Biochar Paper Research

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Science of the Total Environment 675 (2019) 203–212

Contents lists available at ScienceDirect

Science of the Total Environment

journal homepage: www.elsevier.com/locate/scitotenv

Catalytic oxidation of biorefinery corncob lignin via zirconium(IV)


chloride and sodium hydroxide in acetonitrile/water: A
functionality study
Fei Lin a,1, Chao Liu a,1, Xing Wang b, Changsong Hu a, Shiliang Wu a, Rui Xiao a,⁎
a
Key Laboratory of Energy Thermal Conversion and Control of Ministry of Education, School of Energy and Environment, Southeast University, Nanjing 210096, PR China
b
Liaoning Key Laboratory of Pulp and Paper Engineering, School of Light Industry and Chemical Engineering, Dalian Polytechnic University, Dalian 116034, PR China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Oxidation was carried out for the effi-


cient utilization of biorefinery corncob
lignin.
• ZrCl4 and NaOH were employed as the
co-catalyst for lignin oxidation in
CH3CN/H2O.
• Oxidative cleavage on inter-unit link-
ages not on aromatic rings was recog-
nized by NMR.

a r t i c l e i n f o a b s t r a c t

Article history: In order to realize the efficient utilization of biorefinery corncob lignin, the promising catalytic oxidation strategy
Received 18 January 2019 was carried out by using ZrCl4 and NaOH as the co-catalyst and dioxygen as the oxidant in MeCN/H2O. GC/MS,
Received in revised form 14 April 2019 GC-FID, and MALDI-TOF/MS were employed to recognize the produced monomers and oligomers, and GPC
Accepted 15 April 2019
was used to monitor the molecular weight changes of lignin fragments. In addition, specific structural evolution
Available online 16 April 2019
of corncob lignin during ZrCl4/NaOH-catalyzed oxidation were revealed by quantitative 13C (Q13C) and 2D HSQC
Editor: Daniel CW Tsang NMR techniques. Results showed that the total yields of produced oxidation monomers reached 6.8 wt%, and ar-
omatic aldehydes were the major species, in which vanillin and 4-hydroxybenzaldehyde were the two dominant
Keywords: products. After ZrCl4/NaOH-catalyzed oxidation, the weight-average molecular weight of corncob lignin and its
Lignin products decreased from 2000 Da to 300 Da after oxidation with 16 h. Moreover, Q13C NMR analysis showed
Oxidation the decrease percentage of C\\O aliphatic carbons (including methoxyl carbons), the increase percentage of
Zirconium C\\C aliphatic and carbonyl carbons, and the relative stable percentage of aromatic carbons with reaction
GPC prolonged. These results combined with the further confirmation from HSQC indicated the oxidative cleavage
NMR
of C\\O aliphatic linkages and removal of methoxy groups within corncob lignin, as well as the formation of
MALDI-TOF/MS
C\\C aliphatic bonds and carbonyl groups. The work presented a comprehensive insight into the catalytic oxida-
tive depolymerization of biorefinery corncob lignin.
© 2019 Elsevier B.V. All rights reserved.

⁎ Corresponding author.
E-mail address: ruixiao@seu.edu.cn (R. Xiao).
1
Authors contributed equally to the production of this paper.

https://doi.org/10.1016/j.scitotenv.2019.04.224
0048-9697/© 2019 Elsevier B.V. All rights reserved.
204 F. Lin et al. / Science of the Total Environment 675 (2019) 203–212

1. Introduction on various supporters are carefully screeded, and the corresponding


mechanisms on lignin oxidation have been thoroughly explored (Liu
Corn is one of the three major crops (wheat, paddy, and corn) in the et al., 2019). On the other hand, zirconium-based catalysts have been re-
world. Meanwhile, as the residue of corn processing, the production of ported for lignin conversion under subcritical conditions (Nandiwale
corncob is huge (Yang et al., 2012). In China, the annual output of corn- et al., 2017; Barakat et al., 2018), but their effects on lignin oxidation
cob is approximately 50 million tons (Liang et al., 2018). Therefore, are rarely reported, except for the cobalt/manganese/zirconium/bro-
achieving the value-added utilization of agricultural residual corncob mide (Co/Mn/Zr/Br) catalyst in acetic acid-water solvent developed by
has great potential economic benefits (Vu et al., 2017). Owing to the Partenheimer (2009) Therefore, we intended to explore whether
high cellulose content, corncob can be used as a feedstock for cellulosic zirconium-based catalysts can still work well on lignin oxidation.
ethanol production (Fan et al., 2013). During the bio-ethanol produc- In this work, the oxidative depolymerization of biorefinery corncob
tion, the lignin fraction is removed with the processes of sulfite or alkali lignin catalyzed by zirconium chloride (ZrCl4) together with sodium hy-
pretreatment and washing procedure (Fan et al., 2013). Furthermore, droxide (NaOH) in acetonitrile/water co-solvent using dioxygen was in-
with the development of second generation bio-ethanol, a large number vestigated. The produced monomers and oligomers were separately
of lignin-rich solid residues is produced (Behling et al., 2016; Doherty identified by GC/MS, GC-FID, and MALDI-TOF/MS. In addition, the
et al., 2011). As well known, lignin is an aromatic polymer, which con- changes of average molecular weight and chemical structure of corncob
tributes it as the only renewable aromatic resource in nature (Lange lignin during ZrCl4/NaOH-catalyzed oxidative depolymerization were
et al., 2013; Li et al., 2015; Cao et al., 2018). It can be used as an alterna- analyzed by GPC, quantitative 13C, and 2D HSQC NMR. This paper pre-
tive fossil resources to produce valuable aromatic platform chemicals, sented a comprehensive insight into product formation and structure
liquid fuels, and biomass-based materials. Therefore, in order to make evolution during lignin oxidative depolymerization.
the deep and efficient usage of corncob residues, it is necessary to con-
vert the biorefinery residual lignin into value-added products. 2. Experimental
Pyrolysis (thermolysis) (Lei et al., 2019; Huang et al., 2018;
Patwardhan et al., 2011), reduction (hydrogenolysis) (Hu et al., 2019; 2.1. Material and chemicals
Zhang et al., 2014), and chemical oxidation (Liu et al., 2019; Kalliola
et al., 2011; R. Ma et al. 2015) are the three major thermochemical Feedstock used in this study was corncob lignin, which was supplied
methods regarding lignin depolymerization. Pyrolysis can rapidly trans- by Shandong Longlive Bio-Technology Co. Ltd. (Dezhou, China). The lig-
fer lignin into bio-crude oil. However, the components of obtained bio- nin was a byproduct from xylo-oligosaccharides and bio-ethanol pro-
oil are complicated, which leads to its poor physical and chemical prop- duction obtained with the following processes. First, the corncob was
erties (Mu et al., 2013). Catalytic reduction of lignin is an alternative hydrothermally treated to degrade hemicelluloses into xylo-
way to prepare phenolic compounds, but many of the employed cata- oligosaccharides. The residue was further extracted with an alkaline so-
lysts are noble metal catalysts. Meanwhile, the uncontrolled hydrogena- lution to dissolve lignin and obtain a cellulose-rich residue which was
tion under harsh conditions would saturate the aromatic rings of lignin used as the feedstock for bio-ethanol production. Finally, the alkaline
and produced monomers (Lee et al., 2012). Compared with pyrolysis lignin solution was adjusted to acidic condition to precipitate the lignin
and reduction, catalytic oxidation well reserves the aromatic units, (Yang et al., 2014). Composition, element analysis, and proximate anal-
and is one of the viable methods for lignin valorization with environ- ysis of the biorefinery corncob lignin were summarized in Table 1. Zirco-
mentally friendly oxidants such as oxygen, air, or H2O2 (Das et al., nium(IV) chloride (ZrCl4, 98%, anhydrous) purchased from J&K
2017). Lignin oxidation presents various attractive advantages such as scientific Ltd. (Beijing, China) and sodium hydroxide (NaOH, analytical
product accumulation, low energy consumption, mild conditions and reagent) brought from Sinopharm Chemical Reagent Co. Ltd. (Shanghai,
high conversion, which has been considered to be the most promising China) were employed as the catalysts. n-Dodecane (standard for GC,
chemical depolymerization method in lignin conversion and utilization ≥99.5%) shopped from Aladdin Industrial Corporation (Shanghai,
(Gao et al., 2018). China) was used as the internal standard for the calibrations of pro-
Oxidative cracking reactions during lignin depolymerization occur duced monomers. Acetonitrile (MeCN), ethyl acetate, acetic anhydride,
on aryl ether bonds, carbon-carbon bonds, aromatic rings, or other link- pyridine, tetrahydrofuran (THF), and deuterated dimethyl sulfoxide
ages within lignin macromolecules (Li et al., 2015). Lignin oxidation (DMSO-d6) were all commercial and used as received. Deionized
produces a suite of platform compounds including phenolic aldehydes, water was produced in our lab. High purity O2 (99.999%) was provided
ketones, and acids, benzoquinones, and aliphatic carboxylic acids, and by Nanjing Shangyuan Gas Company (Nanjing, China).
these compounds can be further used for the production of species,
drugs, chemicals, and medical products (Liu et al., 2019). However, 2.2. General procedure of lignin oxidation and product processing
most of the current oxidation researches are carried out on lignin
model compounds (Deng et al., 2015; Mottweiler et al., 2015; Sedai Lignin powder (0.4 g) was charged into a 25 mL batch reactor (Bei-
et al., 2011, 2013; Zhou et al., 2017). For example, Hanson et al. jing Century Sen Long Instruments Company, Beijing, China), together
(2012) employed lignin model compound syringyl glycerin guaiacyl with the ZrCl4 catalyst (15 wt%), the NaOH catalyst (8 wt%), and a mag-
ether as the substrate to investigate the selective oxidative C\\O or netic stirrer. Subsequently, MeCN/H2O co-solvent (15 mL, 1:1, v/v) was
C\\C bond cleavage catalyzed by oxovanadium complexes. What is added. The reactor was purged three times with O2 and charged with
worse, these developed oxidative catalytic systems are not applied in 1.5 MPa of O2, and the reaction mixture was stirred at 140 °C with
technical lignin oxidation, except for those conventional catalytic wet 1000 rpm for the desired reaction time. After reaction, the reactor was
air oxidation of lignin (Sales et al., 2004). These deficiencies require us taken out from the heater and quenched instantly with ice water. The
to employ technical lignin directly as the feedstock to explore its oxida- remaining oxygen pressure was released, and the reactor was opened.
tive depolymerization. Therefore, we chose biorefinery corncob lignin Then, the reaction mixture was acidized to a pH of 2, and filtered with
as the substrate to do the catalytic oxidation. filter paper to remove residues and precipitates, which were further
In addition to the metal salt and oxide catalysts (Liu et al., 2019), dried for the calculation of the conversion rate. Before analyzing, the re-
vanadium- (Y. Ma et al. 2015; Son and Toste, 2010), copper- (Salonen action mixture was filtered again using a 0.45 μm PTFE syringe filter.
et al., 2018; Wang et al., 2016), and cobalt-based catalysts (Biannic One milliliter of the filtered solution was taken and diluted with THF
and Bozell, 2013; Rinesch and Bolm, 2018) are commonly employed (to make the concentration b1 mg/mL) to characterize the produced
catalysts in lignin oxidation. The catalytic activity of these vanadium-, oligomers using matrix assisted laser desorption ionization-time of
copper-, and cobalt-based catalysts with different ligands and loaded flight mass spectrometry (MALDI-TOF/MS). Half of the remaining
F. Lin et al. / Science of the Total Environment 675 (2019) 203–212 205

Table 1
Composition, element analysis, and proximate analysis of biorefinery corncob lignin (dry basis).

Sample Composition (wt%) Element analysis (wt%) Proximate analysis (wt%)

Lignin Carbohydrate C H Oa N S Moisture Volatiles Fixed carbon Ash

Corncob lignin 98.67 0.63 62.33 5.44 30.38 0.87 0.28 Trace 67.28 32.02 0.70
a
Determined by difference.

filtered solution was further rotary evaporated under reduced pressure yield and selectivity of monomers were further obtained with the fol-
to remove co-solvent, and then extracted with ethyl acetate. This ex- lowing formulas.
traction was added with n-dodecane, diluted to the constant volume,  
and analyzed by gas chromatography/mass spectrometry (GC/MS) mass of the residue ðgÞ
Conversion ð%Þ ¼ 1−  100%
and GC-flame ionization detector (FID) to identify and quantify the ox- mass of initial corncob lignin ðgÞ
idation monomers. Other filtered solution was freeze-dried by adding  
large amounts of water (to dilute MeCN). The obtained solid product mass of the monomer ðgÞ
Yield ð%Þ ¼  100%
was divided into two parts. One was used for molecular weight mea- mass of initial corncob lignin ðgÞ
surement with the help of gel permeation chromatography (GPC)  
after acetylation. The other was directly dissolved in DMSO-d6 for nu- mass of the monomer ðgÞ
Selectivity ð%Þ ¼  100%
clear magnetic resonance (NMR) testing to reveal the chemical struc- mass of total identified monomers ðgÞ
ture evolution during corncob lignin oxidation. The overall schematic
procedure for product processing was illustrated in Fig. 1. Oligomers from corncob lignin oxidation were characterized by
MALDI-TOF/MS spectroscopy, which was conducted in a Bruker
2.3. Characterization and data processing Ultraflextreme MALDI in negative mode with α-cyano-4-
hydroxycinnamic acid (HCCA) as the matrix. The irradiation source
Monomers from corncob lignin oxidation were first qualitatively an- was a N2 laser with a wavelength of 337 nm, and the duration of one
alyzed by GC/MS using an Agilent 7890B gas chromatograph and an laser pulse was 3 ns. The spectra were acquired in the mass range m/z
Agilent 5977B mass spectrometer equipped with an HP-5MS capillary 0 to 1400 with delayed extraction of 200 ns and acceleration voltage
column (30 m length, 0.25 mm inner diameter, and 0.25 μm film thick- was 20 kV.
ness). The column flow rate (high purity helium, 99.999%) was Molecular weight distribution of corncob lignin and its oxidation
1 mL/min, and the injection temperature was 290 °C. 1 μL of sample products was tested by an Agilent 1100 gel permeation chromatograph
was injected with an autosampler in split mode (split ratio 5:1). The col- equipped with two cascaded PL-gel columns (5 μm, 500 Å) and a refrac-
umn temperature was programmed from 50 °C to 280 °C at a heating tive index detector. First, corncob lignin and its oxidation fragments
rate of 6 °C/min, with an initial time delay of 2 min and holding at 280 were acetylated using pyridine/acetic anhydride (1:1, v/v) to improve
°C for 10 min. The mass spectrometer was set at an ionizing voltage of their solubility in THF (An et al., 2015). Thereafter, THF (HPLC grade)
70 eV, and the mass range was scanned from m/z 33 to 400. The mono- was used as the mobile phase with a flow velocity of 1.0 mL·min−1,
mers were identified by comparing the MS spectra with standard and a series of polystyrene samples with a molecular mass range be-
compounds. tween 580 Da and 19,900 Da served as standards.
Quantitative analysis of produced monomers from corncob lignin NMR determinations for corncob lignin and its oxidation products
oxidation was carried out on a Shimadzu GC-2014 gas chromatograph were performed in a Bruker AVIII 400 MHz spectrometer. All samples
coupled with an HP-5MS capillary column and a FID detector. The tem- were dissolved in DMSO-d6 with a 25 wt% concentration. For quantita-
perature program of GC oven started from 50 °C holding for 2 min to 280 tive 13C (Q13C) NMR analysis, the operation was performed in the FT
°C holding for 30 min with a heating rate of 6 °C/min. The split ratio was mode at 100.6 MHz with an inverse-gated decoupling pulse sequence
set at 5:1. Mass of produced monomers was calculated based on the ef- to avoid NOE effects. Q13C NMR spectra were acquired using an
fective carbon number (ECN) method as described in the literature (Lan inverse-gated decoupling (Waltz-16) pulse sequence with a 30° pulse
et al., 2018), in which the ECNs of monomers were corrected with the angle, 2 s pulse delay, 1.4 s acquisition time and 30,000 scans. 2D
standard n-dodecane in ethyl acetate. Conversion rate of corncob lignin, HSQC NMR determination was carried out according to the standard

Fig. 1. Schematic procedure for product processing.


206 F. Lin et al. / Science of the Total Environment 675 (2019) 203–212

100 10 80 10
Conversion (a) Conversion (b)
Unknown monermers O2 N2 Unknown monermers
80 Identified monermers 8 Identified monermers 8
60
Conversion (%)

Conversion (%)
Yield (wt%)

Yield (wt%)
60 6 6
40
40 4 4

20
20 2 2

0 0 0 0
No Catalyst ZrCl4 NaOH ZrCl4/NaOH ZrCl4/NaOH 1 2 4 8 16
Catalyst Reaction time (h)

Fig. 2. Conversion rate of corncob lignin and corresponding yields of monomers under various conditions: (a) effect of catalyst with the reaction time of 2 h; (b) effect of reaction time with
ZrCl4/NaOH co-catalyst. Other conditions: 0.4 g of corncob lignin, 15 wt% of ZrCl4, 8 wt% of NaOH, 15 mL of MeCN/H2O (1:1, v/v), 1.5 MPa of O2, and 140 °C.

pulse program “hsqcedetgpsisp2”. The spectral widths of 20,000 Hz and H2O, resulting in the 100% of conversion. With the combined effect of
5000 Hz were used for the 13C and 1H dimensions, respectively. ZrCl4 and NaOH, 3.7 wt% of identified monomers and 3.1 wt% of un-
known monomers were generated, together with the conversion rate
3. Results and discussion of 42.3%. In addition, the role of O2 atmosphere was presented by
using 1.5 MPa of N2 atmosphere as a comparison (Fig. 2a). Under N2 at-
3.1. Catalytic oxidation of corncob lignin by ZrCl4 and NaOH in MeCN/H2O mosphere, only 0.57 wt% of identified monomers and 0.89 wt% of un-
known monomers were produced from ZrCl4/NaOH-catalyzed corncob
To confirm the co-catalytic effect of ZrCl4 and NaOH on corncob lig- lignin conversion, which illustrated the oxidation conversion of
nin oxidation, catalytic oxidation of corncob lignin with ZrCl4, NaOH, biorefinery corncob lignin with ZrCl4 and NaOH as the co-catalyst in
ZrCl4/NaOH, and without catalyst were separately carried out in MeCN/H2O system.
MeCN/H2O system with O2 as the oxidant. The conversion rate of corn- The effect of reaction times on corncob lignin oxidation and product
cob lignin and corresponding yields of generated monomers were yields was also investigated, and the results were depicted in Fig. 2b.
shown in Fig. 2a. As can be seen in the bar chart, 10.9% of corncob lignin When the reaction time prolonged from 1 h to 16 h, the conversion
was converted and only 0.3 wt% of unknown monomers was produced rate of corncob lignin kept increasing from 40.6% to 69.7%. However,
without any catalysts. When using ZrCl4 as the catalyst, the conversion the yields of identified monomers fluctuated little before 4 h. They
rate increased to 19.9%, and the monomer yields reached 5.9 wt% in- were 3.6 wt% for 1 h, 3.7 wt% for 2 h, and 3.6 wt% for 4 h. Then the yields
cluding 2.5 wt% of identified monomers and 3.4 wt% of unknown mono- of identified monomers dropped to 2.0 wt% for 8 h and 1.5 wt% for 16 h.
mers. NaOH can also promote the formation of 5.8 wt% of monomers, in Opposite to the formation of identified monomers in a short time, the
which 2.4 wt% for identified monomers and 3.4 wt% for unknown unknown monomers were generated after a long time. Their yields
monomers. The conversion rate of corncob lignin catalyzed by NaOH went from 1.4 wt% (1 h) to 3.4 wt% (4 h), and then kept at this level
was up to 100%. That is because NaOH dissolved in MeCN/H2O can for longer reaction times. Comparing the conversion rate and corre-
transform the corncob lignin into fragments with phenoxy anions sponding yields of monomers, we should notice that large amounts of
which can induce the cleavage of inter-unit linkages (Sun et al., 2018). oligomers were produced during ZrCl4/NaOH-catalyzed corncob lignin
These processes depolymerized corncob lignin into low-molecular- oxidation. These oligomers would be further characterized by MALDI-
weight fragments and increased their dissolving capacity in MeCN/ TOF/MS, GPC, and NMR in the following sections.

7
60
n-Dodecane (a) (b)
13 14 50
9 11
1h 2 3 4 6

40
Selectivity (%)

12
2h 1 5 8 10 15
Intensity

30

4h
20

8h
H compounds
10 G compounds
S compounds
16 h
0
5 10 15 20 25 30 0 2 4 6 8 10 12 14 16
Retention time (min) Time (h)

Fig. 3. GC/MS spectra (a) and total selectivity of H, G, and S products from lignin oxidative depolymerization (b) in ZrCl4/NaOH catalyzed system under O2 in MeCN/H2O.
F. Lin et al. / Science of the Total Environment 675 (2019) 203–212 207

Table 2
Names and structures of depolymerized monomeric products from ZrCl4/NaOH-catalyzed corncob lignin oxidation detected by GC/MS and GC-FID.

Label Name Structure Label Name Structure Label Name Structure

1 Phenol 2 Guaiacol 5 Syringol

3 4-Vinylphenol 4 4-Vinylguaiacol 11 2,6-Dimethoxy-1,4-benzoquinone

6 4-Hydroxybenzaldehyde 7 Vanillin 13 Syringaldehyde

8 4-Hydroxyacetophenone 9 Acetovanillone 14 Acetosyringone

10 4-hydroxybenzoic acid 12 Vanillic acid 15 Syringic acid

3.2. Monomers from ZrCl4/NaOH-catalyzed oxidation of corncob lignin aldehydes, aromatic ketones, aromatic acids, benzoquinones, and phe-
nolics. Aromatic aldehydes, ketones, and acids were formed from the
Monomers produced from ZrCl4/NaOH-catalyzed oxidation of corn- oxidative cleavage of inter-unit linkages within corncob lignin macro-
cob lignin in MeCN/H2O co-solvent were identified with GC/MS and molecules (Tarabanko and Tarabanko, 2017); whereas the generation
quantified with GC-FID. The GC/MS spectra were displayed in Fig. 3a, of 2,6-dimethoxy-1,4-benzoquinone contributed to the catalytic oxida-
and the corresponding names and structures of identified species tion of aromatic rings (Stark et al., 2010). In addition, the production of
were listed in Table 2. These products mainly included aromatic phenol, guaiacol, and their 4-vinyl substituents should benefit from the

80 40
(a) (b)
70 35

60 30
Aromatic aldehydes Phenol
Selectivity(%)

Selectivity(%)

50 25
Aromatic ketones 4-Vinylphenol
Aromatic acids 4-Hydroxybenzaldehyde
40 20
Benzoquinones 4-Hydroxyacetophenone
Phenolics 4-Hydroxybenzoic acid
30 15

20 10

10 5

0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
Time (h) Time (h)
40 14
(c) (d) Syringol
35 2,6-Dimethoxy-1,4-benzoquinone
12
Syringaldehyde
30 Acetosyringone
10 Syringic acid
Selectivity(%)

Selectivity(%)

25 Guaiacol
4-Vinylguaiacol 8
20 Vanillin
Acetovanillone 6
15 Vanillic acid
4
10

5 2

0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14 16
Time (h) Time (h)

Fig. 4. Selectivity of depolymerization products from ZrCl4/NaOH-catalyzed corncob lignin oxidation with various times: (a) total products; (b) H products; (c) G products, and (d) S
products.
208 F. Lin et al. / Science of the Total Environment 675 (2019) 203–212

presence of ferulate and p-coumarate within corncob lignin (which


would be further revealed by 2D HSQC NMR).
As a herbaceous lignocellulosic material, corncob lignin is the typical
GSH lignin (Ouyang et al., 2018). Therefore, it is necessary to divide
Lignin
these compounds into p-hydroxyphenyl (H), guaiacyl (G), and syringyl
(S) groups to investigate the demethoxylation reactions during corncob
lignin oxidation. The groups of identified products were presented in
Table 2, and the corresponding total selectivity of H, G, and S com-
pounds with time were shown in Fig. 3b. The general trends were that 1h
the total selectivity of S and G compounds kept decreasing, whereas
the total selectivity of H compounds increased continuously. As we pre-
viously reported in lignin pyrolysis, S units could degrade into G units,
but the formation of H phenols basically attributed from the
demethoxylation of initial G units (Liu et al., 2016). Therefore, the de- 2h

Intensity
crease of the total selectivity of S and G compounds together with the
increase of the total selectivity of H compounds did account for the
demethoxylation reactions of corncob lignin in ZrCl4/NaOH-catalyzed
oxidation system, and the demethoxylation rate of G units was faster
than that of S units. 4h
Selectivity of specific compounds with time was illustrated in Fig. 4,
including the selectivity of total products in various groups (Fig. 4a), the
selectivity of H products (Fig. 4b), the selectivity of G products (Fig. 4c),
and the selectivity of S products (Fig. 4d). As displayed in Fig. 4a, the
8h
major products produced from ZrCl4/NaOH-catalyzed corncob lignin
oxidation were aromatic aldehydes. Their total selectivity was stable
at about 67%. The second major products were aromatic ketones,
whose total selectivity rose from 10.7% (1 h) to 20.0%. (16 h). Aromatic
acids, benzoquinones, and phenolics were regarded as the minor prod- 16 h
ucts, whose selectivity were all b10%. The high selectivity of aromatics
(including aldehydes, ketones, acids, phenolics) and the low selectivity 0 200 400 600 800 1000 1200 1400
of benzoquinones indicated that the ZrCl4/NaOH-catalyzed oxidation m/z
mainly occurred on the inter-unit linkages within corncob lignin macro-
molecules not on the aromatic rings, which was consistent with re- Fig. 5. MALDI-TOF/MS spectra of corncob lignin and its products after ZrCl4/NaOH-
catalyzed oxidation with various times.
ported advanced lignin oxidation (Ma et al., 2018) and would be
further confirmed by NMR techniques in the following sections.
Fig. 4b presented that the H compounds were the main products fragments mainly consisted of dimers and trimers, illustrating the
from ZrCl4/NaOH-catalyzed corncob lignin oxidation, in which 4- good oxidative depolymerization of corncob lignin. On the other hand,
hydroxybenzaldehyde was the dominant product. Its selectivity in- two cluster peaks around 700 m/z and 800 m/z were observed in
creased from 19.8% (1 h) to 36.0% (16 h). The selectivity trends of ZrCl4/NaOH-catalyzed oxidation fragments. These peaks suggested the
other H products were remarkable resemblance with those of total existence of oligomers produced from ZrCl4/NaOH-catalyzed oxidation
products divided into groups as displayed in Fig. 4a. As for the G prod- of corncob lignin. Furthermore, the maximum intensity of these two
ucts (Fig. 4c), vanillin was the major product. However, in contrast to cluster peaks was found in the four-hour oxidation fragments, indicat-
the increasing selectivity of 4-hydroxybenzaldehyde, the selectivity of ing the highest concentration of produced oligomers. With reaction
vanillin decreased from 42.5% to 19.6% with the reaction carried out time prolonged to 8 h and 16 h, the intensity of these two cluster
form 1 h to 16 h. Furthermore, except for the relatively stable selectivity peaks decreased. This meant the long reaction time would further
of acetovanillone at about 5%, the selectivity of guaiacol, 4-vinylguaiacol,
1.6
and vanillic acid were low. Similar with the G products, the selectivity of 200 Da Lignin
syringaldehyde and 2,6-dimethoxy-1,4-benzoquinone continue to 1h
1.4
decrease with time, which were the two main products among S com- 2h
pounds (see Fig. 4d). 1.2 4h
8h
3.3. Oligomers from ZrCl4/NaOH-catalyzed oxidation of corncob lignin 1.0 16 h
W(log M)

Except for the oxidation monomeric products, there were still large 0.8
amounts of dimers and oligomers produced from ZrCl4/NaOH-
catalyzed oxidation of corncob lignin, which cannot be identified by 0.6
GC/MS and GC-FID. In order to measure these fragments, MALDI-TOF/
0.4
MS was employed to scan the corncob lignin and its oxidation products
dissolved in THF (HPLC grade), and the corresponding spectra from 0 to
0.2
1400 m/z were presented in Fig. 5. As can be seen, the species and inten-
sity of peaks in corncob lignin solution were lower than those in oxida- 0.0
tion fragments. This might be explained by two reasons. One was the 102 103 104 105
poor solubility of corncob lignin in THF, and the other reason was the Molar mass (Da)
absence of low-molecular-weight components in corncob lignin.
After ZrCl4/NaOH-catalyzed oxidation, abundant fragments b400 m/ Fig. 6. GPC curves of corncob lignin and its products after ZrCl4/NaOH-catalyzed oxidation
z were produced. Judging from the range of molecular weight, these with various times.
F. Lin et al. / Science of the Total Environment 675 (2019) 203–212 209

Table 3 Furthermore, the GPC curves can also present the degree of poly-
Average molecular weights and polydispersity indices of corncob lignin and its products merization of peak fragments roughly. As McClelland et al. (2017) re-
after ZrCl4/NaOH-catalyzed oxidation with various times.
ported, the average molecular weight of a polymeric unit can be
Reaction Weight-average Number-average Polydispersity calculated according to the S:G:H ratio of lignin fragment. The S:G:H
time molecular molecular weight index ratio of corncob lignin was 0.20:0.72:0.08 (as revealed in HSQC in
weight (Mw) (Mn)
Section 3.5.2), and the molecular weights of the p-hydroxyphenyl (1-
Corncob 2000 600 3.3 (4-hydroxyphenyl)-1,3-propanediol), guaiacyl (1-(4-hydroxy-3-
lignin
methoxyphenyl)-1,3-propanediol) and syringyl (1-(4-hydroxy-3,5-
1h 930 360 2.6
2h 640 320 2.0 dimethoxyphenyl)-1,3-propanediol) polymeric units are 168.22 Da,
4h 470 260 1.8 198.22 Da, and 228.24 Da, respectively. Therefore, the average molecu-
8h 430 270 1.6 lar weight of a polymeric unit in corncob lignin is 201.82 Da, and we can
16 h 300 210 1.4 approximately regard fragments below 200 Da as monomers, fragments
between 200 Da and 400 Da as dimers, and fragments above 400 Da as
oligomers. Based on this distinction, all of the depolymerized corncob
destroy some oligomer substructures into low-molecular-weight spe- lignin fragments were mainly consisted of oligomers and dimers. With
cies. In general, the MALDI-TOF/MS confirmed the dimers and oligo- prolonged oxidation time, the degree of polymerization of corncob lig-
mers from ZrCl4/NaOH-catalyzed oxidative depolymerization of nin fragments decreased, suggesting the continuous oxidative depoly-
corncob lignin macromolecules. Furthermore, the changes of average merization process catalyzed by ZrCl4 and NaOH.
molecular weight of corncob lignin and its oxidation products would The values of weight-average (Mw), number-average (Mn) molecu-
be detected with GPC in subsequent section. lar weights, and polydispersity indexes (Mw/Mn) of corncob lignin and
its oxidation fragments were illustrated in Table 3. The Mw of corncob
3.4. Molecular weight changes of corncob lignin during ZrCl4/NaOH- lignin was 2000 Da, whereas after oxidative depolymerization with
catalyzed oxidation 1 h, the lignin molecular chain was approximately half cleaved, with
the Mw dropping to 930 Da. According to the obtained average molec-
The GPC curves of corncob lignin and its oxidation products cata- ular weight of a polymeric unit in corncob lignin (201.82 Da), the aver-
lyzed by ZrCl4 and NaOH with various hours were shown in Fig. 6. As age degree of polymerization of corncob lignin decreased from 9.9 to
can be seen, the corncob lignin had a very broad range of molecular 4.6. With the reaction time prolonged, the Mw of lignin fragments
weight distribution from ca. 100 Da to 40,000 Da. Furthermore, the went on declining to 640 Da (2 h) and 470 Da (4 h). The corresponding
corncob lignin had two primary peaks located at 369 Da and 1185 Da, average degrees of polymerization were 3.2 and 2.3, respectively. These
respectively. After ZrCl4/NaOH-catalyzed oxidation, the molecular implied that after four-hour oxidation, the depolymerization fragments
weights of lignin fragments were shortened, and the corresponding dis- were mainly made up of dimers (as well as small amount of trimers).
tribution ranges were narrowed. For the one-hour product, its molecu- Since the main components in ZrCl4/NaOH-catalyzed oxidation system
lar weight ranged from ca. 100 Da to 15,000 Da, and its curve contained with 4 h were already dimers, the reduction of molecular weight
three peaks at 187 Da, 288 Da, and 588 Da, respectively. With the reac- would not be so obvious as the reaction proceeded. As can been seen
tion progressed to 2 h and 4 h, the molecular weight distribution ranges in Table 3, the Mw of lignin fragments after 8 h and 16 h were separately
decreased to ca. 100–6000 Da and 100–4000 Da. The primary peaks of 430 Da and 300 Da, and the corresponding degrees of depolymerization
the two-hour and four-hour lignin fragments shifted to 274 Da and were 2.1 and 1.5, respectively. The overall analysis further confirmed
267 Da. Meanwhile, the obvious shoulder peaks in these two curves that the depolymerization samples were composed of dimers, which
were observed. They were 188 Da and 469 Da for 2 h, and 166 Da and was in consistent with the results from MALDI-TOF/MS. In addition,
461 Da for 4 h. When the reaction time reached 8 h and 16 h, the molec- after ZrCl4/NaOH-catalyzed oxidation, the polydispersity (Mw/Mn) of
ular weight distribution ranges further dropped to ca. 100–3000 Da and lignin fragments decreased from 3.3 (corncob lignin) to 2.6 (1 h), and
100–1800 Da. The corresponding primary peaks moved to 255 Da (8 h) then continually declined to 1.8 (4 h) and 1.4 (16 h). This trend sug-
and 232 Da (16 h). In addition, the curve of sixteen-hour oxidation gested that these oxidation fragments had better uniformity with the
product had a significant shoulder peak at 159 Da. increase of oxidation time.

Carbonyl Aromatic C-O aliphatic C-C aliphatic 60


(a) Region Region Region (b) Carbonyls Aromatics Methoxyls
C-O aliphatics C-C aliphatics
DMSO 50
OMe

Lignin
40
Percentge (%)

1h
30
2h

4h 20

8h 10

16h 0
0 2 4 6 8 10 12 14 16
200 150 100 50 0
Chemical shift (ppm) Reaction time (h)

Fig. 7. Quantitative 13C NMR spectra (a) and results (b) of corncob lignin and its products after ZrCl4/NaOH-catalyzed oxidation with various times.
210 F. Lin et al. / Science of the Total Environment 675 (2019) 203–212

3.5. Chemical structure evolution of corncob lignin during ZrCl4/NaOH- 2014). As reflected in Fig. 7a, after ZrCl4/NaOH-catalyzed oxidation, a se-
catalyzed oxidation ries of new signals in C\\C aliphatic and carbonyl regions were ob-
served, meanwhile the intensity of peaks in C\\O aliphatic region
3.5.1. Quantitative 13C NMR analysis decreased, mainly the intensity of methoxy groups. These suggested
Quantitative 13C (Q13C) NMR was employed to analyze the oxidation the oxidative cleavage of inter-unit C\\O linkages, the formation of
products from corncob lignin catalyzed by ZrCl4 and NaOH to distin- C\\C bonds and carbonyl groups, as well as the removal of methoxyl
guish different carbons in various chemical positions. The spectra and groups. Furthermore, the decrease of peak intensity in aromatic region
corresponding integral results were shown in Fig. 7. The Q13C NMR was not significant, but the chemical positions of these peaks shifted, in-
spectra can be divided into four regions, including C\\C aliphatic region dicating the oxidative modification of aromatic units of depolymerized
(0–55 ppm), C\\O aliphatic region (55–95 ppm) (containing methoxyl lignin fragments.
region, 55–57 ppm), aromatic region (95–165 ppm), and carbonyl re- The detailed percentages of carbons in aromatic, carbonyl, C\\C ali-
gion (165–215 ppm) (Chen et al., 2016; Wen et al., 2013a; Yang et al., phatic, C\\O aliphatic, and methoxyl regions of lignin fragments can be

Fig. 8. 2D HSQC NMR spectra of corncob lignin and its products after ZrCl4/NaOH-catalyzed oxidation with various times.
F. Lin et al. / Science of the Total Environment 675 (2019) 203–212 211

found in Fig. 7b. As for the corncob lignin feedstock, it contained 52.7% of Moreover, the signal at δC/δH 87/5.5 ppm indicated the existence of
aromatic carbon, 18.2% of C\\O aliphatic carbon, 11.4% of C\\C aliphatic Cγ/Hγ in the phenylcoumaran substructure (substructure C, β-5′ link-
carbon, 9.0% of methoxyl carbon, and 8.7% of carbonyl carbon. During age). As displayed in Fig. 8, the intensity of OMe dropped continuously
ZrCl4/NaOH-catalyzed oxidation, the percentage of aromatic carbon de- with reaction time, which was consistent with the result from Q13C
creased a little and maintained at ca. 45%. The ca. 7% loss of aromatic car- NMR. Similarly, the intensity of β-O-4′ and β-5′ signals also decreased
bon, i.e., about 1% loss of the aromatic unit, suggested that the main significantly with the oxidation carried out. This indicated the oxidative
oxidation did not occur on the aromatic unit, which was in consistent cleavage of inter-unit bonds, especially β-O-4′ linkages, in corncob lig-
with the low yield of 2,6-dimethoxy-1,4-benzoquinone, as displayed nin, which also matched well with the percentage decline of C\\O ali-
in Fig. 4d. The percentages of C\\O aliphatic and methoxyl carbon de- phatic carbons.
creased rapidly from 18.2% vs. 9.0% (corncob lignin) to 9.1% vs. 4.4% In the aromatic region, 13C\\1H correlation signals of syringyl (S),
(4 h). After oxidation with 4 h, the decreasing rate of the percentages guaiacyl (G), hydroxyphenyl (H), ferulate (FA), and p-coumarate
of C\\O aliphatic and methoxyl carbon slowed down, achieving 6.5% (pCA) units could be found. Signals at δC/δH 103.8/6.71 ppm were signed
for the C\\O aliphatic carbon and 1.5% for the methoxyl carbon at to the C2,6/H2,6 correlations of S units. Signals at δC/δH 113.4/6.69 ppm
16 h. This phenomenon indicated that main parts of the oxidative cleav- and 115.5/6.79 ppm corresponded to the C2/H2 and C5,6/H5,6 correla-
age of inter-unit C\\O linkages took place before 4 h. Meanwhile, the tions in G units. Signals at δC/δH 127.7/7.05 ppm correlated to the C2,6/
percentage of C\\C aliphatic carbon rose steadily from 11.4% for the H2,6 from H units. In addition, signals of C2,6/H2,6 in S′ units (Cα-
corncob lignin feedstock to 19.0% for the lignin fragment oxidized oxidized S units) located at δC/δH 106.6/7.20 ppm. The C2/H2 and C6/
with 8 h, and then stayed at this percentage, as Fig. 7b shown. The in- H6 correlations of FA appeared at δC/δH 111.7/7.27 and 122.0/
crease of new C\\C bonds might attribute to the condensation reactions 7.12 ppm, respectively, and signals at δC/δH 129.9/7.50 and 143.9/
during lignin oxidation proved by MALDI-TOF/MS. In addition, owing to 7.48 ppm could be assigned to the C2,6/H2,6 and Cα/Hα of pCA. Further-
the oxidation of primary and secondary hydroxyl groups (e.g., γ-OH and more, the Cβ/Hβ signals for FA and pCA were overlapped at 115.8/
α-OH) in lignin side chains to carbonyl groups (Dabral et al., 2017; 6.35 ppm. The percentages of S, G, H, FA, and pCA in biorefinery corncob
Rahimi et al., 2013), the percentage of carbonyl carbon kept increasing lignin were separately 19.8%, 72.1%. 8.1%, 8.4%, and 9.5%. After ZrCl4/
and reached 28.7% at 16 h. To a certain extent, this matched well with NaOH-catalyzed oxidation, the intensity of all these aromatic signals
the produced monomers as discussed in Section 3.2. Combination dropped dramatically. New signals of S′2,6 were formed in the 1 h and
with the decline of the percentages of C\\O aliphatic carbon, we can 2 h products, but only correlations of G5 can be observed with oxidation
draw the conclusion that during ZrCl4/NaOH-catalyzed oxidation, corn- time prolonged to 4 h, All of these reflected the decrease of typical aro-
cob lignin underwent the oxidative cleavage of inter-unit C\\O linkages, matic units. However, as illustrated by the Q13C NMR, the percentages of
as well as the formation of C\\C bonds and carbonyl groups. aromatic carbons declined by only ca. 7% from lignin feedstock to oxida-
tion products. This contrast indicated again that ZrCl4/NaOH-catalyzed
3.5.2. 2D HSQC NMR analysis oxidation occurred on the side chains of corncob lignin, and the chemi-
Compared with Q13C NMR, 2D HSQC NMR can further reflect the cal shifts of oxidized aromatic units migrated, which could not be iden-
specific changes on aromatic units and inter-unit linkages within lignin tified with HSQC.
fragments, which has been widely used in recent years to investigate
the lignin structural evolution (Lei et al., 2019; Liu et al., 2018; 4. Conclusion
McClelland et al., 2017; Wang et al., 2017; Wen et al., 2013b). The 2D
HSQC NMR spectra of corncob lignin and its oxidation products cata- Syringyl, guaiacyl, and p-hydroxyphenyl aldehydes, ketones, acids,
lyzed by ZrCl4 and NaOH were shown in Fig. 8, and the corresponding and quinones, as well as phenol, guaiacol, and their 4-vinyl substituents
assignments of 13C\\1H correlation signals were listed in Table 4. The were identified monomeric products from the oxidative depolymeriza-
presented HSQC spectra mainly included two parts: the C\\O aliphatic tion of biorefinery corncob lignin catalyzed by ZrCl4 and NaOH in aceto-
region (δC/δH 50–95/2.5–6.0 ppm) and the aromatic region (δC/δH nitrile/water under O2. The optimal yields reached 3.7 wt% for identified
100–150/6.0–8.5 ppm). monomers and 3.1 wt% for unknown monomers. Among these mono-
In the C\\O aliphatic region, signals of the methoxy groups (OMe) mers, vanillin and 4-hydroxybenzaldehyde were the two dominant
were easily identified at δC/δH 56.1/3.8 ppm. The β-O-4′ aryl ether products. Oligomers were accompanied produced together with the
bond (substructure A) was determined to be the predominant inter- monomers, and the compositions were recognized by MALDI-TOF/MS.
unit linkage, and the related signals of α-, β- and γ-positions appeared Weight-average molecular weight (Mw) of oxidation products ranged
at δC/δH 72.3/4.8 ppm, 72.3/4.8 ppm, and 59.3/3.5 ppm, respectively. from 930 Da (1 h) to 300 Da (16 h), indicating the presence of dimers
and oligomers. Quantitative 13C and 2D HSQC NMR together demon-
Table 4 strated that the ZrCl4/NaOH-catalyzed oxidation occurred on the inter-
Assignments of 13C\ \1H correlation signals in the 2D HSQC NMR spectra of corncob lignin unit bonds, especially the β-O-4′ linkages, but not on the aromatic
and its oxidation products.
units, owing to the decreasing C\\O aliphatic carbon percentages but
Labels δC/δH (ppm) Assignments the stable aromatic carbon percentages of oxidation fragments. In addi-
OMe 55.5/3.63 C-H in methoxyls tion, a series of new C\\C aliphatic linkages and carbonyl groups were
Aγ 59.9/3.51 Cγ-Hγ in β-O-4′ substructures (A) formed, and the oxidized aromatic units cannot be measured with
Cγ 62.2/3.69 Cγ-Hγ in phenylcoumaran substructures (C) HSQC.
Aα 71.8/4.87 Cα-Hα in β-O-4′ substructures (A)
Aβ 85.8/4.12 Cβ-Hβ in β-O-4′ substructures (A)
S2,6 103.8/6.71 C2,6-H2,6 in syringyl units (S)
Acknowledgements
S′2,6 106.6/7.20 C2,6-H2,6 in Cα-oxidized syringyl units (S′)
FA2 111.1/7.27 C2-H2 in ferulate units (FA) This work was supported by the China National Funds for Distin-
G2 113.4/6.69 C2-H2 in guaiacyl units (G) guished Young Scientists (Grant no. 51525601), the Local Innovation
FAβ 115.1/6.27 Cβ-Hβ in ferulate units (FA)
and Research Teams Project of Guangdong Pearl River Talents Program
G5,6 115.5/6.79 C5,6-H5,6 in guaiacyl units (G)
pCAβ 116.4/6.40 Cβ-Hβ in p-coumarate units (pCA) (Grant no. 2017BT01N092), the Scientific Research Foundation of Grad-
FA6 122.0/7.12 C6-H6 in ferulate units (FA) uate School of Southeast University (Grant no. YBJJ1707), the Postgrad-
H2,6 127.7/7.05 C2,6-H2,6 in p-hydroxyphenyl units (H) uate Research & Practice Innovation Program of Jiangsu Province (Grant
pCA2,6 129.9/7.50 C2,6-H2,6 in p-coumarate units (pCA) no. KYCX17_0080), and the Fundamental Research Funds for the Cen-
pCAα 143.9/7.48 Cα-Hα in p-coumarate units (pCA)
tral Universities.
212 F. Lin et al. / Science of the Total Environment 675 (2019) 203–212

References Ma, R., Guo, M., Zhang, X., 2018. Recent advances in oxidative valorization of lignin. Catal.
Today 302, 50–60.
An, Y.X., Li, N., Wu, H., Lou, W.Y., Zong, M.H., 2015. Changes in the structure and the ther- McClelland, D.J., Motagamwala, A.H., Li, Y., Rover, M.R., Wittrig, A.M., Wu, C., Buchanan,
mal properties of kraft lignin during its dissolution in cholinium ionic liquids. ACS J.S., Brown, R.C., Ralph, J., Dumesic, J.A., Huber, G.W., 2017. Functionality and molecu-
Sustain. Chem. Eng. 3, 2951–2958. lar weight distribution of red oak lignin before and after pyrolysis and hydrogenation.
Barakat, N.A.M., Hassanc, A.A.Y., Matar, S.M.E.S., Awad, M.O.A., Ali, A.S.Y., 2018. ZrO2/TiO2 Green Chem. 19, 1378–1389.
nanofiber catalyst for effective liquefaction of agricultural wastes in subcritical meth- Mottweiler, J., Rinesch, T., Besson, C., Buendia, J., Bolm, C., 2015. Iron-catalysed oxidative
anol. Sep. Sci. Technol. 53, 2628–2638. cleavage of lignin and β-O-4 lignin model compounds with peroxides in DMSO.
Behling, R., Valange, S., Chatel, G., 2016. Heterogeneous catalytic oxidation for lignin val- Green Chem. 17, 5001–5008.
orization into valuable chemicals: what results? What limitations? What trends? Mu, W., Ben, H., Ragauskas, A., Deng, Y., 2013. Lignin pyrolysis components and
Green Chem. 18, 1839–1854. upgrading–technology review. BioEnergy Res 6, 1183–1204.
Biannic, B., Bozell, J.J., 2013. Efficient cobalt-catalyzed oxidative conversion of lignin Nandiwale, K.Y., Danby, A.M., Ramanathan, A., Chaudhari, R.V., Subramaniam, B., 2017.
models to benzoquinones. Org. Lett. 15, 2730–2733. Zirconium-incorporated mesoporous silicates show remarkable lignin depolymeriza-
Cao, L., Yu, I.K.M., Liu, Y., Ruan, X., Tsang, D.C.W., Hunt, A.J., Ok, Y.S., Song, H., Zhang, S., tion activity. ACS Sustain. Chem. Eng. 5, 7155–7164.
2018. Lignin valorization for the production of renewable chemicals: state-of-the Ouyang, X., Chen, L., Zhang, S., Yuan, Q., Wang, W., Linhardt, L., 2018. Effect of simulta-
art review and future prospects. Bioresour. Technol. 269, 465–475. neous steam explosion and alkaline depolymerization on corncob lignin and cellulose
Chen, W., McClelland, D.J., Azarpira, A., Ralph, J., Luo, Z., Huber, G.W., 2016. Low temper- structure. Chem. Biochem. Eng. Q. 32, 177–189.
ature hydrogenation of pyrolytic lignin over Ru/TiO2: 2D HSQC and 13C NMR study of Partenheimer, W., 2009. The aerobic oxidative cleavage of lignin to produce
reactants and products. Green Chem. 18, 271–281. hydroxyaromatic benzaldehydes and carboxylic acids via metal/bromide catalysts
Dabral, S., Hernández, J.G., Kamer, P.C., Bolm, C., 2017. Organocatalytic chemoselective in acetic acid/water mixtures. Adv. Synth. Catal. 351, 456–466.
primary alcohol oxidation and subsequent cleavage of lignin model compounds Patwardhan, P.R., Brown, R.C., Shanks, B.H., 2011. Understanding the fast pyrolysis of lig-
and lignin. ChemSusChem 10, 2707–2713. nin. ChemSusChem 4, 1629–1636.
Das, L., Xu, S., Shi, J., 2017. Catalytic oxidation and depolymerization of lignin in aqueous Rahimi, A., Azarpira, A., Kim, H., Ralph, J., Stahl, S.S., 2013. Chemoselective metal-free aer-
ionic liquid. Front. Energ. Res. 5, 21. obic alcohol oxidation in lignin. J. Am. Chem. Soc. 135, 6415–6418.
Deng, W., Zhang, H., Wu, X., Li, R., Zhang, Q., Wang, Y., 2015. Oxidative conversion of lig- Rinesch, T., Bolm, C., 2018. Cobalt-catalyzed oxidation of the β-O-4 bond in lignin and lig-
nin and lignin model compounds catalyzed by CeO2-supported Pd nanoparticles. nin model compounds. ACS Omega 3, 8386–8392.
Green Chem. 17, 5009–5018. Sales, F., Abreu, C., Pereira, J., 2004. Catalytic wet-air oxidation of lignin in a three-phase
Doherty, W.O., Mousavioun, P., Fellows, C.M., 2011. Value-adding to cellulosic ethanol: reactor with aromatic aldehyde production. Braz. J. Chem. Eng. 21, 211–218.
lignin polymers. Ind. Crop. Prod. 33, 259–276. Salonen, H.E.P., Mecke, C.P.A., Karjomaa, M.I., Joensuu, P.M., Koskinen, A.M.P., 2018. Cop-
Fan, C., Qi, K., Xia, X.X., Zhong, J.J., 2013. Efficient ethanol production from corncob resi- per catalyzed alcohol oxidation and cleavage of β-O-4 lignin model systems: from de-
dues by repeated fermentation of an adapted yeast. Bioresour. Technol. 136, velopment to mechanistic examination. ChemistrySelect 3, 12446–12454.
309–315. Sedai, B., Díaz-Urrutia, C., Baker, R.T., Wu, R., Silks, L.A.P., Hanson, S.K., 2011. Comparison
Gao, R., Li, Y., Kim, H., Mobley, J.K., Ralph, J., 2018. Selective oxidation of lignin model com- of copper and vanadium homogeneous catalysts for aerobic oxidation of lignin
pounds. ChemSusChem 11, 2045–2050. models. ACS Catal. 1, 794–804.
Hanson, S.K., Wu, R., Silks, L.A.P., 2012. C-C or C-O bond cleavage in a phenolic lignin Sedai, B., Díaz-Urrutia, C., Baker, R.T., Wu, R., Silks, L.A.P., Hanson, S.K., 2013. Aerobic oxi-
model compound: selectivity depends on vanadium catalyst. Angew. Chem. Int. Ed. dation of β-1 lignin model compounds with copper and oxovanadium catalysts. ACS
124, 3466–3469. Catal. 3, 3111–3122.
Hu, J., Zhang, S., Xiao, R., Jiang, X., Wang, Y., Sun, Y., Lu, P., 2019. Catalytic transfer Son, S., Toste, F.D., 2010. Non-oxidative vanadium-catalyzed C-O bond cleavage: applica-
hydrogenolysis of lignin into monophenols over platinum-rhenium supported on ti- tion to degradation of lignin model compounds. Angew. Chem. Int. Ed. 49,
tanium dioxide using isopropanol as in situ hydrogen source. Bioresour. Technol. 279, 3791–3794.
228–233. Stark, K., Taccardi, N., Bosmann, A., Wasserscheid, P., 2010. Oxidative depolymerization of
Huang, Y., Liu, H., Yuan, H., Zhan, H., Zhuang, X., Yuan, S., Yin, X., Wu, C., 2018. Relevance lignin in ionic liquids. ChemSusChem 3, 719–723.
between chemical structure and pyrolysis behavior of palm kernel shell lignin. Sci. Sun, Z., Fridrich, B., de Santi, A., Elangovan, S., Barta, K., 2018. Bright side of lignin depoly-
Total Environ. 633, 785–795. merization: toward new platform chemicals. Chem. Rev. 118, 614–678.
Kalliola, A., Kuitunen, S., Liitiä, T., Rovio, S., Ohra-aho, T., Vuorinen, T., Tamminen, T., 2011. Tarabanko, V.E., Tarabanko, N., 2017. Catalytic oxidation of lignins into the aromatic alde-
Lignin oxidation mechanisms under oxygen delignification conditions. Part 1. Results hydes: general process trends and development prospects. Int. J. Mol. Sci. 18, 2421.
from direct analyses. Holzforschung 65, 567–574. Vu, T.M., Trinh, V.T., Doan, D.P., Van, H.T., Nguyen, T.V., Vigneswaran, S., Ngo, H.H., 2017.
Lan, W., Amiri, M.T., Hunston, C.M., Luterbacher, J.S., 2018. Protection group effects during Removing ammonium from water using modified corncob-biochar. Sci. Total Envi-
a,γ-diol lignin stabilization promote high-selectivity monomer oroduction. Angew. ron. 579, 612–619.
Chem. Int. Ed. 57, 1356–1360. Wang, M., Lu, J., Zhang, X., Li, L., Li, H., Luo, N., Wang, F., 2016. Two-step, catalytic C–C bond
Lange, H., Decina, S., Crestini, C., 2013. Oxidative upgrade of lignin–recent routes oxidative cleavage process converts lignin models and extracts to aromatic acids. ACS
reviewed. Eur. Polym. J. 49, 1151–1173. Catal. 6, 6086–6090.
Lee, C.R., Yoon, J.S., Suh, Y.W., Choi, J.W., Ha, J.M., Suh, D.J., Park, Y.K., 2012. Catalytic roles Wang, X., Guo, Y., Zhou, J., Sun, G., 2017. Structural changes of poplar wood lignin after
of metals and supports on hydrodeoxygenation of lignin monomer guaiacol. Catal. supercritical pretreatment using carbon dioxide and ethanol-water as co-solvents.
Commun. 17, 54–58. RSC Adv. 7, 8314–8322.
Lei, M., Wu, S., Liang, J., Liu, C., 2019. Comprehensive understanding the chemical struc- Wen, J.L., Xue, B.L., Sun, S.L., Sun, R.C., 2013a. Quantitative structural characterization and
ture evolution and crucial intermediate radical in situ observation in enzymatic hy- thermal properties of birch lignins after auto-catalyzed organosolv pretreatment and
drolysis/mild acidolysis lignin pyrolysis. J. Anal. Appl. Pyrolysis 138, 249–260. enzymatic hydrolysis. J. Chem. Technol. Biotechnol. 88, 1663–1671.
Li, C., Zhao, X., Wang, A., Huber, G.W., Zhang, T., 2015. Catalytic transformation of lignin Wen, J.L., Sun, S.L., Xue, B.L., Sun, R.C., 2013b. Recent advances in characterization of lignin
for the production of chemicals and fuels. Chem. Rev. 115, 11559–11624. polymer by solution-state nuclear magnetic resonance (NMR) methodology. Mate-
Liang, C., Hu, Y., Wang, Y., Wu, L., Zhang, W., 2018. Production of levulinic acid from corn rials 6, 359–391.
cob residue in a fed-batch acid hydrolysis process. Process Biochem. 73, 124–131. Yang, Q., Shi, J., Lin, L., Zhuang, J., Pang, C., Xie, T., Liu, Y., 2012. Structural characterization
Liu, C., Hu, J., Zhang, H., Xiao, R., 2016. Thermal conversion of lignin to phenols: relevance of lignin in the process of cooking of cornstalk with solid alkali and active oxygen.
between chemical structure and pyrolysis behaviors. Fuel 182, 864–870. J. Agric. Food Chem. 60, 4656–4661.
Liu, C., Wang, X., Lin, F., Zhang, H., Xiao, R., 2018. Structural elucidation of industrial Yang, S., Wen, J.L., Yuan, T.Q., Sun, R.C., 2014. Characterization and phenolation of
bioethanol residual lignin from corn stalk: a potential source of vinyl phenolics. biorefinery technical lignins for lignin-phenol-formaldehyde resin adhesive synthe-
Fuel Process. Technol. 169, 50–57. sis. RSC Adv. 4, 57996–58004.
Liu, C., Wu, S., Zhang, H., Xiao, R., 2019. Catalytic oxidation of lignin to valuable biomass- Zhang, J., Asakura, H., van Rijn, J., Yang, J., Duchesne, P., Zhang, B., Chen, X., Zhang, P.,
based platform chemicals: a review. Fuel Process. Rechnol. 191, 181–201. Saeys, M., Yan, N., 2014. Highly efficient, NiAu-catalyzed hydrogenolysis of lignin
Ma, R., Xu, Y., Zhang, X., 2015a. Catalytic oxidation of biorefinery lignin to value-added into phenolic chemicals. Green Chem. 16, 2432–2437.
chemicals to support sustainable biofuel production. ChemSusChem 8, 24–51. Zhou, Z.Z., Liu, M., Li, C.J., 2017. Selective copper–N-heterocyclic carbene (copper-NHC)-
Ma, Y., Du, Z., Liu, J., Xia, F., Xu, J., 2015b. Selective oxidative C–C bond cleavage of a lignin catalyzed aerobic cleavage of β-1 lignin models to aldehydes. ACS Catal. 7,
model compound in the presence of acetic acid with a vanadium catalyst. Green 3344–3348.
Chem. 17, 4968–4973.

You might also like