxide (GO) (Co–GO)
for the oxidation of lignin [61]. After 24 h of reaction at 80 ◦ C in air, organosolv lignin was converted
to vanillin (the main product), vanillic acid, 1-(4-hydroxy-3-methoxyphenyl) ethanol, and other
monomeric phenolic compounds (Table 1, Entry 4). The oxidation of aliphatic groups resulted in the
formation of carbonyl compounds, cleavage of β-O-4 bonds, demethylation of methoxy groups and
formation of phenolic OH groups.
Recently, the first transition metal system using O2 as an oxidant for cleaving the resinol structure
was reported [62]: An oxidative solvolysis of lignin with inexpensive transition-metal-containing
hydrotalcites (HTc–Cu–V) or combinations of V(acac)3 (acac = acetylacetonate) and Cu(NO3 )2 ·3H2 O
as catalysts. As the homogeneous copper and vanadium species from V(acac)3 /Cu(NO3 )2 ·3H2 O and
the leached HTc–Cu–V were most probably very similar in nature, after 40 h of reaction at 135 ◦ C,
the oxidative solvolysis of lignin afforded products of a mass of approximately 300 Da by using
copper-vanadium double-layered hydrotalcites or combined V(acac)3 (acac = acetylacetonate) and
Cu(NO3 )2 ·3H2 O as catalysts (Table 1, Entry 5).
Methyltrioxo rhenium (MTO) acts as an active and efficient catalyst for the oxidation of phenolic
and non-phenolic lignin model compounds representing the main bond patterns in native lignin
resulting in the side-chain oxidation and aromatic ring cracking reaction [81]. Crestini et al. [63]
developed immobilized MTO catalysts such as poly(4-vinylpyridine)/MTO and polystyrene/MTO as
an easy recovery, reuse and low environmental impact catalyst, for the oxidation of lignin (Table 1,
Entry 6). Different reaction pathways occurred between soluble MTO and immobilized MTO catalysts
in the oxidative solvolysis of lignin: compared to homogeneous MTO, immobilized MTO catalysts
with lower Lewis acidity might direct their reactivity toward aliphatic C–H insertion and Dakin-like
reactions instead of oxidizing the aromatic ring. Consequently, these differences gave rise to the
final product with a higher amount of free phenolic guaiacyl groups in the presence of immobilized
MTO catalysts.

3.1.2. Metal-Free-Organic Catalysts


Some metal-free-organic catalysts have been identified as effective catalysts for the oxidation of
β-O-4 and α-O-4 types of lignin model compounds. Gao et al. [64] used nitrogen-containing graphene
material (LCN) to convert birch wood lignin into depolymerized products in water and tert-butyl
hydroperoxide solvent. LCN exhibited a similar catalytic behavior in the oxidation of lignin model
compounds and birch wood lignin. The catalytic oxidation of benzylic C–H and C–OH bonds into
carbonyl groups broke the Cα –Cβ and Cα –O bonds, and further catalyzed the decomposition of some
unstable aromatic compounds. Finally, 45.8 wt % bio-oil could be obtained from birch wood lignin
after 24 h of reaction at 140 ◦ C in the presence of 0.1 MPa O2 (Table 1, Entry 7).

3.1.3. Acid/Base Catalysts


NaOH and KOH are utilized most frequently in the alkaline oxidation of lignin. Ouyang et al. [82]
reported the oxidative degradation of soda lignin assisted by microwave on NaOH catalyst (pH = 11).
Similarly, Wu et al. [83] reported the oxidative degradation of different lignin samples in KOH with
H2 O2 or K2 S2 O8 to methanol, formate, carbonate, and oxalate. Even at near-ambient temperatures
(60 ◦ C), >90% consumption of lignin was observed after 24 h. The microwave irradiation efficiently
helped the degradation of lignin to produce high molecular weight compounds. The origin of the
lignin, reaction temperature, oxidant concentration, and pH value also had impacts on the conversion
efficiency. For example, higher reaction temperature and O2 concentration facilitated the yield of
the final products but also gave rise to the repolymerization of lignin fragments. Maziero et al. [84]
also reported that sugarcane bagasse lignin oxidized with 9.1% H2 O2 (m/v) at pH = 13.3 had the
highest fragmentation, oxidation degree and stability. The oxidation of alkali lignin was performed on
Polymers 2017, 9, 240 9 of 25

NaOH catalyst (pH = 11) in the presence of O2 [65]. After 20 min of reaction, the maximum total yield
of products (formic acid, acetic acid, succinic acid, oxalic acid, and glutaconic acid) of 44 wt % was
obtained at 225 ◦ C and 1 MPa O2 (Table 1, Entry 8).
Ouyang et al. [66] reported the oxidative solvolysis of lignin by using H2 O2 as the oxidant and
combination of CuO, Fe2 (SO4 )3 and NaOH as catalysts. Cu2+ could result in the removal of side chains
from basic units of lignin forming some phenolic compounds while with OOH radicals generated by
dissociation of H2 O2 , Fe3+ could form new reactive intermediates [85], facilitating oxidative solvolysis
of lignin. The yield of monophenolic compounds with a higher amount of syringyl unit compounds
(syringaldehyde, syringic acid, acetosyringone) reached 17.92 wt % in methanol/water (1:1 (v/v))
solvent at 150 ◦ C after 1 h reaction (Table 1, Entry 9). Wood lignin from loblolly pine was oxidized by
1,10-phenanthroline and copper (II) sulfate pentahydrate and NaOH catalysts in methanol solvent [67].
Finally, 3.5 wt % vanillic acid and 12.6 wt % vanillin could be obtained after 24 h of reaction at 80 ◦ C in
the presence of 0.27 MPa O2 (Table 1, Entry 10).

3.1.4. Metal Salt Catalysts


Polyoxometalates (POMs) consists of a number of metal oxygen cluster anions. They are soluble in
both water and organic solvents, and its redox potential is high enough to oxidize lignin basic units yet
low enough to be reoxidized by O2 . A series of POMs are capable of selectively degrading the residual
lignin and, in turn, fully converting residual lignin into CO2 and H2 O by wet oxidation [86]. The
POM treatment of lignin leads to a sharp reduction in the content of α-OH/β-O-4 inter-unit linkages,
following demethylation. A high increase in carbonyl groups implies the loss of aromaticity in lignin,
probably due to the conversion of aromatic rings to quinone moieties [86,87]. Voitl and Rohr [68]
conducted the oxidative solvolysis of lignin by using aqueous POMs in the presence of alcohols for
converting kraft lignin into chemicals. After 20 min of reaction, a total yield of products (vanillin and
methyl vanillate) of 5.2 wt % could be obtained at 170 ◦ C and 0.5 MPa O2 (Table 1, Entry 11). In a
stirred batch reactor, vanillin (4.6 wt %) and methyl vanillate (4.2 wt %) could be obtained at 170 ◦ C
and 1.0 MPa O2 (Table 1, Entry 12) [69].
Transition metal salts are another kind of homogeneous catalysts for lignin oxidation, which have
a wide range of cation redox potential. Werhan et al. [70] reported the acidic oxidation of kraft lignin
with different transition metal salts (CuSO4 , FeCl3 , CuCl2 , CoCl2 ). The maximum yield of vanillin
gained was invariably higher for the transition metal salts than for the POMs. CoCl2 showed the
best performance among the investigated catalysts with a maximum yield of 6.3 wt % (vanillin and
methyl vanillate) at 170 ◦ C and 1.0 MPa O2 (Table 1, Entry 13). In addition, the H2 O2 mediated
oxidative solvolysis of kraft lignin was also investigated in the presence of different metal salt catalysts
(Fe2 (SO4 )3 , Fe(NO3 )·9H2 O, Mn(SO4 )·H2 O, Bi2 (SO4 )3 , Bi(NO3 )3 ·5H2 O, H2 WO4 , and Na2 WO4 ·2H2 O)
in acetone/water with the use of ultrasound irradiation [71]. The Na2 WO4 ·2H2 O catalyst was the most
efficient regarding the total yield of vanillin-based monomers (vanillin, acetovanillone, vanillic acid,
and guaiacol) with a maximum yield of 0.51 wt % at 45 ◦ C (Table 1, Entry 14). In addition, the use
of ultrasound irradiation in this experiments results in high oxidative coupling of phenoxy radicals
generated from LDP.
The metal/bromide catalyzed aerobic oxidation of alkylaromatic compounds in acetic acid
solvent have shown to be a well-established method to produce aromatic carboxylic acids [88].
Thus, Partenheimer [72] investigated the aerobic oxidative solvolysis of lignin via Co/Mn/Zr/Br
catalysts in acetic acid-water solvent. After 2 h of reaction at 180 ◦ C in the presence of 13.8 MPa air, the
highest yields of aromatic compounds with a total of 10.9 wt % could be obtained from organosolv
lignin (Table 1, Entry 15).
In summary, most of these homogeneous systems lack selectivity for lignin, resulting in a relatively
low yield of the target products. The separation and recyclability of homogeneous catalysts is still
challenging for the catalytic oxidative solvolysis of lignin. Current research is focused on developing
Polymers 2017, 9, 240 10 of 25

novel homogeneous catalysts and another kind of catalysts—heterogeneous catalysts—for the catalytic
oxidative solvolysis of lignin.

3.1.5. Heterogeneous Catalyst


Considering the limitations of homogeneous catalytic oxidative solvolysis of lignin, heterogeneous
catalytic systems have been developed with regard to their advantages in achieving a relatively
high yield of the product, combined with an easy separation and recyclability of the catalyst.
Microwave assisted catalytic oxidation of organosolv lignin over lanthanum modified SBA-15 catalyst
was investigated [73]. The catalytic oxidation with H2 O2 led to different yields of aldehydes, and
the respective acids and aceto derivatives. The highest yield of vanillin was 9.94 wt % and that of
syringaldehyde was 15.66 wt % (Table 1, Entry 16). Then, the oxidative solvolysis of organosolv lignin
in methanol using Pd/CeO2 catalyst was performed [74]. After 24 h of reaction at 185 ◦ C under O2 ,
several aromatic monomers including vanillin, guaiacol and 4-hydroxybenzaldehyde were identified
and the yields of these monomers were 5.2, 0.87 and 2.4 wt %, respectively (Table 1, Entry 17). After that,
Deng et al. [75,76] synthesized the perovskite-type LaMnO3 and LaCoO3 catalysts for catalytic wet
aerobic oxidation of lignin. Both the perovskite-type oxides, LaMnO3 and LaCoO3 , were efficient and
recyclable heterogeneous catalysts for catalytic oxidative solvolysis of lignin. When LaMnO3 was used,
the lignin conversion was 57.0% after 3 h at 120 ◦ C in the presence of 0.5 MPa O2 , and the maximum
yields of vanillin, p-hydroxybenzyl aldehyde and syringaldehyde were 4.32 (30 min), 2.03 (120 min),
and 9.33 wt % (30 min) respectively (Table 1, Entry 18). When LaCoO3 was used, the lignin conversion
after 3.0 h of reaction increased 46.7%, and the maximum yield of vanillin, p-hydroxybenzyl aldehyde
and syringaldehyde were 4.55 (60 min), 2.23 (120 min), and 9.99 wt % (50 min) respectively (Table 1,
Entry 19).
It can be concluded that both homogeneous and heterogeneous catalysts for the oxidative
solvolysis of lignin should be further developed due to the low selectivity and production for
specific compounds.

3.2. Effects of Solvent System


The oxidative solvolysis of lignin has been conducted with the help of molecular O2 or aqueous
H2 O2 as oxidants. The oxidation of raw lignin as well as lignin model compounds using molecular O2
or aqueous H2 O2 as oxidants is highly promising. It was reported that the removal of O2 in the system
of H2 O2 as an oxidant had no effect on the oxidation reactions [83]. The in-situ oxygen produced by
the decomposition of H2 O2 during the reaction might be more reactive than the external oxygen.
Solvent molecules could coordinate or dissociate to the vacant metal coordination sites in
homogeneous catalysis [89]. Different solvents in the oxidative solvolysis of lignin result in a significant
change of the yields of aromatic products. Chan et al. [58] conducted the oxidative solvolysis of
organosolv lignin into phenolic products in acetonitrile and ethyl acetate. Both acetonitrile and ethyl
acetate were optimal solvents, whereas THF had to be added to contribute to the total miscibility of all
the reaction components. Ma et al. [90] reported the selective oxidative C–C bond cleavage of a lignin
model compound over a vanadium catalyst in different solvents. In triethylamine, the C–H bond
cleavage was preferred over the vanadium-catalyzed oxidation with O2 , while C–C bond cleavage
was dominant in acetic acid. It was believed that the coordination of the carboxylic group in acetic
acid solvent to vanadium (V) catalyst was vital for C–C bond cleavage.
The oxidative solvolysis of lignin generates a number of radicals resulting in the repolymerization
reaction to limit the production of aromatic compounds. Therefore, some research explored the addition
of some radical scavengers in order to quench the reactive lignin fragments before repolymerization
took place. The alcohol as a quenching agent could generate the radicals such as CH3 O· or ·CH3 .
The role of those alcohols was investigated in the acidic degradation of lignin [68]. Methanol and
ethanol might prevent the condensation through the competitive reaction with intermediate carbonium
Polymers 2017, 9, 240 11 of 25

ions. Radical coupling of lignin fragments with CH3 O· and ·CH3 produced from methanol occurred
via acid-catalyzed formation of dimethyl ether (Equations (1) and (2)).

2CH3 OH → CH3 OCH3 + H2 O (1)

CH3 OCH3 → CH3 O· + ·CH3 BDE = 84 kcal·mol−1 (2)

The effect of varied alcohol solvents on the oxidative solvolysis of lignin is still uncertain.
Pan et al. [91] found that acetonitrile and methanol exhibit a positive effect on the oxidative degradation
of lignin, but dioxane and ethanol exhibit a negative effect. It was reported that a composite-solvent
composed of water and an organic solvent could both promote depolymerization of lignin and prevent
repolymerization of lignin fragments (monomers), increasing the yield of the aromatic products [66].
In addition, under specific reaction conditions the solvent will play a different role as opposed
to its normal role. Deng et al. [74] proposed that methanol as a solvent in oxidative solvolysis of
lignin model compounds, which could afford active H species for the hydrogenolysis of the β-O-4
bond, producing phenol and acetophenone. Similarly, Shilpy et al. [92] investigated the influence of
the property of the solvent on the oxidative solvolysis of vanillyl alcohol and vanillin. The highest
conversion rate was observed in acetic acid (91%), followed by ethyl alcohol (70%), acetonitrile (67%),
N,N-dimethyl formaldehyde (35%), tetrahydrofuran (20%) and acetone (1%). Thus it was proposed
that the major role was played by peracetic acid that was in-situ generated by reacting acetic acid
with H2 O2 over an acid catalyst [93]. Then, peracetic acid accelerated and acted as an oxidant in the
oxidation reaction.
Most of the solvents in the LDP process act as agents for dissolving lignin, products and
homogeneous catalysts. In some cases, simple alcohols could exert an unexpected influence for
protecting the produced aromatic fragments during the oxidative solvolysis process of lignin.
The multi-function of solvents should be further investigated for the low-cost and high-efficient
oxidative solvolysis of lignin.

4. Oxidative Solvolysis of Lignin-Related Model Compounds


The complexity and variability of lignin structure promoted the use of lignin-related model
compounds for LDP studies, giving rise to a better understanding of the cracking mechanism via
specific inter-unit linkage thus promoting the production of aromatic compounds. In this section,
oxidative solvolysis of lignin-related model compounds (both aromatic monomers and oligomers) will
be overviewed, concentrating on the reforming of aromatic monomers and the cracking of inter-unit
linkages in oligomers.

4.1. Oxidative Reforming of Lignin-Derived Monomers


To understand the reactivity of the functional groups on lignin, much attention have been paid to
the oxidative reforming of aromatic monomers. The representative lignin-related aromatic monomers
such as apocynol, veratryl alcohol, 3-methoxy-4-hydroxybenzyl alcohol and 4-hydroxybenzyl alcohol
were studied in recent years.

4.1.1. Apocynol
Sushanta et al. [94,95] reported an efficient oxidative solvolysis of apocynol (M1) over SBA-15
and Co (salen)/SBA-15 by using H2 O2 as an oxidant. The reaction pathway for M1 conversion on
SBA-15 and Co (salen)/SBA-15 is shown in Figure 4. The first oxidation product was acetovanillone,
then vanillin was formed through the side chain cleavage while 2-methoxybenzoquinone as a product
of oxidation of the phenolic group together with oxidative degradation of the p-alkyl side chain.
Co (salen)/SBA-15 showed a highly efficient performance (nearly 100%) for M1 conversion after
40 min microwave heating and SBA-15 showed an unusual oxidative ability and excellent selectivity
for acetovanillone. Badamali et al. [96] also demonstrated the oxidation of M1 by mesoporous MCM-41,
Polymers 2017, 9, 240 12 of 25

HMS, SBA-15 and amorphous silica with using H2 O2 as an oxidant under microwave irradiation.
With the same conversion pathway on SBA-15 or Co (salen)/SBA-15, acetovanillone, vanillin and
2-methoxybenzoquinone
Polymers 2017, 9, 240 were the main products (Figure 4). Silica as a catalyst led to the 12 ofhighest
25
conversion (94%) of M1 among the studied catalysts with the equal selectivity to acetovanillone
selectivity to acetovanillone It
and 2-methoxybenzoquinone. and 2-methoxybenzoquinone.
could be concluded that the It could
surfacebe hydroxyl
concluded groups
that theand
surface
internal
hydroxyl
Polymers groups
2017, 9, 240 and internal pore structure of the employed silicas play a key role in
pore structure of the employed silicas play a key role in the catalytic oxidative process. In addition,the catalytic
12 of 25
oxidative process. In addition, Hanson et al. [97] reported that dipicolinate vanadium complexes
Hanson et al. [97] reported that dipicolinate vanadium complexes oxidized 2-phenoxyethanol in air.
selectivity to acetovanilloneinand
oxidized 2-phenoxyethanol air.2-methoxybenzoquinone. It could
After the reaction at 100 °C for onebeweek,
concluded that the surface
approximately 20% of
Afterhydroxyl
the reaction at 100 ◦ C for one week, approximately 20% of the 2-phenoxyethanol was consumed.
groups
the 2-phenoxyethanol and internal pore
was consumed.structure of the
Phenol employed
(18%), formic silicas play aand
acid (6%), keyseveral
role in other
the catalytic
minor
Phenol (18%),
oxidative
unidentifiedformic
process. acid
products (6%), and
In addition, several other minor unidentified products were detected.
Hanson et al. [97] reported that dipicolinate vanadium complexes
were detected.
oxidized 2-phenoxyethanol in air. After the reaction at 100 °C for one week, approximately 20% of
the 2-phenoxyethanol was consumed. Phenol (18%), formic acid (6%), and several other minor
unidentified products were detected.

Figure 4. Reaction pathway for apocynol (M1) conversion on the catalysts.


Figure 4. Reaction pathway for apocynol (M1) conversion on the catalysts.
4.1.2. Veratryl alcohol/3-methoxy-4-hydroxybenzyl alcohol/4-hydroxybenzyl alcohol
4.1.2. Veratryl Alcohol/3-Methoxy-4-Hydroxybenzyl
Figure 4. Reaction pathway for apocynol Alcohol/4-Hydroxybenzyl
(M1) conversion on the catalysts. Alcohol
Mate et al. [98,99] reported the oxidation of veratryl alcohol over a nano-structured, spinel
Mate
Co3O4et al. [98,99]
catalyst reported
in the presence theofoxidation
O2. After of 7 hveratryl alcohol
of reaction at 140
4.1.2. Veratryl alcohol/3-methoxy-4-hydroxybenzyl alcohol/4-hydroxybenzyl alcohol
over°C ainnano-structured,
water, the Co3O4spinel catalystCo3 O4
yielded the highest conversion of 85% together with a ◦96% selectivity
catalyst in the presence of O2 . After 7 h of reaction at 140 C in water, the Co3 O4 catalyst yielded the to veratryl aldehyde. In
Matesolvents,
et al. [98,99] reported the oxidation of3Overatryl
different
highest conversion of the
85%catalytic
together activity
with aof96% the Co selectivity toalcohol
4 catalyst over atonano-structured,
was proved
veratryl aldehyde. occur In
in different spinel
the followingsolvents,
Co 3O4 catalyst
order: toluene in
> the presence
water > of O
ethanol > 2.methanol.
After 7 h The of reaction
highest atconversion
140 °C in (75%)water,wasthe Co 3O4 catalyst
observed inwater
a
the catalytic activity of the Co3 O4 catalyst was proved to occur in the following order: toluene > >
yielded the
nonpolar highest
solvent suchconversion
as toluene of together
85% together
with abovewith a97% 96%selectivity
selectivitytotoveratryl
veratrylaldehyde.
aldehyde.The In
ethanol > methanol. The
different solvents,
highest conversion (75%) wascatalyst
observed inproved
a nonpolar solvent such as toluene
proposed pathwaythe for catalytic
oxidationactivity of the
of veratryl Co3O4over
alcohol Co3Owas4 catalyst was to presented
occur in the following
in Scheme 1:
together
order:with above 97%
tolueneveratryl selectivity
> wateralcohol
> ethanol to veratryl
> methanol. aldehyde.
The highest The proposed pathway for oxidation
in a of
The adsorbed formed an intermolecular bond conversion
between Co3+ (75%)
and O was observed
2 while the surface
veratryl
oxidealcohol
nonpolar over Co
ion solvent
formed 3 Ointermolecular
such
an 4ascatalyst
toluenewas presented
together
bond with in
with theScheme
above 97% 1:
hydrogen The adsorbed
selectivity
of the to veratryl
veratryl
alcoholic alcohol
aldehyde.
group; then Theformed
the
an intermolecular bondforbetween 3+
proposed pathway
abstraction of hydrogen by theCo
oxidation oxide and
ion O
of veratryl 2 while
alcohol
lead to the theformation
over surface oxide
Co3O4 catalyst ion
of Co3+was formed
species; an
OHpresented inintermolecular
afterScheme
that the1:
The adsorbed
bondactivated
with theCo veratryl
hydrogen
3+OH alcohol formed an intermolecular bond between Co 3+ and O2 while the surface
species could produce another intermolecular hydrogen bond and abstraction of ion
of the alcoholic group; then the abstraction of hydrogen by the oxide
oxide
lead an
to the ion formed
resultedan
formation
electron ofinCothe3+ OH
intermolecular
species;
formation bond with
after
of Co 2+ theand
, Hthat
2O hydrogen
the activated
the Co3+
mainofproduct,
the alcoholic
OH group;
species
veratryl then produce
could
aldehyde; the
the
abstraction
reduced
another of hydrogen
Co specieshydrogen
2+
intermolecular by the
were reoxidizedoxide
bond and ion lead to
withabstraction the
O2 to regenerate formation
of anthe of Co OH
3+3+
active resulted
electron species;
Co O species.

in the formation the
after that of Co2+ ,
activated Co3+OH species could produce another intermolecular2+hydrogen bond and abstraction of
H2 O and the main product, veratryl aldehyde; the reduced Co species were reoxidized with O2 to
an electron resulted 3+ in the formation of Co2+, H2O and the main product, veratryl aldehyde; the
regenerate the active Co O− species.
reduced Co species were reoxidized with O2 to regenerate the active Co O species.
2+ 3+ −

Scheme 1. Proposed pathways for oxidation of veratryl alcohol over Co3O4 catalyst [99], reproduced
with permission from Elsevier.

Gu et 1. Proposed
Scheme al. [100,101]
pathwaysperformed
for oxidationthe oxidation
of veratryl of Co
alcohol over a 3Olignin
4 catalystmodel compound,
[99], reproduced
Scheme 1. Proposed pathwaysalcohol
3-methoxy-4-hydroxybenzyl for oxidation
with permission from Elsevier. (M2), of veratryl
over alcohol overLa/SBA-15
a mesoporous Co3 O4 catalyst [99], reproduced
catalyst. Sixty-eight
with permission
percent from
conversion of Elsevier.
M2 to vanillin was obtained after 30 min of reaction under 200 W microwave
Gu et al. [100,101] performed the oxidation of a lignin model compound,
3-methoxy-4-hydroxybenzyl alcohol (M2), over a mesoporous La/SBA-15 catalyst. Sixty-eight
percent conversion of M2 to vanillin was obtained after 30 min of reaction under 200 W microwave
Polymers 2017, 9, 240 13 of 25

Gu et al. [100,101] performed the oxidation of a lignin model compound, 3-methoxy-4-


hydroxybenzyl alcohol (M2), over a mesoporous La/SBA-15 catalyst. Sixty-eight percent conversion
Polymers 2017, 9, 240 13 of 25
of M2 to vanillin was obtained after 30 min of reaction under 200 W microwave irradiation
with the presence
irradiation with theofpresence
H2 O2 asofoxidant. ReactionReaction
H2O2 as oxidant. pathway for M2for
pathway conversion
M2 conversion on La/SBA-15
on La/SBA-15 was
shown in Figure 5. Pan et al. [91] also reported the microwave-assisted
was shown in Figure 5. Pan et al. [91] also reported the microwave-assisted oxidative degradation of oxidative degradation
of 3-methoxy-4-hydroxybenzyl
3-methoxy-4-hydroxybenzyl alcohol
alcohol (M2)(M2) and 4-hydroxybenzyl
and 4-hydroxybenzyl alcoholalcohol
(M3) with (M3) 14 with
types14 types
of metal
of metal salts in the presence
salts in the presence of H2O2. After2 10 of H O . After 10 min of reaction with
2 min of reaction with a CrCl3 catalyst 3at 80 °C in a CrCl catalyst at

80 C in H2 O/acetonitrile
H2O/acetonitrile (1:1, v/v),(1:1, M2v/v), wasM2converted
was converted to vanillin
to vanillin (59.8%),
(59.8%), vanillicacid
vanillic acid (9.7%),
(9.7%),
5-hydroxyl-4-methoxyl-7-ketone-4-heptylic acid (17.0%), etc., through
5-hydroxyl-4-methoxyl-7-ketone-4-heptylic acid (17.0%), etc., through the pathway depicted in the pathway depicted in Figure 5.
Different from the conversion on La/SBA-15, vanillic acid underwent
Figure 5. Different from the conversion on La/SBA-15, vanillic acid underwent a ring-opening a ring-opening reaction to
produce
reaction 5-hydroxyl-4-methoxyl-7-ketone-4-heptylic
to produce 5-hydroxyl-4-methoxyl-7-ketone-4-heptylic acid. Under the same acid. reaction
Under the conditions, M3 was
same reaction
converted to hydroquinone (33.2%), 4-hydroxybenzaldehyde
conditions, M3 was converted to hydroquinone (33.2%), 4-hydroxybenzaldehyde (22.2%), (22.2%), 4-hydroxybenzonic acid (7.6%),
etc. The reaction pathway for M3 conversion
4-hydroxybenzonic acid (7.6%), etc. The reaction pathway on CrCl 3 was shown in Figure 6. 4-hydroxybenzaldehyde
for M3 conversion on CrCl3 was shown in
and 4-hydroxybenzonic acid were the first
Figure 6. 4-hydroxybenzaldehyde and 4-hydroxybenzonic acid two oxidation products from M3. the
were Thenfirstthe decarboxylation
two oxidation
of
products from M3. Then the decarboxylation of 4-hydroxybenzonic acid resulted in thewith
4-hydroxybenzonic acid resulted in the formation of phenol and then phenol reacted hydroxyl
formation of
radicals
phenol and formed
then by a microwave
phenol reacted with irradiation
hydroxyl ofradicals
H2 O2 , to produce
formed by ahydroquinone.
microwave irradiation In addition,
of H2the
O2,
pathway
to produce M2 on immobilized
for hydroquinone. MTO [63]the
In addition, andpathway
on CoTiOfor 3 [92]
M2was on also different from
immobilized MTO the[63]
conversion
and on
on La/SBA-15 and on
CoTiO3 [92] was also different CrCl 3 from the conversion on La/SBA-15 and on CrCl3 (Figure 5).catalyst
(Figure 5). After 6 h of reaction with immobilized MTO After 6at h
room temperature in CH COOH/H O , M2 was converted
of reaction with immobilized MTO catalyst at room temperature in CH3COOH/H2O2, M2 was
3 2 2 to vanillin (14.2%), vanillic acid (18.9%),
muconolactone
converted to vanillin(5.5%), etc. The
(14.2%), muconolactone
vanillic acid (18.9%), wasmuconolactone
mostly gained (5.5%), via an excessive oxidative ring
etc. The muconolactone
opening
was mostly reaction to avia
gained muconic acid intermediate
an excessive oxidative ring followed
openingby the formation
reaction to a of a lactone
muconic acidmoiety, perhaps
intermediate
catalyzed by MTO. For the CoTiO catalyst, with the help of H O
followed by the formation of a lactone moiety, perhaps catalyzed by MTO. For the CoTiO3 catalyst,
3 2 2 in acetic acid and isopropanol
solvent,
with the ahelp remarkable
of H2O2 selectivity
in acetic acid of 99.8% to vanillin solvent,
and isopropanol was obtained at 85 ◦ C.
a remarkable The major
selectivity of products
99.8% to
were
vanillinvanillin, vanillicatacid
was obtained 85 °C. andTheguaiacol.
major products Schemewere 2 illustrated the proposed
vanillin, vanillic acid and mechanism for the
guaiacol. Scheme
oxidation of vanillyl alcohol over CoTiO [92]. Initially the catalyst
2 illustrated the proposed mechanism for the oxidation of vanillyl alcohol over CoTiO3 [92]. Initially
3 reacted with H 2 O 2 to form a
hydroperoxyl radical, which acts as a source of highly active oxidant, and
the catalyst reacted with H2O2 to form a hydroperoxyl radical, which acts as a source of highly active then attacks the catalyst to
form an intermediate.
oxidant, and then attacks When the vanillyl
catalyst alcohol
to formwas activated on the
an intermediate. When catalyst
vanillylsurface,
alcohol thewas
electron-rich
activated
hydroperoxyl attacks the partially positive charged carbon and the hydrogen
on the catalyst surface, the electron-rich hydroperoxyl attacks the partially positive charged carbon of the hydroxyl group
through oxygen to form
and the hydrogen of theanhydroxyl
activatedgroupoxygen species.oxygen
through This facilitated
to form an theactivated
removal of the hydroxyl
oxygen species.group
This
from the hydroperoxyl and produced water as a by-product. At
facilitated the removal of the hydroxyl group from the hydroperoxyl and produced water as a the same time, the vanilloxy cation
released
by-product. a proton that same
At the was attacked
time, the by electron-rich
vanilloxy cation oxygen of the hydroperoxyl
released a proton that groupwasto attacked
produce the by
final product,oxygen
electron-rich vanillicofacid.
the hydroperoxyl group to produce the final product, vanillic acid.

Figure 5. Reaction pathway for 3-methoxy-4-hydroxybenzyl alcohol (M2) conversion on the


Figure 5. Reaction pathway for 3-methoxy-4-hydroxybenzyl alcohol (M2) conversion on the catalysts.
catalysts.
Polymers 2017, 9, 240 14 of 25
Polymers 2017, 9, 240 14 of 25
Polymers 2017, 9, 240 14 of 25

Figure 6. Reaction pathway for 4-hydroxybenzyl alcohol (M3) conversion on the catalyst [91],
Figure 6. Reaction
Figure
pathway
6. Reaction
for
pathwayfrom
4-hydroxybenzyl alcohol
for 4-hydroxybenzyl
alcohol(M3)(M3) conversion on the catalyst [91],
conversion on the catalyst [91],
reproduced with permission American Chemical Society.
reproduced withwith
reproduced permission from
permission fromAmerican
AmericanChemical Society.
Chemical Society.

Scheme 2. The suggested mechanism for formation of the main oxidation product (vanillin and
Scheme 2. The
vanillic acid) suggested
from vanillyl mechanism forpresence
alcohol in the formation of the3 [92],
of CoTiO mainreproduced
oxidation product (vanillin from
with permission and
Scheme 2. The suggested mechanism for formation of the main oxidation product (vanillin and vanillic
vanillic acid) from
Royal Society vanillyl alcohol in the presence of CoTiO3 [92], reproduced with permission from
of Chemistry.
acid)Royal
from Society
vanillyl ofalcohol in the presence of CoTiO3 [92], reproduced with permission from Royal
Chemistry.
Society of Chemistry.
4.2. Oxidative Cracking of Typical Inter-Unit Linkages in Lignin-Derived Oligomers
4.2. Oxidative Cracking of Typical Inter-Unit Linkages in Lignin-Derived Oligomers
4.2.1. C–CCracking
4.2. Oxidative Linkage of Typical Inter-Unit Linkages in Lignin-Derived Oligomers
4.2.1. C–C Linkage
Homogeneous copper and vanadium complexes are well-known to be effective catalysts for
4.2.1. C–CHomogeneous
Linkage copper of and vanadium complexesC–C arebond
well-known
aerobic oxidation reactions diols with concomitant cleavageto[102,103].
be effective catalysts that
Recognizing for
aerobic
copper oxidation
Homogeneous and vanadium reactions
copper of diols
complexes
and with
couldconcomitant
vanadium potentially
complexes C–C
beare bond
used cleavage
for the C–C
well-known [102,103].
tobond Recognizing
cleavage
be effective in that for
lignin
catalysts
copper and vanadium
depolymerization, some complexes
scientists could
began potentially
to synthesize be oxovanadium
used for the C–C (e.g., bond
Complexcleavage
1 in in lignin
Figure 3)
aerobic oxidation reactions of diols with concomitant C–C bond cleavage [102,103]. Recognizing
depolymerization,
and copper catalysts someandscientists began
explore their to synthesize
fundamental oxovanadium
reactivity. Hanson (e.g.,
et al. Complex
[97,104] 1synthesized
in Figure 3)a
that copper and vanadium complexes could potentially be used for the C–C bond cleavage in lignin
and
series copper catalystscomplexes
of vanadium and explore andtheir fundamental
homogeneous reactivity.
copper Hanson
catalysts et al. their
to explore [97,104] synthesized
reactivity toward a
depolymerization,
series of oxidation
some scientistsand
vanadiumofcomplexes
began to synthesize
homogeneous
oxovanadium
copper catalysts
(e.g., Complex 1 in Figure
to explore their reactivity toward
3) and
aerobic a simple lignin model compound, 1,2-diphenyl-2-methoxyethanol (M4). The
copper catalysts
aerobic
vanadium
and explore
oxidation
complex of(dipic)V their
a simple fundamental
lignin
V(O)(O iPr)model
(Complex
reactivity.
compound,
1) reacted
Hanson et al. [97,104] synthesized
1,2-diphenyl-2-methoxyethanol
with M4 in DMSO with the addition (M4).a The series
of
of
vanadium
vanadium complexes
air at 100 °C, complex and
producing homogeneous
(dipic)V V copper
(O)(O Pr) (Complex
i
predominantly catalysts
benzaldehyde1) reacted to
andwith explore
M4 in but
methanol, their
DMSO reactivity toward
with the itaddition
in pyridine produced aerobic
of
oxidation
air at of100a °C,
predominantly simple ligninacid
producing
benzoic modeland compound,
predominantly
methyl benzaldehyde
benzoate1,2-diphenyl-2-methoxyethanol
(Figureand7).methanol,
The ketone butbenzoin (M4).
in pyridine
methyl itThe vanadium
produced
ether was
complex (dipic)VV (O)(O
predominantly
an intermediate benzoic i Pr)
in this acid and methyl
(Complex
reaction. The benzoate
reacted (Figure
with M4 in 7).DMSO
The ketone
1) CuCl/2,2,6,6-tetramethylpiperidine-1-oxyl withbenzoin methyl
radicalofether
the addition air at
(TEMPO) was100 ◦ C,
an
producing intermediate
mixture could also
predominantly in this reaction.
react with M4
benzaldehyde The in CuCl/2,2,6,6-tetramethylpiperidine-1-oxyl
pyridine
and methanol,underbut O2inatpyridine
100 °C, itaffording radical
produced (TEMPO)
a predominantly
mixture of
mixture
benzaldehyde could also
(84%) react
and with
methyl M4
benzoatein pyridine
(88%) under
(Figure 7). O 2 at 100
Unlike
benzoic acid and methyl benzoate (Figure 7). The ketone benzoin methyl ether was an intermediate the °C,
vanadium affording
catalysta mixture
system, of
the
benzaldehyde
copper-catalyzed (84%) and methyl
reactions werebenzoate
proposed (88%) (FigureC–C
for direct 7). Unlike the vanadium
bond radical
cleavage of M4catalyst
without system,
ketone the
or also
in this reaction. The CuCl/2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) mixture could
copper-catalyzed
aldehyde intermediates. reactionsRecently,
were proposedfour◦ newfor direct C–C bondcomplexes
vanadium(V) cleavage of ofM4 without ketone or
amino-bis(phenolate)
react with M4 in pyridine under O2 at 100 C, affording a mixture of benzaldehyde (84%) and methyl
aldehyde
ligands (e.g., intermediates.
Complex 5 inRecently,Figure 3)four havenew been vanadium(V)
synthesized, and complexes
all complexesof amino-bis(phenolate)
were employed for
benzoate
ligands
(88%) (Figure
(e.g., aerobic
Complex
7). 5Unlike
in Figure
the vanadium catalyst system, thecomplexes
copper-catalyzed reactions were
the catalytic oxidative C–C3)bond havecleavage
been synthesized,
of M4 in air andat all
100 °C, wherewere employedand
benzaldehyde for
proposed
the
methyl for direct
catalytic
benzoate C–C
aerobic
were bond
oxidative cleavage
identifiedC–C of
as bond M4 without
cleavage
the major ketone
of M4[105].
products or aldehyde
in airInatDMSO, intermediates.
100 °C,awhere reactionbenzaldehyde Recently,
with Complex and5 four
new methyl
vanadium(V)
produced benzoate complexes
were identified
benzaldehyde of amino-bis(phenolate)
at 90%, as the major
methanol products
at 46%, methyl ligands
[105]. (e.g.,
In DMSO,
benzoate Complex
at 36%, benzoin5
a reaction in Figure
with
methyl ether3)at
Complex 5 have
been produced
synthesized,
18%, smalland
and abenzaldehyde all complexes
amount atof90%,
benzoic were
methanol
acid. Inemployed
at 46%, methyl
pyridine, for benzoate
methyl thebenzoate
catalytic
at 36%, aerobic
(67%), benzoin oxidative
benzoin methyl
methyl C–C
ether atbond
ether
18%,
cleavage ofand
M4 a in
small
air amount
at 100 ◦ C,
ofwere
benzoic
where acid. In
benzaldehydepyridine, andmethyl
methyl benzoate
benzoate (67%), benzoin methyl as ether
(15%), and methanol (2%) obtained. Benzaldehyde was detected as a were
minor identified
product (1%) the
andmajor
(15%),
no and
benzoic methanol
acid was (2%)
detected.were obtained. Benzaldehyde was detected
products [105]. In DMSO, a reaction with Complex 5 produced benzaldehyde at 90%, methanol at 46%, as a minor product (1%) and
methylno benzoate
benzoic acid was detected.
at 36%, benzoin methyl ether at 18%, and a small amount of benzoic acid. In pyridine,
methyl benzoate (67%), benzoin methyl ether (15%), and methanol (2%) were obtained. Benzaldehyde
was detected as a minor product (1%) and no benzoic acid was detected.
Polymers 2017, 9, 240 15 of 25
Polymers 2017, 9, 240 15 of 25

Polymers 2017, 9, 240 15 of 25

Figure 7. The aerobic oxidation of 1,2-diphenyl-2-methoxyethanol (M4) with vanadium and copper
Figure 7. The aerobic oxidation of 1,2-diphenyl-2-methoxyethanol (M4) with vanadium and
catalysts.
copper catalysts.
Degradation research indicated that β-1 structures are widespread, especially in hardwood
lignin FigureThe
[106].
Degradation 7.research
The
β-1aerobic
linkage oxidation
has notofyet
indicated 1,2-diphenyl-2-methoxyethanol
thatbeen
β-1adequately
structures addressed (M4)
bywith
are widespread, vanadium
reductive and copper
cleavage.
especially However,
in hardwood
Hanson catalysts.
et al. [47,104,107] successfully used their CuOTf/2,6-lutidine/TEMPO catalyst system and
lignin [106]. The β-1 linkage has not yet been adequately addressed by reductive cleavage. However,
(HQ)2VV(O)(OiPr) (Complex 2) to achieve the aerobic oxidation of nonphenolic and phenolic β-1
Hanson et al. [47,104,107]
Degradation successfully
research indicatedused their
that β-1 CuOTf/2,6-lutidine/TEMPO
structures are widespread, especiallycatalyst system and
in hardwood
lignin
V modelsi (M5 and M6). It was found that the selectivity of the aerobic oxidation of β-1 lignin
(HQ)2 Vlignin(O)(O Pr)The
[106]. (Complex
β-1 linkage to not
2)has achieve
yet beenthe aerobic addressed
adequately oxidationbyofreductive
nonphenolic
cleavage. and phenolic β-1
However,
model compounds depended on both the catalyst and substrate. In the absence of phenolic groups,
Hanson (M5
ligninComplex
models et al. and
[47,104,107]
M6). Itsuccessfully
was found used
that their
the CuOTf/2,6-lutidine/TEMPO
selectivity of the aerobic catalyst
oxidationsystemofand
β-1 lignin
2 showed a high selectivity for Cα–H bond cleavage, however the copper catalyst system
model (HQ) 2VV(O)(OiPr) (Complex 2) to achieve the aerobic oxidation of nonphenolic and phenolic β-1
compounds
worked
ligninon Cα–Cdepended
models
on both the catalyst and substrate. In the absence of phenolic groups,
β bond cleavage (Figure 8a). The introduction of a phenolic functional group in the
(M5 and M6). It was found that the selectivity of the aerobic oxidation of β-1 lignin
Complex 2 showed
substrate enabled a high
the selectivity
cracking forCC
of both
the α –H
α–C bond with
aryl bond cleavage,
the helphowever boththe
of absence copper and
vanadium catalyst
coppersystem
model compounds depended on the catalyst and substrate. In the of phenolic groups,
worked on
catalysts
ComplexC –C
(Figure
α 2 showed
β bond
8b). As cleavage
for the (Figure
copper 8a).
catalyst, The
the introduction
reaction of C–C of
bond a phenolic
cleavage
a high selectivity for Cα–H bond cleavage, however the copper catalyst system functional
might proceed group
via in
either
the substrate a one-electron
workedenabled
on Cα–Cβthe bondoxidation
cracking or a
cleavageof(Figure selective
the Cα –CThe
8a). oxidation of
bond with of
aryl introduction the
the primary
help of both
a phenolic alcohol, followed
vanadium
functional group inandby a
the copper
retro-aldol
catalysts (Figurereaction,
substrate 8b). Asthe
enabled while,
for the for
cracking the
copper
of the vanadium
catalyst,
Cα–Caryl the catalyst,
bond with the
reactiontheofreaction
C–Cofbond
help wascleavage
both in accordance
vanadium might with a via
proceed
and copper
eithertwo-electron
a catalysts
one-electron oxidation
(Figure pathway
8b). As
oxidation fororthe [47,104].
a copper
selective catalyst, the reaction
oxidation of theofprimary
C–C bond cleavagefollowed
alcohol, might proceed via
by a retro-aldol
reaction,either a one-electron oxidation or a selective oxidation of the primary alcohol, followed by a
while, for the vanadium catalyst, the reaction was in accordance with a two-electron oxidation
retro-aldol reaction, while, for the vanadium catalyst, the reaction was in accordance with a
pathway [47,104].
two-electron oxidation pathway [47,104].

Figure 8. (a) The aerobic oxidation of non-phenolic β-1 linkage lignin model compound (M5) with
vanadium and copper catalysts; and (b) the aerobic oxidation of phenolic β-1 linkage lignin model
Figure 8. (M6)
compound (a) The aerobic
with oxidation
vanadium of non-phenolic
and β-1 linkage
copper catalysts [104], lignin model compound
reproduced (M5) with
withcompound
permission from with
Figure 8. (a) The aerobic oxidation of non-phenolic β-1 linkage lignin model (M5)
vanadium and copper catalysts; and (b) the aerobic oxidation of phenolic β-1 linkage lignin model
American Chemical Society.
vanadium and copper catalysts; and (b) the aerobic oxidation of phenolic β-1 linkage lignin model
compound (M6) with vanadium and copper catalysts [104], reproduced with permission from
compound (M6) with
American vanadium
Chemical Society.and copper catalysts [104], reproduced with permission from American
Chemical Society.
Polymers 2017, 9, 240 16 of 25

Polymers 2017, 9, 240 16 of 25


4.2.2. C–O Linkage
4.2.2. C–O Linkage
The benzylic C–H or C–OH bonds in lignin would be easily attacked and transformed to carbonyl
The oxidative
groups under benzylic C–H or C–OH
conditions. bonds
Next, theinCαlignin
–O orwould
Cβ –O be easily
bond canattacked andtotransformed
be cracked to
the corresponding
carbonyl groups under oxidative conditions. Next, the Cα–O or Cβ–O bond can be cracked to the
fragments [3]. The oxidation of α-O-4 or β-O-4 linkages containing lignin model compounds has been
corresponding fragments [3]. The oxidation of α-O-4 or β-O-4 linkages containing lignin model
extensively investigated. The key for achieving the production of specific chemicals from selective
compounds has been extensively investigated. The key for achieving the production of specific
oxidation of lignin depends on the advance of catalysts.
chemicals from selective oxidation of lignin depends on the advance of catalysts.
4.2.3. α-O-4 Containing Model Compound
4.2.3. α-O-4 Containing Model Compound
Haibach et al.et[108]
Haibach investigated
al. [108] investigatedthe thedehydroaryloxylation
dehydroaryloxylation ofofaryl
aryl alkyl
alkyl ethers
ethers in p-xylene
in p-xylene usingusing
pincer iridium catalysts at a moderate temperature (150 or 200 ◦ C). This system showed a rare
pincer iridium catalysts at a moderate temperature (150 or 200 °C). This system showed a rare
atom-economical
atom-economical process
processforfor
C–OC–O ether
etherbond
bond cleavage, producinga alimited
cleavage, producing limited number
number of substituted
of substituted
alkyl alkyl
aryl ethers. Recently,
aryl ethers. Recently,Gao et et
Gao al.al.[64]
[64]used
usednitrogen-containing graphene
nitrogen-containing graphene material
material (LCN)
(LCN) as anas an
effective catalyst for the oxidation of α-O-4 types of a lignin model compound (M7)
effective catalyst for the oxidation of α-O-4 types of a lignin model compound (M7) in the presence of in the presence
of tert-butyl
tert-butyl hydroperoxide
hydroperoxide (TBHP), (TBHP),
to produceto produce aromatic
aromatic aldehydes
aldehydes and acids
and acids in high
in high yield.yield.
The The
reaction
reaction pathway for M7 conversion on LCN is shown in Figure 9. A free-radical
pathway for M7 conversion on LCN is shown in Figure 9. A free-radical mechanism was involved, mechanism was
involved, initiated by a benzylic C–H bond activation, followed by a Cα–O bond cleavage, and
initiated by a benzylic C–H bond activation, followed by a Cα –O bond cleavage, and finally completed
finally completed by a further oxidation of intermediate aromatics. Similarly, an oxidative C–O
by a further oxidation of intermediate aromatics. Similarly, an oxidative C–O bond cleavage of M7
bond cleavage of M7 was demonstrated on Complex 5 at 80 °C, but the conversion was significantly
was demonstrated ◦
limited [105]. on Complex 5 at 80 C, but the conversion was significantly limited [105].

Figure 9. Reaction pathway for α-O-4 lignin model dimer (M7) conversion on the catalyst [64],
Figure 9. Reaction pathway for α-O-4 lignin model dimer (M7) conversion on the catalyst [64],
reproduced with permission from John Wiley and Sons.
reproduced with permission from John Wiley and Sons.
4.2.4. β-O-4 Containing Model Compound
4.2.4. β-O-4 Containing Model Compound
The oxidative solvolysis of a lignin model compound, 2-phenoxy-1-phenylethanol (M8), was
performed
The with solvolysis
oxidative a CrCl3 catalyst
of awhile
lignin using
modelH2Ocompound,
2 as an oxidant at 80 °C [91]. After 10 min of reaction
2-phenoxy-1-phenylethanol (M8), was
in H2O/acetonitrile
performed with a CrCl3solvent,
catalystM8 while wasusing
converted
H2 O2 as to anphenol
oxidant(3.1%),
at 80benzoic
◦ C [91]. acidAfter
(7.4%),
10 min
2-hydroxy-1-phenylethanone (17.3%), 2-phenoxy-1-phenylethanone (20.1%),
of reaction in H2 O/acetonitrile solvent, M8 was converted to phenol (3.1%), benzoic acid (7.4%), etc. The reaction
pathway for M8 on CrCl3 is shown in Figure 10. M8 was firstly oxidized to produce
2-hydroxy-1-phenylethanone (17.3%), 2-phenoxy-1-phenylethanone (20.1%), etc. The reaction
2-phenoxy-1-phenylethanone. The β-O-4 bond in 2-phenoxy-1-phenylethanone was cleaved through
pathway for M8 on CrCl3 is shown in Figure 10. M8 was firstly oxidized to produce
two pathways: (1) between the Cβ–O atoms with 2-hydroxy-1-phenylethanone and phenol as the
2-phenoxy-1-phenylethanone.
major product; and (2) between The β-O-4the Cbond in 2-phenoxy-1-phenylethanone was cleaved through
α–Cβ atoms with benzoic acid as the major product. It is
two pathways: (1) between the
estimated that the cleavage of the β C –O atoms
Cβ–O bond was with 2-hydroxy-1-phenylethanone
easier than that of the Cα–Cβ bond. Hanson and phenol
et al. as
the major product;
[109] used and (2) between the Cα
the CuOTf/2,6-lutidine/TEMPO –Cβ atoms
catalyst systemwith benzoic
to conduct the acid
aerobicas oxidation
the majorof product.
M8
It is obtaining
estimatedthethat theconversion
overall cleavage ofof67% theafter
Cβ 40
–Oh bond was
at 100 °C. Theeasier
major than
productsthatwereof the Cα –Cβ bond.
the formylated
Hansonsubstrate
et al. [109]2-phenoxy-1-phenylformate,
used the CuOTf/2,6-lutidine/TEMPO and the catalyst
TEMPO-functionalized
system to conduct the ketone
aerobic
2-phenoxy-1-phenyl-2-(2,2,6,6-tetramethylpiperidin-1-yloxy)ethanone, while
oxidation of M8 obtaining the overall conversion of 67% after 40 h at 100 C. The major products phenol,
◦ benzoic acid
were and 2-phenoxyacetophenone
the formylated were produced only in small amounts.
substrate 2-phenoxy-1-phenylformate, Recently, a new strategy for
and the TEMPO-functionalized ketone
catalytic oxidation cleavage of aryl ethers in aralkyl aryl ethers involving a hemiacetal-like structure
2-phenoxy-1-phenyl-2-(2,2,6,6-tetramethylpiperidin-1-yloxy)ethanone, while phenol, benzoic acid
was proposed: VO(acac)2-catalyzed aerobic oxidation of M8 in the presence of acetic acid [90,110].
and 2-phenoxyacetophenone were produced only in small amounts. Recently, a new strategy
Benzaldehyde (5%), benzoic acid (14%), phenyl formate (9%), phenol (54%),
for catalytic oxidation cleavage
2-phenoxy-1-phenylethanone (2%) of aryl ethers
and some in aralkylproducts
esterification aryl ethers
(16%)involving
were obtained a hemiacetal-like
after 8 h at
structure
80 °C.was proposed:
It was proposed VO(acac) 2 -catalyzed
that a cleavage aerobic
of the linkage (C–Coxidation
bond and M8 bond)
ofC–O in thebetween
presencethe of
twoacetic
acid aromatic
[90,110].ringsBenzaldehyde (5%), atbenzoic
could occur through least twoacid (14%),route
pathways: phenyl
A andformate
a new route (9%), phenol10).
B (Figure (54%),
2-phenoxy-1-phenylethanone (2%) and some
For route A, 2-phenoxy-1-phenylethanone wasesterification
the intermediate products (16%)
that further weretoobtained
cleaved after 8 h
benzoic acid,
◦ C. It was
at 80 phenyl formate and phenol.
proposed that aFor route B,of
cleavage anthe
initial oxidative
linkage (C–C activation
bond and of the
C–O Cβ–H
bond)bond produced
between a two
the
β-hydroxyl hemiacetal. Then, through a further cleavage of the C–C bond,
aromatic rings could occur through at least two pathways: route A and a new route B (Figure 10). the following oxidative
For route A, 2-phenoxy-1-phenylethanone was the intermediate that further cleaved to benzoic acid,
phenyl formate and phenol. For route B, an initial oxidative activation of the Cβ –H bond produced a
Polymers 2017, 9, 240 17 of 25

β-hydroxyl hemiacetal.
Polymers 2017, 9, 240 Then, through a further cleavage of the C–C bond, the following 17 of oxidative
25

activation of the Cα –H bond, and the eventual cleavage of the C–O bond can generate benzaldehyde,
activation of the Cα–H bond, and the eventual cleavage of the C–O bond can generate benzaldehyde,
benzoicbenzoic
acid, phenyl formate
acid, phenyl andand
formate phenol.
phenol.

Figure 10. Reaction pathway for β-O-4 lignin model dimer (M8) conversion on the catalysts.
Figure 10. Reaction pathway for β-O-4 lignin model dimer (M8) conversion on the catalysts.
The catalytic oxidative solvolysis of
1-(4-hydroxy-3-methoxyphenyl)-2-(2-methoxyphenoxy)-propane-1,3-diol (M9) and
The catalytic oxidative solvolysis of 1-(4-hydroxy-3-methoxyphenyl)-2-(2-methoxyphenoxy)-
1-(3,4-dimethoxyphenyl)-2-(2-methoxyphenoxy)-propane-1,3-diol (M10) was performed over a 2:1
propane-1,3-diol (M9) and 1-(3,4-dimethoxyphenyl)-2-(2-methoxyphenoxy)-propane-1,3-diol (M10)
mixture of 1,10-phenanthroline and copper (II) sulfate pentahydrate in NaOH solution [67]. This
was performed over a 2:1 mixture
reaction was conducted in the of 1,10-phenanthroline
presence of O2 in methanoland at 80copper (II) sulfate
°C. Vanillin was thepentahydrate
major identified in NaOH
solutionproduct
[67]. This ◦ C. Vanillin was the
fromreaction
M9 whilewas conducted
veratric acid wasinthe the presence
major product offrom
O2 inM10 methanol at 80
(Figure 11). The β-O-4 bond
was cleavedproduct
major identified mainly between
from M9 thewhile
Cα–Cβ atoms
veratricandacid
different
wasmechanisms
the major productwere proposedfrombecause of
M10 (Figure 11).
the phenolic hydroxyl group para to the side chain. Blandez et al.
The β-O-4 bond was cleaved mainly between the Cα –Cβ atoms and different mechanisms were [111] found that graphene oxide
promoted the oxidative degradation of M9 mainly to guaiacol, 2-methoxyquinone, vanillic acid and
proposed because of the phenolic hydroxyl group para to the side chain. Blandez et al. [111] found that
coniferyl aldehyde. After 24 h of reaction at 140 °C in acetonitrile solvent in the presence of O2, M9
graphenewasoxide promoted
converted thealdehyde
to coniferyl oxidative degradation
(20%), M9 vanillin
vanillic acidof(2%), mainly(6%), to guaiacol, 2-methoxyquinone,
2-methoxyquinone (5%)
vanillicand
acidguaiacol
and coniferyl aldehyde. After 24 h of reaction at 140 ◦ C in acetonitrile solvent in
(96%). Reaction pathway for M9 on graphene oxide was shown in Figure 11.
the presence
{Fe-DABCO} , M9 was
of O2oxidative converted
cleavage of M10 to coniferyl
with peroxidesaldehyde
in DMSO was (20%), vanillic [60].
also reported acidGuaiacol
(2%), vanillin
(6%), 2-methoxyquinone (5%) and guaiacol (96%). Reaction pathway for M9 on°C
and veratraldehyde with 42% and 35% yield were obtained as the major products at 100 after 16 oxide
graphene
h. Thein
was shown reaction
Figure pathway for M10 on {Fe-DABCO}
11. {Fe-DABCO} oxidative is shown
cleavage in Figure
of M10 11. Transition-metal-containing
with peroxides in DMSO was
hydrotalcites (HTc) and V(acac)3/Cu(NO3)2·3H2O mixtures were employed for the catalytic cleavage
also reported [60]. Guaiacol and veratraldehyde with 42% and 35% yield were obtained as the
of M10 with O2 as an oxidant [62]. After 16 h of reaction at 135 °C in pyridine solvent in the presence
major products ◦ C after 16 h. The reaction pathway for M10 on {Fe-DABCO} is shown in
of O2, M10atwas100mainly converted to veratraldehyde (38%), veratric acid (31%) and phenol. The
Figure 11. Transition-metal-containing
reaction pathway for M10 on HTc is shown hydrotalcites
in Figure 11.(HTc) and V(acac)3 /Cu(NO3 )2 ·3H2 O mixtures
were employedA ligand
for is
theancatalytic
ion or molecule
cleavage of M10group)
(functional with O linked
2 as anto aoxidant
central metal
[62]. atom
Afterto16 structure a
h of reaction at
135 ◦ C in
coordination
pyridine complex.
solvent in Compared to otherofligands,
the presence O2 , M10 salen
wascomplexes
mainlyare relatively to
converted stable, inexpensive (38%),
veratraldehyde
and can be readily synthesized. A series of vanadium (V) complexes (e.g., Complex 3 in Figure 3)
veratric acid (31%) and phenol. The reaction pathway for M10 on HTc is shown in Figure 11.
bearing a Schiff base ligand for the non-oxidative solvolysis of a lignin model compound (M11) was
A demonstrated
ligand is an at ion or molecule (functional group) linked to a central metal atom to structure a
80 °C [112]. Different from the conversion pathway on copper complex, the β-O-4
coordination complex. Compared
bond was cleaved mainly at the to other
Cβ–O ligands,
bond (Figuresalen11).
complexes
Thus M11 arewas
relatively
converted stable, inexpensive
to alkene,
and can2-methoxyphenol
be readily synthesized. A series
and ketone after 24 h ofofreaction
vanadiumin CD3(V)CN complexes
solvent. Another (e.g., Complex
series 3 in Figure 3)
of dipicolinate
bearingvanadium (V) complexes
a Schiff base ligand for(e.g.,the Complex 2 in Figure
non-oxidative 3) forofthe
solvolysis oxidative
a lignin model degradation
compound of (M11)
1-phenyl-2-phenoxyethanol
was demonstrated at 80 ◦ C [112].(M8) was studied from
Different at 100the°C in air [97]. Differently,
conversion pathwayM8 onwas mainly
copper complex,
converted to formic acid, benzoic acid, phenol, and 2-phenoxyacetophenone (Figure 10).
the β-O-4 bond was cleaved mainly at the Cβ –O bond (Figure 11). Thus M11 was converted to
Ruthenium-triphos complexes also exhibited an outstanding catalytic activity and selectivity in the
alkene,redox-neutral
2-methoxyphenol
C–C bond and ketone
cleavage after[113].
of M10 24 h Theofuse
reaction in CD3 CN solvent.
of Ru(CO)(Cl)(H)–(triphos) [114]Another
resulted inseries of
dipicolinate vanadium
the products from the (V)Cαcomplexes (e.g., in
–Cβ bond cleavage Complex in Figure
66% yield,2whereas 3) for the
the products fromoxidative
the Cβ–O degradation
bond
of 1-phenyl-2-phenoxyethanol
cleavage were formed in 4% yield (M8)at was160 °Cstudied
after 4 h of 100 ◦ C(Figure
at reaction in air 11).[97].
It should be noted thatM8 was
Differently,
mainly converted to formic acid, benzoic acid, phenol, and 2-phenoxyacetophenone (Figure 10).
Ruthenium-triphos complexes also exhibited an outstanding catalytic activity and selectivity in the
redox-neutral C–C bond cleavage of M10 [113]. The use of Ru(CO)(Cl)(H)–(triphos) [114] resulted in
the products from the Cα –Cβ bond cleavage in 66% yield, whereas the products from the Cβ –O bond
cleavage were formed in 4% yield at 160 ◦ C after 4 h of reaction (Figure 11). It should be noted that the
Polymers 2017, 9, 240 18 of 25

Polymers 2017, 9, 240 18 of 25

phenolic group played an important role: (1) Cα –OH was required for both the Cα –Cβ and the Cβ –O
the phenolic group played an important role: (1) Cα–OH was required for both the Cα–Cβ and the Cβ–
bondPolymers 2017, 9, 240
cleavage; and (2) Cγ –OH was essential for accessing to the Cα –Cβ bond cleavage. 18 of 25
O bond cleavage; and (2) Cγ–OH was essential for accessing to the Cα–Cβ bond cleavage.
the phenolic group played an important role: (1) Cα–OH was required for both the Cα–Cβ and the Cβ–
O bond cleavage; and (2) Cγ–OH was essential for accessing to the Cα–Cβ bond cleavage.

Figure 11. Reaction pathway for β-O-4 lignin model dimer (M9, M10, and M11) conversion on the
Figure 11. Reaction pathway for β-O-4 lignin model dimer (M9, M10, and M11) conversion on
catalysts.
the catalysts.
To compare the differences in selectivity of products between Complex 2 and Complex 3, the
Figureoxidation
aerobic 11. Reaction
of apathway for β-O-4
lignin model lignin model
compound dimer (M9,with
M10, and M11) conversion on 3the
To compare
catalysts. the differences in selectivity of(M13) reacting
products between Complex
Complex 2 and
2Complex
and Complex was
3, the
investigated [115]. Complex 3 afforded the production of alkene, 2-methoxyphenol and ketone from
aerobic Coxidation of a lignin model compound (M13) reacting with Complex 2 and Complex 3 was
β–O bond cleavage, while Complex 2 gave the production of 2,6-dimethoxybenzoquinone and
investigated To compare
[115]. the differences
Complex
acrolein derivative, infrom
afforded
and 3ketone selectivity
the of products
a Cproduction between2-methoxyphenol
of alkene, Complex 2 and Complex
α–Caryl bond cleavage (Figure 12). An application of an
and ketone3, the from
Cβ –O aerobic
bond
isotopeoxidation
cleavage, of while
a lignin
tracer technique modelthat
Complex
showed compound
gave the
2Complex (M13) reacting
production
3 broke the C–O with Complex 2 andcompound
of 2,6-dimethoxybenzoquinone
bond in the model Complex 3via was and
investigated
cracking [115].
of the Complex
benzylic C 3 afforded
–H bond, the
while production
Complex 2 of alkene,
broke the
acrolein derivative, and ketone from a Cα –Caryl bond cleavage (Figure 12). An application of an
α 2-methoxyphenol
Cα –C aryl bond without and ketone
cleavage from
of
Cβ–O bond cleavage, while Complex 2 gave the production Cof 2,6-dimethoxybenzoquinone and
isotope the benzylic
tracer Cα–H bond.
technique showedFurther thatexperiments
Complexshowed 3 broke that noC–O
the α–C aryl bond cleavage was observed
bond in the model compound via
acrolein derivative,
in the aerobic and ketone
oxidation from a Cmodel
of a non-phenolic α–Caryl compound
bond cleavage (M12)(Figure
with Complex 12). An 2. Itapplication
was consistent of an
cracking of the benzylic Cα –H bond, while Complex 2 broke the Cα –Caryl bond without cleavage of
isotope tracer
with the technique
result of aerobicshowed
oxidationthatofComplex 3 broke
a β-1 linkage themodel
lignin C–O bond
compound in theonmodel
Complex compound
2 (Figurevia
the benzylic
cracking Cofα –H
8). Mechanisms bond.
the benzylic Further
of Complex
Cα–H2bond,experiments
and Complex 3showed
while Complexcatalyses 2 that nothe
during
broke CαC–C
reactions
α–C aryl bond
were
aryl bond cleavage
proposed
without awas
incleavage
recentobserved
of
in thethe
aerobic oxidation
publication
benzylic Cα[116]. of a
Parker
–H bond. non-phenolic
et al. [117]
Further model
attempted
experiments compound
to determine
showed (M12)
that no theCα–C with
effect Complex
of ligand
aryl bond
2.
structure
cleavage It was consistent
on the
was observed
with in
thetheresult
activity
aerobicof oxidation
of aerobic
vanadium oxidation
ofSchiff-base of catalysts
a non-phenolic a β-1model
linkage
(e.g., lignin model
Complex
compound compound
3) towards
(M12) on2.Complex
the degradation
with Complex 2 (Figure 8).
It wasofconsistent
β-O-4
Mechanisms
withcontained
the resultdimers.
of Complex Ortho-bulky,
of aerobic2 oxidation electron-donating
and Complex of a β-1 3 linkage ligand
catalyses ligninsubstituents
during
model were determined
reactions
compound were to make
proposed
on Complex thea recent
in
2 (Figure
most active
8). Mechanisms
publication catalyst,
[116]. Parker which
of Complex can be enhanced
2 andattempted
et al. [117] by the
Complex 3tocatalyses addition
determine duringof bulky aliphatic
reactions
the effect substituents
were proposed
of ligand such
structureinona the as activity
recent
tert-butyl
publication and adamantyl
[116]. Parker groups.
et al.(e.g., The
[117]Complex trityl-substituted
attempted3)to complex
determine was not that active due to the
of vanadium Schiff-base catalysts towards the the effect of ligand
degradation of β-O-4 structure
containedon the dimers.
enhanced steric hindrance, confining the substrate access to the active sites of catalyst [117].
activity of
Ortho-bulky, vanadium Schiff-base
electron-donating ligandcatalysts (e.g., were
substituents Complex 3) towards
determined the degradation
to make of β-O-4
the most active catalyst,
contained dimers. Ortho-bulky, electron-donating ligand substituents were determined to make the
which can be enhanced by the addition of bulky aliphatic substituents such as tert-butyl and adamantyl
most active catalyst, which can be enhanced by the addition of bulky aliphatic substituents such as
groups. The trityl-substituted complex was not that active due to the enhanced steric hindrance,
tert-butyl and adamantyl groups. The trityl-substituted complex was not that active due to the
confining the substrate
enhanced accessconfining
steric hindrance, to the active sites of catalyst
the substrate access to[117].
the active sites of catalyst [117].

Figure 12. Reaction pathway for β-O-4 lignin model dimer (M13) conversion on the catalysts.

Figure 12. Reaction pathway for β-O-4 lignin model dimer (M13) conversion on the catalysts.
Figure 12. Reaction pathway for β-O-4 lignin model dimer (M13) conversion on the catalysts.
Polymers 2017, 9, 240 19 of 25

The oxidation of lignin-related model compounds, 2-phenoxy-1-phenylethanol (M8) and


2-(4-methoxyphenoxy)-1-phenylethanol
Polymers 2017, 9, 240 (M14), on a Pd/CeO2 catalyst was performed 19 inofmethanol
25
under O2 [74]. The reaction mechanism for the conversion of M8 over Pd/CeO2 in methanol with
the presence The of oxidation
Polymers 2017,
of lignin-related
O9,2240is shown in Schememodel 3. The compounds,
Cα –OH of 2-phenoxy-1-phenylethanol
2-phenoxy-1-phenylethanol (M8)19 of
andwas first
25
2-(4-methoxyphenoxy)-1-phenylethanol (M14), on a Pd/CeO2 catalyst was performed in methanol
oxidized due to the catalytic function of Pd nanoparticles, to form 2-phenoxy-1-phenylethanone,
under O2 [74].
The The reaction
oxidation mechanism model
of lignin-related for the conversion
compounds,of 2-phenoxy-1-phenylethanol
M8 over Pd/CeO2 in methanol(M8) withandthe
whose β-O-4 bond became weakened due toCthe
presence of O2 is shown in Scheme 3. The α–OH
presence of Cα =O. The subsequent hydrogenolysis
2-(4-methoxyphenoxy)-1-phenylethanol (M14), on of 2-phenoxy-1-phenylethanol
a Pd/CeO 2 catalyst was performed was first oxidized
in methanol
of the β-O-4
due tobond
under the
O 2 [74].
ofThe2-phenoxy-1-phenylethanone
catalytic functionmechanism
reaction of Pd nanoparticles, to resulted in the form of acetophenone
form 2-phenoxy-1-phenylethanone,
for the conversion of M8 over Pd/CeO2 in methanol whose and
β-O-4
with the
phenol
over thepresence
Pd/CeO
bond becameof 2Ocatalyst.
is shownOn
2weakened due
in the other
to the
Scheme hand,
presence
3. The ofthe
Cα–OH Cαof oxidative
=O. cracking
The subsequent of the β-O-4
hydrogenolysis
2-phenoxy-1-phenylethanol ofbond
was first the over CeO2
β-O-4
oxidized
bondto
might lead
due toofthe2-phenoxy-1-phenylethanone
the formation
catalytic of phenol
function and resulted in
an oxidized
of Pd nanoparticles, the form
to form of acetophenone
intermediate and phenol over
(e.g., 2-hydroxyacetophenone).
2-phenoxy-1-phenylethanone, whose the
β-O-4
Pd/CeO
bond
Meanwhile, 2 catalyst.
abecame
small amountOn the
weakened other
ofdue tohand, the oxidative
the presence
acetophenone canofbe cracking
Cαproduced
=O. of during
the β-O-4
The subsequent bond over CeO
hydrogenolysis
this process. 2 might
of lead
the β-O-4
Finally, the oxidized
to the of
bond formation of phenol and an oxidized
2-phenoxy-1-phenylethanone intermediate
resulted in the (e.g.,of
form 2-hydroxyacetophenone).
acetophenone and phenolMeanwhile,
over the
intermediate experienced the C–C bond cracking, producing benzoic acid and methyl benzoate. Phenol,
a small2 amount
Pd/CeO catalyst. On of acetophenone
the other hand,can be produced
the oxidative during
cracking this
of the process.
β-O-4 bond Finally,
over CeOthe oxidized
2 might lead
acetophenone and experienced
intermediate methyl benzoate the C–C were
bond produced
cracking, with thebenzoic
producing yield of 48%,
acid and 38% andbenzoate.
methyl 14% from the
to the formation of phenol and an oxidized intermediate (e.g., 2-hydroxyacetophenone). Meanwhile,
oxidation of M8,
aPhenol,
small whileof4-methoxyphenol,
acetophenone
amount and methyl benzoate
acetophenone can be acetophenone
were produced
produced and methyl
withthis
during the yield benzoate
of 48%,
process. 38%
Finally,were
and produced
the 14% from with
oxidized
yields ofintermediate
82%,
the 38% experienced
oxidation and
of M8,40% from
while the oxidation
4-methoxyphenol, of M14 at 185and
acetophenone ◦ C methyl
for 24 h. and
It could
benzoate beproduced
were concluded that
the C–C bond cracking, producing benzoic acid methyl benzoate.
with yields
ortho-methoxy
Phenol, of 82%, 38%
substituents
acetophenone and
and 40% from
further
methyl the oxidation
elevated
benzoate of M14with
the performance
were produced at 185 of°Cyield
the for Pd/CeO
the 24
of h. It could
48%, 38% be concluded
catalyst,
and 14% which was
from
2
thatoxidation
the ortho-methoxy of substituents
M8, while further elevatedacetophenone
4-methoxyphenol, the performance
consistent with the other report stating that ortho-methoxy substituents could weaken the was
and ofmethyl
the Pd/CeO 2 catalyst,
benzoate were which
produced
stability of
consistent
with yieldswithof 82%,the other
38% andreport
40%stating thatoxidation
from the ortho-methoxy
of M14substituents
at 185 °C for could
24 h.weaken
It couldthe
be stability
concluded of
the aromatic ring, promoting its over-oxidation
the aromatic ring, promoting its over-oxidation [64].
[64].
that ortho-methoxy substituents further elevated the performance of the Pd/CeO2 catalyst, which was
consistent with the other report stating that ortho-methoxy substituents could weaken the stability of
the aromatic ring, promoting its over-oxidation [64].

Scheme 3. Proposed pathway for oxidative conversion of 2-phenoxy-1-phenylethanol (M8) over a


Scheme 3. Proposed pathway for oxidative conversion of 2-phenoxy-1-phenylethanol (M8) over a
Pd/CeO2 catalyst [74], reproduced with permission from Royal Society of Chemistry.
Pd/CeO2 catalyst [74], reproduced with permission from Royal Society of Chemistry.
Scheme 3. Proposed pathway for oxidative conversion of 2-phenoxy-1-phenylethanol (M8) over a
Moreover,
Pd/CeO Wang et al. [118] reported a two-step strategy for lignin C–C bond conversion via a
2 catalyst [74], reproduced with permission from Royal Society of Chemistry.
β-O-4 alcohol
Moreover, Wang oxidation to ketone
et al. [118] over the
reported a VOSO
two-step4/TEMPO [(2,2,6,6-tetramethylpiperidin-1-yl)oxyl)]
strategy for lignin C–C bond conversion via a
catalyst, followed
Moreover, by aet ketone
Wang al. [118]oxidation
reported over
a Cu(OAc)
two-step 2/1,10-phenanthroline
strategy for lignin C–C bondcatalyst to acidsvia
conversion anda
β-O-4 alcohol oxidation to ketone over the VOSO4 /TEMPO [(2,2,6,6-tetramethylpiperidin-1-yl)oxyl)]
phenols.
β-O-4 The proposed
alcohol oxidationreaction
to ketone mechanism is shown
over the VOSO in Scheme
4/TEMPO 4. The oxidation of a Cα–OH alcohol
[(2,2,6,6-tetramethylpiperidin-1-yl)oxyl)]
catalyst,tofollowed
a ketone by a ketone
activated Coxidation overCu(OAc)
Cu(OAc) 2 /1,10-phenanthroline catalyst to acids and
catalyst, followed by athe
ketone β–Hoxidation
bond. The over Cu(OAc) 2/1,10-phenanthroline
2/1,10-phenanthroline reacted
catalystwith
to oxygen to
acids and
phenols.phenols.
Thea proposed
form reaction
copper-oxo-bridged mechanism
dimer. The is shown
activation of in
the CScheme
-C bond 4. The
in the oxidation
form
The proposed reaction mechanism is shown in Scheme 4. The oxidation of a Cα–OH alcohol
α β of a of a
hydroxylC α –OH
ketone alcohol
to a ketone
to activated
structure-like
a ketone the Cthe
activated β –H
intermediate bond.
β–H The
Csignificantly
bond. The Cu(OAc)
decreased
Cu(OAc)its /1,10-phenanthroline
bond energy, resulting
22/1,10-phenanthroline inreacted
reacted the with oxygen
formation
with oxygen of
to to
benzoic acid and
form a copper-oxo-bridged phenyl formate.
dimer. Transfer
TheThe of
activation a hydroxyl
activationof ofthe group
Cαα-C-C generated benzoic acid, restoring the
form a copper-oxo-bridged dimer. the C β bond
β bond in form
in the the form of a hydroxyl
of a hydroxyl ketone ketone
initial copper complex.
structure-like Finally,
intermediate the oxidative
significantly decarboxylation
decreased its bond of phenyl
energy, formate
resulting generated phenol
structure-like intermediate significantly decreased its bond energy, resulting ininthe
theformation
formation ofbenzoic
of
and CO2acid
benzoic . and phenyl formate. Transfer of a hydroxyl group generated benzoic acid, restoring the
acid and phenyl formate. Transfer of a hydroxyl group generated benzoic acid, restoring the initial
initial copper complex. Finally, the oxidative decarboxylation of phenyl formate generated phenol
copper complex.
and CO2.
Finally, the oxidative decarboxylation of phenyl formate generated phenol and CO2 .

Scheme 4. Proposed reaction mechanism for two-step, catalytic C–C bond oxidative cleavage process
[118], reproduced with permission from American Chemical Society.
Scheme 4. Proposed reaction mechanism for two-step, catalytic C–C bond oxidative cleavage process
Scheme[118], Proposed with
4. reproduced reaction mechanism
permission for two-step,
from American catalytic C–C bond oxidative cleavage
Chemical Society.
process [118], reproduced with permission from American Chemical Society.
Polymers 2017, 9, 240 20 of 25

The oxidative solvolysis of lignin-related model compounds with homogeneous catalysts was
mostly reported for enhancing the production of aromatic aldehydes, aromatic ketones, benzoquinones,
carboxylic acids, etc. The oxidative solvolysis of model compounds with heterogeneous catalysts may
achieve a high selectivity of specific products. However, the performance of both homogeneous and
heterogeneous catalysts in the cleavage of inter-unit linkages is still ambiguous in the oxidative solvent
system, confining the understanding of inherent mechanism of whole lignin depolymerization process.

5. Summary and Outlook


The catalytic oxidation of lignin in solvent systems is attracting more and more attention for
producing renewable aromatic chemicals. A number of functionalized monomeric products were
obtained from oxidative depolymerization of lignin, including vanillin, vanillic acid, syringaldehyde
and syringic acid. Considering the ambiguities in the inherent mechanism of (catalytic) oxidative
solvolysis of lignin process and the limited selectivity and yield of renewable chemicals, several issues
should be addressed: (1) identification of the oligomer structure for specifying the inherent oxidative
depolymerization mechanism of natural lignin; (2) design of highly efficient catalyst for C–C type
inter-unit linkages; (3) strategies for suppressing the repolymerization of the produced fragments;
(4) low-cost separation and purification technology of the specific aromatic compounds from an LDP
system; and (5) stability of the heterogeneous catalysts during LDP and its reactivation.

Acknowledgments: The authors greatly acknowledge the funding support from the projects supported by
National Natural Science Foundation of China (Grant Nos. 51676034 and 51628601) and Natural Science
Foundation of Jiangsu Province (project reference: BK20161423).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Li, C.; Zhao, X.; Wang, A.; Huber, G.W.; Zhang, T. Catalytic transformation of lignin for the production of
chemicals and fuels. Chem. Rev. 2015, 115, 11559–11624. [CrossRef] [PubMed]
2. Sawin, J.L.; Sverrisson, F. Renewables 2014: Global Status Report; REN21 Secretariat: Paris, France, 2014; p. 214.
3. Lange, H.; Decina, S.; Crestini, C. Oxidative upgrade of lignin—Recent routes reviewed. Eur. Polym. J. 2013,
49, 1151–1173. [CrossRef]
4. Huber, G.W.; Iborra, S.; Corma, A. Synthesis of transportation fuels from biomass: Chemistry, catalysts, and
engineering. Chem. Rev. 2006, 106, 4044–4098. [CrossRef] [PubMed]
5. Melero, J.A.; Iglesias, J.; Garcia, A. Biomass as renewable feedstock in standard refinery units. Feasibility,
opportunities and challenges. Energy Environ. Sci. 2012, 5, 7393–7420. [CrossRef]
6. Gallezot, P. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 2012, 41, 1538–1558.
[CrossRef] [PubMed]
7. Tuck, C.O.; Perez, E.; Horvath, I.T.; Sheldon, R.A.; Poliakoff, M. Valorization of biomass: Deriving more
value from waste. Science 2012, 337, 695–699. [CrossRef] [PubMed]
8. Besson, M.; Gallezot, P.; Pinel, C. Conversion of biomass into chemicals over metal catalysts. Chem. Rev.
2014, 114, 1827–1870. [CrossRef] [PubMed]
9. Mamman, A.S.; Lee, J.-M.; Kim, Y.-C.; Hwang, I.T.; Park, N.-J.; Hwang, Y.K.; Chang, J.-S.; Hwang, J.-S.
Furfural: Hemicellulose/xylose derived biochemical. Biofuels Bioprod. Biorefin. 2008, 2, 438–454. [CrossRef]
10. Doherty, W.O.S.; Mousavioun, P.; Fellows, C.M. Value-adding to cellulosic ethanol: Lignin polymers.
Ind. Crops Prod. 2011, 33, 259–276. [CrossRef]
11. Serrano-Ruiz, J.C.; Luque, R.; Sepúlveda-Escribanoa, A. Transformations of biomass-derived platform
molecules: From high added-value chemicals to fuelsvia aqueous-phase processing. Chem. Soc. Rev. 2011, 40,
5266–5281. [CrossRef] [PubMed]
12. Rinaldi, R.; Palkovits, R.; Schüth, F. Depolymerization of cellulose using solid catalysts in ionic liquids.
Angew. Chem. Int. Ed. 2008, 47, 8047–8050. [CrossRef] [PubMed]
13. Gosselink, R.J.A.; de Jong, E.; Guran, B.; Abächerli, A. Co-ordination network for lignin-standardisation,
production and applications adapted to market requirements. Ind. Crops Prod. 2004, 20, 121–129. [CrossRef]
Polymers 2017, 9, 240 21 of 25

14. Calvo-Flores, F.G.; Dobado, J.A. Lignin as renewable raw material. ChemSusChem 2010, 3, 1227–1235.
[CrossRef] [PubMed]
15. Ragauskas, A.J.; Beckham, G.T.; Biddy, M.J.; Chandra, R.; Chen, F.; Davis, M.F.; Davison, B.H.; Dixon, R.A.;
Gilna, P.; Keller, M.; et al. Lignin valorization: Improving lignin processing in the biorefinery. Science 2014,
344, 709–721. [CrossRef] [PubMed]
16. Naik, S.N.; Goud, V.V.; Rout, P.K.; Dalai, A.K. Production of first and second generation biofuels:
A comprehensive review. Renew. Sustain. Energy Rev. 2010, 14, 578–597. [CrossRef]
17. Hu, L.H.; Pan, H.; Zhou, Y.H.; Zhang, M. Methods to improve lignin’s reactivity as a phenol substitute and
as replacement for other phenolic compounds: A brief review. BioResources 2011, 6, 3515–3525.
18. Deuss, P.J.; Barta, K. From models to lignin: Transition metal catalysis for selective bond cleavage reactions.
Coord. Chem. Rev. 2016, 306, 510–532. [CrossRef]
19. Kang, S.; Li, X.; Fan, J.; Chang, J. Hydrothermal conversion of lignin: A review. Renew. Sustain. Energy Rev.
2013, 27, 546–558. [CrossRef]
20. Carpenter, D.; Westover, T.L.; Czernik, S.; Jablonski, W. Biomass feedstocks for renewable fuel production:
A review of the impacts of feedstock and pretreatment on the yield and product distribution of fast pyrolysis
bio-oils and vapors. Green Chem. 2014, 16, 384–406. [CrossRef]
21. Shen, D.; Xiao, R.; Gu, S.; Luo, K. The pyrolytic behavior of cellulose in lignocellulosic biomass: A review.
RSC Adv. 2011, 1, 1641–1660. [CrossRef]
22. Azadi, P.; Inderwildi, O.R.; Farnood, R.; King, D.A. Liquid fuels, hydrogen and chemicals from lignin:
A critical review. Renew. Sustain. Energy Rev. 2013, 21, 506–523. [CrossRef]
23. Shen, D.; Jin, W.; Hu, J.; Xiao, R.; Luo, K. An overview on fast pyrolysis of the main constituents in
lignocellulosic biomass to valued-added chemicals: Structures, pathways and interactions. Renew. Sustain.
Energy Rev. 2015, 51, 761–774. [CrossRef]
24. Chatel, G.; Rogers, R.D. Review: Oxidation of lignin using ionic liquids—An innovative strategy to produce
renewable chemicals. ACS Sustain. Chem. Eng. 2014, 2, 322–339. [CrossRef]
25. Behling, R.; Valange, S.; Chatel, G. Heterogeneous catalytic oxidation for lignin valorization into valuable
chemicals: What results? What limitations? What trends? Green Chem. 2016, 18, 1839–1854. [CrossRef]
26. Yokoyama, T. Revisiting the mechanism of β-O-4 bond cleavage during acidolysis of lignin. Part 6: A review.
J. Wood Chem. Technol. 2014, 35, 27–42. [CrossRef]
27. Mota, M.I.F.; Rodrigues Pinto, P.C.; Loureiro, J.M.; Rodrigues, A.E. Recovery of vanillin and syringaldehyde
from lignin oxidation: A review of separation and purification processes. Sep. Purif. Technol. 2015, 45,
227–259. [CrossRef]
28. Xu, C.; Arancon, R.A.D.; Labidid, J.; Luque, R. Lignin depolymerisation strategies—Towards valuable
chemicals and fuels. Chem. Soc. Rev. 2014, 43, 7485–7500. [CrossRef] [PubMed]
29. Deuss, P.J.; Barta, K.; de Vries, J.G. Homogeneous catalysis for the conversion of biomass and biomass-derived
platform chemicals. Catal. Sci. Technol. 2014, 4, 1174–1196. [CrossRef]
30. Zakzeski, J.; Bruijnincx, P.C.A.; Jongerius, A.L.; Weckhuysen, B.M. The catalytic valorization of lignin for the
production of renewable chemicals. Chem. Rev. 2010, 110, 3552–3599. [CrossRef] [PubMed]
31. Collinson, S.R.; Thielemans, W. The catalytic oxidation of biomass to new materials focusing on starch,
cellulose and lignin. Coord. Chem. Rev. 2010, 254, 1854–1870. [CrossRef]
32. Dai, J.; Patti, A.F.; Saito, K. Recent developments in chemical degradation of lignin: Catalytic oxidation and
ionic liquids. Tetrahedron Lett. 2016, 57, 4945–4951. [CrossRef]
33. Ma, R.; Xu, Y.; Zhang, X. Catalytic oxidation of biorefinery lignin to value-added chemicals to support
sustainable biofuel production. ChemSusChem 2015, 8, 24–51. [CrossRef] [PubMed]
34. Gupta, K.C.; Kumar Sutar, A.; Lin, C.-C. Polymer-supported schiff base complexes in oxidation reactions.
Coord. Chem. Rev. 2009, 253, 1926–1946. [CrossRef]
35. Amadio, E.; Di Lorenzo, R.; Zonta, C.; Licini, G. Vanadium catalyzed aerobic carbon–carbon cleavage.
Coord. Chem. Rev. 2015, 301–302, 147–162. [CrossRef]
36. Stewart, D. Lignin as a base material for materials applications: Chemistry, application and economics.
Ind. Crops Prod. 2008, 27, 202–207. [CrossRef]
37. Crestini, C.; Crucianelli, M.; Orlandi, M.; Saladino, R. Oxidative strategies in lignin chemistry: A new
environmental friendly approach for the functionalisation of lignin and lignocellulosic fibers. Catal. Today
2010, 156, 8–22. [CrossRef]
Polymers 2017, 9, 240 22 of 25

38. Dong, C.; Feng, C.; Liu, Q.; Shen, D.; Xiao, R. Mechanism on microwave-assisted acidic solvolysis of
black-liquor lignin. Bioresour. Technol. 2014, 162, 136–141. [CrossRef] [PubMed]
39. Shen, D.; Liu, N.; Dong, C.; Xiao, R.; Gu, S. Catalytic solvolysis of lignin with the modified husys in formic
acid assisted by microwave heating. Chem. Eng. J. 2015, 270, 641–647. [CrossRef]
40. Nimz, H. Beech lignin—Proposal of a constitutional scheme. Angew. Chem. Int. Ed. 1974, 13, 313–321.
[CrossRef]
41. Erickson, M.; Larsson, S.; Miksche, G.E. Analysis using gas-chromatography of lignin oxidation-products. 8.
Structure of spruce lignin. Acta Chem. Scand. 1973, 27, 903–914. [CrossRef]
42. Hage, R.; Lienke, A. Applications of transition-metal catalysts to textile and wood-pulp bleaching.
Angew. Chem. Int. Ed. 2006, 45, 206–222. [CrossRef] [PubMed]
43. Hu, J.; Xiao, R.; Shen, D.; Zhang, H. Structural analysis of lignin residue from black liquor and its thermal
performance in thermogravimetric-fourier transform infrared spectroscopy. Bioresour. Technol. 2013, 128,
633–639. [CrossRef] [PubMed]
44. Shen, D.; Cheng, C.; Liu, N.; Xiao, R. Lignin depolymerization (LDP) with solvolysis for selective production
of renewable aromatic chemicals. In Production of Biofuels and Chemicals from Lignin; Fang, Z., Smith, R.L., Jr.,
Eds.; Springer: Berlin, Germany, 2016; p. 290.
45. Choi, Y.S.; Singh, R.; Zhang, J.; Balasubramanian, G.; Sturgeon, M.R.; Katahira, R.; Chupka, G.; Beckham, G.T.;
Shanks, B.H. Pyrolysis reaction networks for lignin model compounds: Unraveling thermal deconstruction
of β-O-4 and α-O-4 compounds. Green Chem. 2016, 18, 1762–1773. [CrossRef]
46. Forsythe, W.G.; Garrett, M.D.; Hardacre, C.; Nieuwenhuyzen, M.; Sheldrake, G.N. An efficient and flexible
synthesis of model lignin oligomers. Green Chem. 2013, 15, 3031–3038. [CrossRef]
47. Sedai, B.; Díaz-Urrutia, C.; Baker, R.T.; Wu, R.; Silks, L.A.P.; Hanson, S.K. Aerobic oxidation of β-1 lignin
model compounds with copper and oxovanadium catalysts. ACS Catal. 2013, 3, 3111–3122. [CrossRef]
48. Reddy, G.V.B.; Sridhar, M.; Gold, M.H. Cleavage of nonphenolic β-1 diarylpropane lignin model dimers by
manganese peroxidase from phanerochaete chrysosporium. Evidence for a hydrogen abstraction mechanism.
Eur. J. Biochem. 2003, 270, 284–292. [CrossRef] [PubMed]
49. Tran, F.; Lancefield, C.S.; Kamer, P.C.J.; Lebl, T.; Westwood, N.J. Selective modification of the β-β linkage in
DDQ-treated kraft lignin analysed by 2D NMR spectroscopy. Green Chem. 2015, 17, 244–249. [CrossRef]
50. Ouyang, X.-P.; Liu, C.-L.; Pang, Y.-X.; Qiu, X.-Q. Synthesis of a trimeric lignin model compound composed of
α-O-4 and β-O-4 linkages under microwave irradiation. Chin. Chem. Lett. 2013, 24, 1091–1094. [CrossRef]
51. Mimms, A.; Kocurek, M.; Pyatte, J.; Wright, E. (Eds.) Kraft Pulping Chemistry and Process; Tappi Press:
Peachtree Corners, GA, USA, 1989.
52. Glasser, W.; Dave, V.; Frazier, C. Molecular weight distribution of (semi-)commercial lignin derivatives.
J. Wood Chem. Technol. 1993, 13, 545–559. [CrossRef]
53. Pan, X.J.; Arato, C.; Gilkes, N.; Gregg, D.; Mabee, W.; Pye, K.; Xiao, Z.Z.; Zhang, X.; Saddler, J. Biorefining of
softwoods using ethanol organosolv pulping: Preliminary evaluation of process streams for manufacture of
fuel-grade ethanol and co-products. Biotechnol. Bioeng. 2005, 90, 473–481. [CrossRef] [PubMed]
54. Scholze, B.; Meier, D. Characterization of the water-insoluble fraction from pyrolysis oil (pyrolytic lignin).
Part I. PY-GC/MS, FTIR, and functional groups. J. Anal. Appl. Pyrolysis 2001, 60, 41–54. [CrossRef]
55. Scholze, B.; Hanser, C.; Meier, D. Characterization of the water-insoluble fraction from fast pyrolysis liquids
(pyrolytic lignin): Part II. GPC, carbonyl goups, and 13 C NMR. J. Anal. Appl. Pyrolysis 2001, 58, 387–400.
[CrossRef]
56. Samuel, R.; Pu, Y.; Raman, B.; Ragauskas, A. Structural characterization and comparison of switchgrass
ball-milled lignin before and after dilute acid pretreatment. Appl. Biochem. Biotechnol. 2010, 162, 62–74.
[CrossRef] [PubMed]
57. Saha, B.; Iten, L.; Cotta, M.; Wu, Y. Dilute acid pretreatment, enzymatic saccharification and fermentation of
wheat straw to ethanol. Process Biochem. 2005, 40, 3693–3700. [CrossRef]
58. Chan, J.M.W.; Bauer, S.; Sorek, H.; Sreekumar, S.; Wang, K.; Toste, F.D. Studies on the vanadium-catalyzed
nonoxidative depolymerization of miscanthus giganteus-derived lignin. ACS Catal. 2013, 3, 1369–1377.
[CrossRef]
59. Díaz-Urrutia, C.; Chen, W.-C.; Crites, C.-O.; Daccache, J.; Korobkov, I.; Baker, R.T. Towards lignin valorisation:
Comparing homogeneous catalysts for the aerobic oxidation and depolymerisation of organosolv lignin.
RSC Adv. 2015, 5, 70502–70511. [CrossRef]
Polymers 2017, 9, 240 23 of 25

60. Mottweiler, J.; Rinesch, T.; Besson, C.; Buendia, J.; Bolm, C. Iron-catalysed oxidative cleavage of lignin and
β-O-4 lignin model compounds with peroxides in DMSO. Green Chem. 2015, 17, 5001–5008. [CrossRef]
61. Zhou, X.-F.; Lu, X.-J. Co(salen) supported on graphene oxide for oxidation of lignin. J. Appl. Polym. Sci. 2016,
44133, 1–9. [CrossRef]
62. Mottweiler, J.; Puche, M.; Rauber, C.; Schmidt, T.; Concepcion, P.; Corma, A.; Bolm, C. Copper-and
vanadium-Catalyzed oxidative cleavage of lignin using dioxygen. ChemSusChem 2015, 8, 2106–2113.
[CrossRef] [PubMed]
63. Crestini, C.; Caponi, M.C.; Argyropoulos, D.S.; Saladino, R. Immobilized methyltrioxo rhenium
(MTO)/H2O2 systems for the oxidation of lignin and lignin model compounds. Bioorg. Med. Chem.
2006, 14, 5292–5302. [CrossRef] [PubMed]
64. Gao, Y.; Zhang, J.; Chen, X.; Ma, D.; Yan, N. A metal-free, carbon-based catalytic system for the oxidation of
lignin model compounds and lignin. ChemPlusChem 2014, 79, 825–834. [CrossRef]
65. Demesa, A.G.; Laari, A.; Turunen, I.; Sillanpää, M. Alkaline partial wet oxidation of lignin for the production
of carboxylic acids. Chem. Eng. Technol. 2015, 38, 2270–2278. [CrossRef]
66. Ouyang, X.; Ruan, T.; Qiu, X. Effect of solvent on hydrothermal oxidation depolymerization of lignin for the
production of monophenolic compounds. Fuel Process. Technol. 2016, 144, 181–185. [CrossRef]
67. Azarpira, A.; Ralph, J.; Lu, F. Catalytic alkaline oxidation of lignin and its model compounds: A pathway to
aromatic biochemicals. BioEnergy Res. 2013, 7, 78–86. [CrossRef]
68. Voitl, T.; von Rohr, P.R. Oxidation of lignin using aqueous polyoxometalates in the presence of alcohols.
ChemSusChem 2008, 1, 763–769. [CrossRef] [PubMed]
69. Voitl, T.; von Rohr, P.R. Demonstration of a process for the conversion of kraft lignin into vanillin and methyl
vanillate by acidic oxidation in aqueous methanol. Ind. Eng. Chem. Res. 2010, 49, 520–525. [CrossRef]
70. Werhan, H.; Mir, J.M.; Voitl, T.; von Rohr, P.R. Acidic oxidation of kraft lignin into aromatic monomers
catalyzed by transition metal salts. Holzforschung 2011, 65. [CrossRef]
71. Napoly, F.; Kardos, N.; Jean-Gérard, L.; Goux-Henry, C.; Andrioletti, B.; Draye, M. H2 O2 -mediated kraft
lignin oxidation with readily available metal salts: What about the effect of ultrasound? Ind. Eng. Chem. Res.
2015, 54, 6046–6051. [CrossRef]
72. Partenheimer, W. The aerobic oxidative cleavage of lignin to produce hydroxyaromatic benzaldehydes and
carboxylic acids via metal/bromide catalysts in acetic acid/water mixtures. Adv. Synth. Catal. 2009, 351,
456–466. [CrossRef]
73. Gu, X.; He, M.; Shi, Y.; Li, Z. La-modified SBA-15/H2 O2 systems for the microwave assisted oxidation of
organosolv beech wood lignin. Maderas Cienc. Tecnol. 2012, 14, 31–41. [CrossRef]
74. Deng, W.; Zhang, H.; Wu, X.; Li, R.; Zhang, Q.; Wang, Y. Oxidative conversion of lignin and lignin model
compounds catalyzed by CeO2-supported pd nanoparticles. Green Chem. 2015, 17, 5009–5018. [CrossRef]
75. Deng, H.; Lin, L.; Sun, Y.; Pang, C.; Zhuang, J.; Ouyang, P.; Li, Z.; Liu, S. Perovskite-type oxide LaMnO3 :
An efficient and recyclable heterogeneous catalyst for the wet aerobic oxidation of lignin to aromatic
aldehydes. Catal. Lett. 2008, 126, 106–111. [CrossRef]
76. Deng, H.; Lin, L.; Sun, Y.; Pang, C.; Zhuang, J.; Ouyang, P.; Li, J.; Liu, S. Activity and stability of
perovskite-type oxide LaCoO3 catalyst in lignin catalytic wet oxidation to aromatic aldehydes process.
Energy Fuels 2009, 23, 19–24. [CrossRef]
77. Dawange, M.; Galkin, M.V.; Samec, J.S.M. Selective aerobic benzylic alcohol oxidation of lignin model
compounds: Route to aryl ketones. ChemCatChem 2015, 7, 401–404. [CrossRef]
78. Zhu, R.; Wang, B.; Cui, M.; Deng, J.; Li, X.; Ma, Y.; Fu, Y. Chemoselective oxidant-free dehydrogenation of
alcohols in lignin using Cp*Ir catalysts. Green Chem. 2016, 18, 2029–2036. [CrossRef]
79. Cedeno, D.; Bozell, J. Catalytic oxidation of para-substituted phenols with cobalt-schiff base
complexes/O2 -selective conversion of syringyl and guaiacyl lignin models to benzoquinones.
Tetrahedron Lett. 2012, 53, 2380–2383. [CrossRef]
80. Badamali, S.; Luque, R.; Clark, J.; Breeden, S. Co(salen)/SBA-15 catalysed oxidation of a β-O-4 phenolic
dimer under microwave irradiation. Catal. Commun. 2011, 12, 993–995. [CrossRef]
81. Crestini, C.; Pro, P.; Neri, V.; Saladino, R. Methyltrioxorhenium: A new catalyst for the activation of hydrogen
peroxide to the oxidation of lignin and lignin model compounds. Bioorg. Med. Chem. 2005, 13, 2569–2578.
[CrossRef] [PubMed]
Polymers 2017, 9, 240 24 of 25

82. Ouyang, X.; Lin, Z.; Deng, Y.; Yang, D.; Qiu, X. Oxidative degradation of soda lignin assisted by microwave
irradiation. Chin. J. Chem. Eng. 2010, 18, 695–702. [CrossRef]
83. Wu, A.; Lauzon, J.M.; Andriani, I.; James, B.R. Breakdown of lignins, lignin model compounds, and
hydroxy-aromatics, to C1 and C2 chemicals via metal-free oxidation with peroxide or persulfate under mild
conditions. RSC Adv. 2014, 4, 17931–17934. [CrossRef]
84. Maziero, P.; de Oliveira Neto, M.; Machado, D.; Batista, T.; Schmitt Cavalheiro, C.C.; Neumann, M.G.;
Craievich, A.F.; de Moraes Rocha, G.J.; Polikarpov, I.; Gonçalves, A.R. Structural features of lignin obtained
at different alkaline oxidation conditions from sugarcane bagasse. Ind. Crops Prod. 2012, 35, 61–69. [CrossRef]
85. Chen, C.; Pakdel, H.; Roy, C. Production of monomeric phenols by thermochemical conversion of biomass:
A review. Bioresour. Technol. 2001, 79, 277–299. [CrossRef]
86. Bujanovic, B.; Reiner, R.S.; Ralph, S.A.; Atalla, R.H. Polyoxometalate delignification of birch kraft pulp and
effect on residual lignin. J. Wood Chem. Technol. 2011, 31, 121–141. [CrossRef]
87. Sik Kim, Y.; Chang, H.-m.; Kadla, J.F. Polyoxometalate (POM) oxidation of milled wood lignin (MWL).
J. Wood Chem. Technol. 2007, 27, 225–241. [CrossRef]
88. Partenheimer, W. Methodology and scope of metal bromide autoxidation of hydrocarbons. Catal. Today 1995,
23, 69–158. [CrossRef]
89. Rothenberg, G. Catalysis: Concepts and Green Applications; Wiely-VCH: Weinheim, Germany, 2008.
90. Ma, Y.; Du, Z.; Liu, J.; Xia, F.; Xu, J. Selective oxidative C–C bond cleavage of a lignin model compound in
the presence of acetic acid with a vanadium catalyst. Green Chem. 2015, 17, 4968–4973. [CrossRef]
91. Pan, J.; Fu, J.; Lu, X. Microwave-assisted oxidative degradation of lignin model compounds with metal salts.
Energy Fuels 2015, 29, 4503–4509. [CrossRef]
92. Shilpy, M.; Ehsan, M.A.; Ali, T.H.; Abd Hamid, S.B.; Ali, M.E. Performance of cobalt titanate towards H2 O2
based catalytic oxidation of lignin model compound. RSC Adv. 2015, 5, 79644–79653. [CrossRef]
93. Tumula, V.; Bondwal, S.; Bisht, P.; Pendem, C.; Kumar, J. Oxidation of sulfides to sulfones with hydrogen
peroxide in the presence of acetic acid and amberlyst 15. React. Kinet. Mech. Catal. 2012, 107, 449–466.
[CrossRef]
94. Badamali, S.; Clark, J.; Breeden, S. Microwave assisted selective oxidation of lignin model phenolic monomer
over SBA-15. Catal. Commun. 2008, 9, 2168–2170. [CrossRef]
95. Badamali, S.; Luque, R.; Clark, J.; Breeden, S. Microwave assisted oxidation of a lignin model phenolic
monomer using Co(salen)/SBA-15. Catal. Commun. 2009, 10, 1010–1013. [CrossRef]
96. Badamali, S.K.; Luque, R.; Clark, J.H.; Breeden, S.W. Unprecedented oxidative properties of mesoporous
silica materials: Towards microwave-assisted oxidation of lignin model compounds. Catal. Commun. 2013,
31, 1–4. [CrossRef]
97. Hanson, S.K.; Baker, R.T.; Gordon, J.C.; Scott, B.L.; Thorn, D.L. Aerobic oxidation of lignin models using a
base metal vanadium catalyst. Inorg. Chem. 2010, 49, 5611–5618. [CrossRef] [PubMed]
98. Mate, V.R.; Shirai, M.; Rode, C.V. Heterogeneous Co3 O4 catalyst for selective oxidation of aqueous veratryl
alcohol using molecular oxygen. Catal. Commun. 2013, 33, 66–69. [CrossRef]
99. Mate, V.R.; Jha, A.; Joshi, U.D.; Patil, K.R.; Shirai, M.; Rode, C.V. Effect of preparation parameters on
characterization and activity of Co3 O4 catalyst in liquid phase oxidation of lignin model substrates.
Appl. Catal. A 2014, 487, 130–138. [CrossRef]
100. Gu, X.; He, M.; Shi, Y.; Li, Z. La-containing SBA-15/H2 O2 systems for the microwave assisted oxidation of a
lignin model phenolic monomer. Maderas Cienc. Tecnol. 2010, 12, 181–188. [CrossRef]
101. Gu, X.; He, M.; Shi, Y.; Li, Z. Production of armotic aldehyde by microwave catalytic oxidation of a lignin
model compound with La-containing SBA-15/H2 O2 systems. BioResources 2010, 5, 2029–2039.
102. Lanfranchi, M.; Prati, L.; Rossi, M.; Tiripicchio, A. The oxidation of ethane-1,2-diol resulting from molecular
oxygen activation by copper: Formation of an oxoethanoate complex precedes carbon-carbon bond cleavage.
J. Chem. Soc. Chem. Commun. 1993, 22, 1698–1699. [CrossRef]
103. Prati, L.; Rossi, M. Stepwise oxidation of 1,2-diols resulting from molecular oxygen activation by copper.
J. Mol. Catal. A 1996, 110, 221–226. [CrossRef]
104. Hanson, S.K.; Baker, R.T. Knocking on wood: Base metal complexes as catalysts for selective oxidation of
lignin models and extracts. Acc. Chem. Res. 2015, 48, 2037–2048. [CrossRef] [PubMed]
Polymers 2017, 9, 240 25 of 25

105. Elkurtehi, A.I.; Walsh, A.G.; Dawe, L.N.; Kerton, F.M. Vanadium aminophenolate complexes and their
catalytic activity in aerobic and H2 O2 -mediated oxidation reactions. Eur. J. Inorg. Chem. 2016, 2016,
3123–3130. [CrossRef]
106. Ede, R.M.; Ralph, J.; Torr, K.M.; Dawson, B.S.W. A 2D NMR investigation of the heterogeneity of distribution
of diarylpropane structures in extracted pinus radiata lignins. Holzforschung 1996, 50, 161–164.
107. Hanson, S.K.; Baker, R.T.; Gordon, J.C.; Scott, B.L.; Silks, L.A.P.; Thorn, D.L. Mechanism of alcohol oxidation
by dipicolinate vanadium(V): Unexpected role of pyridine. J. Am. Chem. Soc. 2010, 132, 17804–17816.
[CrossRef] [PubMed]
108. Haibach, M.C.; Lease, N.; Goldman, A.S. Catalytic cleavage of ether C–O bonds by pincer iridium complexes.
Angew. Chem. Int. Ed. 2014, 53, 10160–10163. [CrossRef] [PubMed]
109. Díaz-Urrutia, C.; Sedai, B.; Leckett, K.C.; Baker, R.T.; Hanson, S.K. Aerobic oxidation of 2-phenoxyethanol
lignin model compounds using vanadium and copper catalysts. ACS Sustain. Chem. Eng. 2016, 4, 6244–6251.
[CrossRef]
110. Ma, Y.; Du, Z.; Xia, F.; Ma, J.; Gao, J.; Xu, J. Mechanistic studies on the VO(acac)2-catalyzed oxidative cleavage
of lignin model compounds in acetic acid. RSC Adv. 2016, 6, 110229–110234. [CrossRef]
111. Blandez, J.F.; Navalón, S.; Alvaro, M.; Garcia, H. Graphenes as metal-free catalysts for the oxidative
depolymerization of lignin models. ChemCatChem 2015, 7, 3020–3026. [CrossRef]
112. Son, S.; Toste, F.D. Non-oxidative vanadium-catalyzed C–O bond cleavage: Application to degradation of
lignin model compounds. Angew. Chem. Int. Ed. 2010, 49, 3791–3794. [CrossRef] [PubMed]
113. Vom Stein, T.; Hartog, T.; Buendia, J.; Stoychev, S.; Mottweiler, J.; Bolm, C.; Leitner, W. Ruthenium-catalyzed
C–C bond cleavage in lignin model substrates. Angew. Chem. Int. Ed. 2015, 54, 5859–5863. [CrossRef]
[PubMed]
114. Sung, K.; Huh, S.; Jun, M. Syntheses of ruthenium(II) complexes containing polyphosphine ligands and their
applications in the homogeneous hydrogenation. Polyhedron 1999, 18, 469–479. [CrossRef]
115. Hanson, S.K.; Wu, R.; Silks, L.A. C–C or C–O bond cleavage in a phenolic lignin model compound: Selectivity
depends on vanadium catalyst. Angew. Chem. Int. Ed. 2012, 51, 3410–3413. [CrossRef] [PubMed]
116. Jiang, Y.-Y.; Yan, L.; Yu, H.-Z.; Zhang, Q.; Fu, Y. Mechanism of vanadium-catalyzed selective C–O and C–C
cleavage of lignin model compound. ACS Catal. 2016, 6, 4399–4410. [CrossRef]
117. Parker, H.J.; Chuck, C.J.; Woodman, T.; Jones, M.D. Degradation of β-O-4 model lignin species by vanadium
Schiff-base catalysts: Influence of catalyst structure and reaction conditions on activity and selectivity.
Catal. Today 2016, 269, 40–47. [CrossRef]
118. Wang, M.; Lu, J.; Zhang, X.; Li, L.; Li, H.; Luo, N.; Wang, F. Two-step, catalytic C–C bond oxidative cleavage
process converts lignin models and extracts to aromatic acids. ACS Catal. 2016, 6, 6086–6090. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like