You are on page 1of 8

Bioresource Technology 154 (2014) 1017

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

An efcient and economical process for lignin depolymerization


in biomass-derived solvent tetrahydrofuran
Jinxing Long, Qi Zhang, Tiejun Wang , Xinghua Zhang, Ying Xu, Longlong Ma
Key Laboratory of Renewable Energy, Guangzhou Institute of Energy Conversion, Chinese Academy of Sciences, Guangzhou 510640, PR China

h i g h l i g h t s
 An efcient and economic process for lignin depolymerization was achieved.
 THF advantaged excellent lignin dissolution ability and catalyst promotion effect.
 13.2% yield of phenolic monomer was obtained under optimized condition.
 This base-catalyzed process can inhibit char formation signicantly.

a r t i c l e

i n f o

Article history:
Received 30 September 2013
Received in revised form 2 December 2013
Accepted 5 December 2013
Available online 14 December 2013
Keywords:
Lignin
Phenolic monomer
Tetrahydrofuran
MgO
Depolymerization

a b s t r a c t
The depolymerization of renewable lignin for phenolic monomer, a versatile biochemical and precursor
for biofuel, has attracted increasing attention. Here, an efcient base-catalyzed depolymerization process
for this natural aromatic polymer is presented with cheap industrial solid alkali MgO and biomassderived solvent tetrahydrofuran (THF). Results showed that more than 13.2% of phenolic monomers were
obtained under 250 C for 15 min, because of the excellent lignin dissolution of THF and its promotion
effect on the catalytic activity of MgO. Furthermore, comparison characterization on the raw material,
products and residual solid using elemental analysis, FT-IR, TGDSC, PyGCMS and chemo-physical
absorption and desorption demonstrated that this base-catalyzed process can inhibit char formation signicantly. Whereas, the fact that thermal repolymerization of oligomer on the pore and surface of catalyst resulting in the declination of the catalytic performance is responsible for the residue formation.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Lignin, one of the principal components in plants, is the second
most abundant natural polymer and the unique renewable aromatic resource on the globe. Its efcient utilization has attracted
increasing attention in the last few decades (Heitner et al., 2010;
Zakzeski et al., 2010), especially in view of the increasing depletion
of fossil fuel and the growing concern about the serious environmental problem. Furthermore, phenolic monomer, one of main
products from the depolymerization of this sustained aromatic
natural polymer, is considered to be a high value-added biochemical (platform chemical) and excellent precursors for biofuel and
biopolymer (Amen-Chen et al., 2001; Barta et al., 2010). However,
current technologies for this sustainable and energy-potential
aromatic materials are still improvable due to its complexity, nonuniformity, and conjunctive bonding to other substances (Heitner
et al., 2010). For example, the yield of phenolic monomer is
relatively low and carbon deposition is remarkable in most of
Corresponding authors. Tel.: +86 20 87048614; fax: +86 20 87057789.
E-mail addresses: wangtj@ms.giec.ac.cn (T. Wang), mall@ms.giec.ac.cn (L. Ma).
0960-8524/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.biortech.2013.12.020

current acid catalytic depolymerization process (Gosselink et al.,


2012).
Due to its excellent catalytic degradation performance and insignicant carbon deposition, base-catalyzed depolymerization is considered to be a promising process for the lignin transformation to
biofuel and useful biochemical such as phenolic monomers (Dalimova, 1996; Gasson et al., 2012; Lavoie et al., 2011; Shabtai et al.,
2003a,b; Toledano et al., 2012). For example, 730% yield of liquid
products with 4.4% identiable phenolic monomers from alkali lignin had been obtained under the catalysis of NaOH (Thring, 1994).
Recently, Lavoie also reported that the yield of phenolic monomer
can be improved largely (1012%) when steam-treated lignin was
used (Lavoie et al., 2011). However, this homogeneous catalyst normally results in the difculty in the post separation. MgO is a typical
cost-efcient solid base catalyst. It has been used as a powerful
alternative for the homogeneous alkali such as NaOH and KOH in
most area, wherein good to excellent catalytic activity had exhibited. Furthermore, it has long been regarded as a robust catalyst
support (Fierro, 2006; McFarland and Metiu, 2013).
In parallel, tetrahydrofuran (THF) is relatively nontoxic and can
be obtained direct from the renewable lignocelluloses. This

J. Long et al. / Bioresource Technology 154 (2014) 1017

biomass-derived chemical is thereby widely used as a popular ecologically friendly solvent in many organic reactions, especially
when a moderately higher-boiling ethereal solvent is required
(Morrison and Boyd, 1972). Furthermore, the oxygen center of
ethers can coordinate to Lewis acids, resulting in the increase of
the basic strength of catalyst (Elschenbroich and Salzer, 1992).
Therefore, we reported here an efcient and economical process
for the depolymerization of pine lignin using this powerful and
green solvent in the presence of MgO catalyst.

11

50 mg (less than 1% as for raw lignin) were collected in capsules


during each run, the volatile gaseous compounds were not taken
into account in the mass balance. The product mixture was ltered,
and the solid fractions (including catalyst and residue) were rst
washed three times with 30 mL THF (10 mL  3). And then, it
was dissolved by 2 mol/L HCl to remove the catalyst. The residual
solid was dried at 100 C until no obvious weight loss was
detected. And the ltrate was diluted by 120 mL deionized water
to precipitate the undegraded and/or partly degraded lignin. The
phenolic monomers were extracted using CH2Cl2.

2. Methods
2.4. Products analysis
2.1. Materials
THF (HPLC grade) were purchased from Acros (Belgium) and
used as received. Other reagents are analytical grade and were supplied by Guanghua Chemical Factory Co., Ltd. (Shantou, China) and
redistilled before use. MgO was precalcined at 550 C for 5 h. The
pine lignin was separated according to the reported producer (Long
et al., 2013) and characterized using Fourier Transform Infrared
(FT-IR), Nuclear Magnetic Resonance (1H-NMR, 13C-NMR) and
Thermogravimetric-differential scanning calorimetric analyzer
(TGDSC). Results showed that there were not obvious absorbance
of carbohydrate (cellulose and hemicellulose) displayed. The dilute
acid hydrolysis experiment also showed that not any sugar was
detected. Elemental analysis demonstrated that it was composed
of 65.33% C, 5.55% H, 28.75% O and 2.45% ash.
2.2. Typical process for lignin depolymerization
0.5 g pine lignin, 4.0 mmol MgO and 40 mL reaction medium
were charged into a 100 mL stainless autoclave (316L stainless,
made by Weihai Chemical Machinery Co., Ltd.) equipped with a
mechanical agitation in sequence. After the air displacement by N2
for three times, the reactor was heated to designed temperature
for a certain time. When the reaction time had elapsed, the mixture
was cooled to room temperature using electronic fan during 30 min.
2.3. Products separation
The typical separation procedure used for the depolymerization
products was shown in Fig. 1. Since gaseous fractions of less than

Gas fraction from this depolymerization process was indentied


and measured on an Agilent 6890 Gas Chromatogram (GC) with a
Thermal Conductivity Detector (TCD) and a Flame Ionization
Detector (FID) using extranet standard method.
The volatile products including phenolic monomers were qualied on an Agilent 5890 GC with an Agilent 5975 inevt mass-selective detector and identied according to the NIST MS library
(capillary column: HP-INNOWAX, 30 m  0.25 mm  0.25 um;
programmed oven temperature: 60 C, hold 2 min at 60 C and
then ramped up with 10 C/min to 260 C and hold for another
10 min; injector: kept at 280 C in spit mode with spit ratio of
5:1; helium as the carrier gas). The molecular weight distribution
of nonvolatile products was measured by Gel Permeation Chromatography (GPC) on an Agilent 1260 HPLC apparatus with a Refractive Index Detector (RID). THF was used as eluent with the ow
rate of 1.0 mL/min. The injected sample volume was 20 lL, and
the column was kept at 30 C. The average molecular weight of
products was measured through extranet standard method wherein polystyrene was applied as standard compound.
2.5. Comparative characterizations of raw material and residual solid
The physicalchemical properties of raw lignin and residual
solid were characterized by FT-IR (PerkinElmer Spectra GX spectrometer, KBr pellets), TG (NETZSCH STA449C, ramped up from 40
to 900 C at 10 C/min), and elemental analysis (vario EL III elemental analyzer, oxygen content was estimated by the conservation of mass based on the assumption that the samples only
contain C, H, N, S and O).

Fig. 1. Procedure for product separation.

12

J. Long et al. / Bioresource Technology 154 (2014) 1017

The
pyrolysisgas
chromatography/mass
spectrometry
(PyGCMS) of raw lignin and residual solid were measured on a
Pyroprobe 5200 High Pressure Reactor connected with an Agilent
5890 GC/Agilent 5975 inevt mass-selective detector. The feedstock
was fast pyrolyzed at 900 C during 15 s, and its products were
determined by GCMS. The analysis method was as same as aforementioned volatile products analysis process.
The BrunauerEmmettTeller (BET) surface area, average pore
diameter and pore volume of catalysts were determined by N2 isothermal (77 K) adsorption using the Micrometrics ASAP-2010
automated system.

the widely used solvent for lignin transformation (Beauchet et al.,


2012). It also showed that both degree of conversion and yield of
phenolic monomers were sharply increased when additional water
was presented. It is considered that lignin liquefaction accompanying with hydrolysis in this process is responsible for it. Both raw
lignin and catalyst used in this process are insoluble in water
resulting in the contraction difculty between catalyst and
feedstock. Therefore, no more than 42% conversion with few of
phenolic monomers was obtained. Table 1 also showed that the
most abundant conversion and yield of phenolic monomer were
achieved when THF was used. The excellent dissolution capacity
of THF for the lignin is responsible for it. Its promotion effect on
the base strength of MgO catalyst resulting in higher catalytic
performance (Elschenbroich and Salzer, 1992) is contributed as
well.
The effect of catalyst dosage, reaction temperature and retention time had also been examined carefully. Result showed that
this process is more catalyst dependence, and then temperature,
whereas the effect of reaction time is relatively insignicant. As
shown in Fig. 2a, both the degree of conversion and the yield of
phenolic monomers were increasing with the increase of catalyst
dosage when it was lower than 4.0 mmol. Normally, higher catalyst dosage indicates much more catalytic active center resulting
in higher catalytic performance. However, this catalyst catalyzes
not only lignin depolymerization but also the repolymerization of
the oligomer of products. Therefore, the yield of phenolic monomer
was slightly declined, whereas it of residual solid was increased
when 5.0 mmol MgO was imposed.
The reaction conditions such as temperature and retention time
were also examined (Fig. 2b and c). Results showed that this process is high temperature dependence. The conversion of lignin
and the yield of phenolic monomer were increased remarkably
with the elevation of reaction temperature when it was lower than
250 C. At the elevated temperature, for example 280 C, 100% lignin can be converted. However, more than 46% of it was converted
to residual solid fraction, meanwhile; the yield of phenolic monomer was slight declined. GPC analysis results of depolymerization
products listed in Table 2 also demonstrated clearly that the
average molecular of this fraction (Mw) was rst decreased gradually (220250 C) and then increased signicantly (280 C) with
the increase of reaction temperature. It had been reported that
the unsaturated bond in lignin molecular can repolymerize when
the temperature is higher than 240 C (Heitner et al., 2010). Therefore, this increase on the yield of residual solid and the average
molecular weight of liquid products can be attributed to the repolymerization of phenolic oligomer driven by thermal effect. Compared with the reaction temperature, the effect of retention time
is relatively insignicant thought the same changing trend can be
observed (Fig. 2c). For instance, more than 13.2% yield of total phenolic monomer was achieved for 15 min. However, this value was
declined slightly to 10.2% thought the yield of residual solid was
increased caused by the repolymerization at prolonged time of
60 min (Fig. 2c).

2.6. Measurements of the conversion of lignin and the yield of phenolic


monomer
The degree of conversion of lignin and the yield of residual solid
fraction were calculated according to eqs. (1) and (2) through the
weight comparison of recovered lignin, residual solid and feed lignin respectively.

Degree of conversion % W F  W R =W F  100%

Yeild of residual solid % W S =W F  100%

WF: the weight of feed lignin; WR: the weight of recovered lignin; WS: the weight of residual solid.
The content of phenolic monomers was measured by GC at
the same capillary column (HP-INNOWAX, 30 m  0.25 mm 
0.25 um) and temperature program as the GCMS analysis.
Acetophenone was used as internet standard compound. The
yields of phenolic monomer, phenol, guaiacol and syringol were
evaluated according to the following equations.

Yield of total phenolic monomer % W T =W F  100%

Yield of phenol % W P =W F  100%

Yield of guaiacol % W G =W F  100%

Yield of syringol % W S =W F  100%

WT: the weight of total phenolic monomer; WF: the weight of


feed lignin; WP: the weight of phenol; WG: the weight of guaiacol;
WS: the weight of syringol.
3. Results and discussion
3.1. Base-catalytic depolymerization of lignin
The effect of reaction solvent on the conversion of lignin and the
products distribution were rst investigated. Results shown in
Table 1 demonstrated that this process was inuenced signicantly
by the reaction medium. For example, more than 84.5% degree of
conversion with 8.45% yield of total phenolic monomer was
achieved in the low carbon alcohol such as methanol and ethanol,

Table 1
The effect of reaction mediuma.
Solvent

CH3OH
C2H5OH
C2H5OH/H2O
THF
H2O
a
b

Conversion of lignin (%)

90.7
84.5
92.5
97.5
42.3

Yield of phenolic monomer (%)

Yield of residual solid (%)

Phenol

Guaiacol

Syringol

Others

0.25
0.25
0.82
0.75
0.23

1.36
1.42
1.58
2.81
0.70

1.19
1.24
1.32
1.45
0.52

6.46
5.54
7.52
7.79
0.74

Condition: lignin 0.5 g; MgO 3.0 mmol; solvent 40 mL; 250 C; 30 min.
No residual solid was found after the dissolution using THF.

37.5
37.2
32.2
38.5
n.d.b

J. Long et al. / Bioresource Technology 154 (2014) 1017

13

Fig. 2. Effect of (a) catalyst dosage, (b) reaction temperature, (c) retention time on the depolymerization of lignin and (d) the molecular weight distribution of products (curve
(a) raw lignin, (b) without catalyst; (c) 1.0 mmol; (d) 2.0 mmol; (e) 4.0 mmol; (f) 5.0 mmol) Conditions: (a) and (d): lignin 0.5 g, THF 40 mL, 250 C, 30 min; (b): lignin 0.5 g,
MgO 4 mmol, THF 40 mL, 30 min; (c) lignin 0.5 g, MgO 4 mmol, THF 40 mL, 250 C.

Table 2
Average molecular weight of depolymerization products from various temperaturesa.
Temperature (C)

Mnb

Mwb

Mzb

Db

220
230
240
250
280
Raw lignin

573
561
585
543
600
892

891
836
806
792
884
1360

1332
1184
1074
1126
1211
1579

1.55
1.49
1.38
1.46
1.47
1.52

Conditions: lignin 0.5 g, MgO 4.0 mmol, THF 40 mL, 30 min.


Mn: number average molecular weight; Mw: weight average molecular weight;
Mz: Z-average molecular weight; D: dispersion degree.
b

3.2. Products analysis


The gaseous product was analyzed by GCTCDFID, and its content was measured using external standard method. Result showed
that it was mainly composed of CH4, CO, CO2 and C2C3 hydrocarbon. However, the total weight of these compounds were less than
1% of the feedstock (lignin), therefore, it was not calculated during
the next material balance. GCMS analysis (Table 3) on the volatile
products exhibited that this process is complex on not only the
process but also the products, in which more than 50 kinds of

chemicals were observed. It also indicates that the basic structural


units (H, G and S unit) of pine lignin were all converted under this
condition, resulting in phenol, guaiacol, syringol and their deviates
respectively. Furfural derivates such as furfural and 2,5-dimethyl
furan, which generally originated from the depolymerization of lignin-carbohydrate complex (LCC) unit of lignin (Heitner et al., 2010)
was also found, but both their species and contents were much
more insignicant than aromatic compounds.
GPC curves of nonvolatile fractions (Fig. 2d) demonstrated that
catalyst has a signicant effect not only on the conversion and
yield of phenolic monomers but also on the products distribution.
For example, relatively high catalyst dosage resulted in higher conversion and yield of phenolic monomers, but, both of them had not
increased continuously while slightly declined with the further increase of catalyst dosage to 5.0 mmol due to the repolymerization
of products and ever the feedstock (Fig. 2a). GPC results conrmed
this conclusion. As shown in Fig. 2d, the average molecular weight
of nonvolatile fractions was decreased gradually when the added
catalyst was less than 4.0 mmol, and then it was increased sharply
with the continuously increasing catalyst dosage. It also demonstrated that the dispersion degree of products was increased at
higher catalyst dosage. For example, the peak of GPC curve moved
apparently to the large molecular weight when 5.0 mmol catalyst

14

J. Long et al. / Bioresource Technology 154 (2014) 1017

Table 3
The most 24 abundant chemicals in volatile fraction.
RT (min)

Compound

Structure

RT (min)

Compound

4.26

1-Butanol

HO

14.84

Phenol

4.63

3-Buten-1-ol

HO

15.07

Phenol, 4-ethyl-2-methoxy

8.55

Acetic acid

15.58

Phenol, 4-methyl-

8.72

Furfural

15.83

Phenol, 2-methoxy-4-propyl

16.45

Phenol, 3-ethyl-

16.70

3-Hydroxy-4-methoxybenzoic acid

Structure

OH
OH
O
OH

OH
O

HO
O

10.01

Pentanoic acid, 4-oxo-, methyl ester

O
O

HO

10.49

Pentanoic acid, 4-oxo-, ethyl ester

O
O
O

10.82

HO

13.34

16.89

Phenol, 2,3,6-trimethyl-

17.26

Phenol, 2,6-dimethoxy-

OH
OH

Phenol, 2-methoxy-

OH

13.45

O
OH

OH

2-Methoxy-5-methylphenol

OH

O
O

Butanoic acid, 4-hydroxy-

O
HO

18.02

Benzoic acid, 4-hydroxy-3-methoxy-

OH

HO
14.19

Phenol, 4-methoxy-3-methyl-

HO

18.50

Benzene, 1,2,3-trimethoxy-5-methyl-

O
O

O
14.33

Phenol, 2-methoxy-4-methyl-

OH

19.08

Methyl (3,4-dimethoxy phenyl) (hydroxy)acetate

OH

14.80

Phenol, 2-methyl-

OH

20.62

2-Propanone, 1-(4-hydroxy-3-methoxyphenyl)-

HO

O
Condition: lignin 0.5 g, MgO 4 mmol, THF 40 mL, 250 C, 30 min; Analysis condition: capillary column: HP-INNOWAX, 30 m  0.25 mm  0.25 um; programmed oven
temperature: 60 C, hold 2 min at 60 C and then ramped up with 10 C/min to 260 C and hold for another 10 min; injector: kept at 280 C in spit mode with spit ratio of 5:1;
helium as the carrier gas.

was presented. At the same time, the chemicals with lower molecular weight were also observed. It implies that both depolymerization and the repolymerization of lignin and the products are
occurred during this process, and this phenomenon becomes
remarkable at higher catalyst concentration.
3.3. Characterization of raw lignin and the residues
In order to examine the reasons for the formation of catalytic
residues in this process, the physicalchemical properties of raw
lignin and the residual solids from various catalyst dosages were
characterized intensively using elemental analysis, FT-IR, TG,
PyGCMS and BET characterization.
Main composed elements listed in Table 4 showed that the raw
lignin composed of 65.33% C, 5.55% H and 28.75% O with a higher
heating value (HHV) of 24.88 MJ/kg. However, the carbon content
of the residual solid from the depolymerization process was much
higher than the raw lignin. Especially, it in the residual solid A was
increased signicant to 76.93%, whereas the hydrogen and oxygen
content were decreased to 4.61% and 18.15% respectively. These
values were very near to their contents in the reported biochar
(Keiluweit et al., 2010). It indicates that the polymerization and

dehydration of aromatic lignin rather than decomposition are


occurred without catalyst. It also explained well that scarce phenolic monomers were achieved under the single thermal effect
(Fig. 2a). Noticeably, compared with the solid from noncatalytic
system, the carbon content was decreased obviously with the addition of MgO and the increase of its amount. Relatively, the H/C and
O/C in the residual solid were increased, and the HHV was
declined. As aforementioned discussion, the repolymerization of
lignin and the oligomer during this process is responsible for it.
The evolution of FT-IR spectra of raw lignin and residual solid
from various catalyst dosages was shown in Fig. 3a and Table 5.
It has been reported that the appearance of peaks at 34503300,
2935 and 2841 cm1 in spectra is due to the stretching of OH,
OCH3 and CH2 bonds, respectively (Fushimi et al., 2009). The
peaks at 1606, 1514, 1462 and 843 cm1 were assigned as the
characteristic vibrations of the benzene structures in lignin
(Fushimi et al., 2009). 1713 cm1 was regarded as the characteristic stretching vibration of the carbonyl (Long et al., 2011).
875 cm1 was considered to be the characteristic vibrations of
LCC (Heitner et al., 2010). The peaks at 1267, 1115 and 843 cm1
were considered to correspond to the characteristic vibrations of
the guaiacyl units. The peak at 1215 cm1 was asscociated with

15

J. Long et al. / Bioresource Technology 154 (2014) 1017


Table 4
The main composed elements of raw lignin and liquefaction residuesa.
Material

Raw lignin
Residue A (0.0 mmol)
Residue B (2.0 mmol)
Residue C (4.0 mmol)
Residue D (5.0 mmol)
a
b
c

Elemental content
b

65.33
76.93
71.41
69.33
68.97

5.55
4.61
4.30
4.29
4.84

28.75
18.15
23.95
26.06
25.85

0.26
0.27
0.25
0.27
0.22

0.11
0.04
0.09
0.05
0.12

Experimental molecular formula

HHV (MJ/kg)c

C9H9.17O2.97N0.03S0.006
C9H6.47O1.59N0.03S0.002
C9H6.50O2.26N0.03S0.004
C9H6.68O2.54N0.03S0.002
C9H7.57O2.53N0.02S0.006

24.88
29.35
26.02
24.92
25.62

On a dry basis (the sample was dried at 120 C until without further weight loss occurred).
The oxygen content was estimated by the conservation of mass based on the assumption that the samples only contain C, H, N, S and O.
Evaluated by Dulong Formula: HHV (MJ/kg) = 0.3383  C + 1.422  (H  O/8).

Fig. 3. FT-IR spectra (a) and TG curves (b) of raw lignin and residual solid from various catalyst dosages curve (a) raw lignin; (b) without catalyst; (c) 2.0 mmol; (d) 4.0 mmol;
(e) 5.0 mmol.

Table 5
FT-IR spectra of lignin and the residual solids.
Band position (cm1)

Functional groups

34503300
2935
2841
1713, 1696
1606, 1514, 1462
1267, 1215, 1115, 1032
875
843

OAH stretching vibration


CAH stretching vibration of OCH3
CAH stretching vibration of CH2
C@O stretching vibration
Aromatic C@C ring breathing
ArAO stretching breathing
CAO stretching vibration
Aromatic CAH ring out-of-plane vibration
breathing

All spectra were assigned according to the reported literature. Mayo et al. (2004).

ring breathing with CAO stretching of both the syringyl and guaiacyl structures (Alriols et al., 2009). Fig. 3a showed that the differences between the FT-IR spectra of raw lignin and the residual
solid from various catalyst dosages were remarkable. To summarize, all spectra of residual solid with catalyst had not exhibited
the characterized absorbance of biochar as compared with the results from Keiluweit (Keiluweit et al., 2010). The peak strength at
843 cm1 which is considered to be the characteristic absorption
of guaiacyl units was decreased obviously during the depolymerization process. Instresting, the characteristic stretching vibration

of LCC (875 cm1) was changed signicantly with the addition of


catalyst. As shown in Fig. 3a, it almost disappeared thoroughly under the thermal effect but intensied signicantly with catalyst,
and the strongest absorption was obtained under the optimized
catalyst amount (4.0 mmol). And thus, FT-IR spectra demonstrated
clearly that the guaiacyl units of lignin is much easier degraded
than sypringol one resulting in high content of volatile guaiacol
and its derivates as detected by GCMS (Table 3). It also demonstrated that LCC part in lignin is more recalcitrant during this process thought is generally exible than the aromatic structure. The
inductive effect of MgO and the bezene ring in lignin molecular is
considered to attributed it.
The thermochemical proprieties of these lignin and residual
solid were further investigated. TG curve of raw lignin showed
two obvious weight loss stages with the temperature range of
150200 C and 240900 C respectively. It can be attributed to
the decomposition of LCC and lignin respectively. When it was
treated by THF, the thermal decomposition property was changed
remarkably. The residue from the noncatalytic system with higher
carbon content began to decompose at 280 C, leaving 71.2% of the
indecomposable solid at 900 C. However, the onset decomposition
temperature (Tonset) of residual solid from catalytic system was
declined to 200 C and it decomposed slowly with increasing temperature when it was lower than 400 C. It is considered that the

16

J. Long et al. / Bioresource Technology 154 (2014) 1017

partly broken interior chemical bond and hydrogen bond in the lignin structural units is responsible for this lower Tonset. Generally,
the weight loss on the TG curve reects the formation of volatile
products. It also had reported that nonvolatile hemichar and tar
was rst formatted during the pyrolysis of biomass (Huang et al.,
2013). And these hemichar and tar can be further decomposition
and dehydration to give much higher carbon content solid such
as char and carbon (Huang et al., 2013). Therefore, a quick weight
loss was observed when the temperature was higher than 450 C
on the TG curve of residual solids (Fig. 3b). Fig. 3b also demonstrated that the residual solid from 5.0 mmol MgO exhibited a
much nearer TG curve as the raw material with higher Tonset and
lower nonvolatilized component compared with the others. It
further conrmed the aforementioned conclusion that catalyst
has a signicant effect on the lignin depolymerization, however,
much higher catalyst amount results in the repolymerization of
oligomer and lignin (Fig. 2a and d).
PyGCMS analysis of the raw lignin and the residual solid from
optimized condition was shown in Fig. S1. Result indicated that the
main component of lignin had not changed signicantly during
this base-catalyzed process in THF. But, the components and contents of aromatic pyrolysis products (RT = 13.225.0 min) were decreased after the treatment of THF because of its depolymerization
for phenolic products (Table 3). Compared with aromatic fraction,
the difference between furfural and its derivates (RT = 8.6
13.2 min), which are generally come from the LCC part of lignin,
was much more insignicant (Fig. S1). It implied that LCC is more
recalcitrant than aromatic fraction during this process. This result
accords well with the afomentioned FT-IR spectra (Fig. 3a) and
GCMS results (Table 3).
In order to understand the main cause for the formation of
residual solid, BrunauerEmmettTeller (BET) characterization of
fresh and used catalyst (Fig. S2) were carried out. It showed that
the surface area of catalyst was declined sharply from 22.18 to
1.24 m3/g (Table S1) after this process. At the same time, the pore
structure of used catalyst was almost thoroughly disappeared.
With the help of solvent and catalyst, lignin polymer is rst decomposed to oligomer (Song et al., 2013), which can enter into the pore
of catalyst. And then it further decomposes for small molecular
chemicals such as phenolic monomers which leave the catalyst
pore by the extraction of solvent. However, this oligomer can also
repolymerize under the thermal effect and catalysis, resulting in
the increase of the molecular diameter. When the size of this reaggregation is larger than the catalyst pore size, clogging of internal
pores would likely be occurred. As a result, the catalyst surface
area, pore volume and pore size were reduced.
And thus, elemental analysis, FT-IR, TG, PyGCMS and BET
results demonstrated that the main functional group, thermochemical property and the main elemental components of residual
solid from catalytic system had not changed obviously. However,
the surface and the pore diameter of catalyst were all decreased
sharply after the catalytic run. Hence, it is considered that the formation of solid in this process is mainly caused by the repolymerization of unsaturated products on the catalyst surface and its pore,
whereas the carbon deposition on this catalyst is insignicant.
4. Conclusion
An efcient and economical catalytic system for the transformation of lignin was achieved. Due to its excellent dissolution
capability for lignin and the promotion effect for catalyst, THF is
found to be an efcient reaction medium for lignin depolymerization. And 13.2% yield of useful phenolic monomers had been obtained. Detailed analysis on the raw material and residual solid
demonstrated that the residue is mainly caused by the repolymerization of unsaturated products on the catalyst surface and pore.

No obvious char was observed. And thus, this process will be a benecial reference for the future utilization of this aromatic sustained
material.
Acknowledgements
The authors gratefully acknowledge the nancial support of the
Natural Science Foundation of China (Nos. 51306191, 51106166),
National Basic Research Program of China (973 program, Project
No. 2012CB215304), the Key Research Program of the Chinese
Academy of Science (No. KGZD-EW-304-3), and the International
S&T Cooperation Program of China (No. 2012DFA61080).
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.biortech.2013.
12.020.
References
Alriols, M.G., Tejado, A., Blanco, M., Mondragon, I., Labidi, J., 2009. Agricultural palm
oil tree residues as raw material for cellulose, lignin and hemicelluloses
production by ethylene glycol pulping process. Chem. Eng. J. 148, 106114.
Amen-Chen, C., Pakdel, H., Roy, C., 2001. Production of monomeric phenols by
thermochemical conversion of biomass: a review. Bioresour. Technol. 79, 277
299.
Barta, K., Matson, T.D., Fettig, M.L., Scott, S.L., Iretskii, A.V., Ford, P.C., 2010. Catalytic
disassembly of an organosolv lignin via hydrogen transfer from supercritical
methanol. Green Chem. 12, 16401647.
Beauchet, R., Monteil-Rivera, F., Lavoie, J.M., 2012. Conversion of lignin to aromaticbased chemicals (L-chems) and biofuels (L-fuels). Bioresour. Technol. 121, 328
334.
Dalimova, G.N., 1996. Alkaline hydrolysis of natural cottonplant stem lignin in the
presence of demethylated lignin. Chem. Nat. Compd. 32, 928929.
Elschenbroich, C., Salzer, A., 1992. Organometallics: A Concise Introduction, second
ed. Wiley-VCH, Weinheim.
Fierro, J.L.G., 2006. Metal Oxides Chemistry and Applications. CRC Taylor & Franci,
Boca Raton.
Fushimi, C., Katayama, S., Tasaka, K., Suzuki, M., Tsutsumi, A., 2009. Elucidation of
the interaction among cellulose, xylan, and lignin in steam gasication of
woody biomass. AIChE J. 55, 529537.
Gasson, J.R., Forchheim, D., Sutter, T., Hornung, U., Kruse, A., Barth, T., 2012.
Modeling the lignin degradation kinetics in an ethanol/formic acid solvolysis
approach. Part 1. Kinetic model development. Ind. Eng. Chem. Res. 51, 10595
10606.
Gosselink, R.J.A., Teunissen, W., van Dam, J.E.G., de Jong, E., Gellerstedt, G., Scott, E.L.,
Sanders, J.P.M., 2012. Lignin depolymerisation in supercritical carbon dioxide/
acetone/water uid for the production of aromatic chemicals. Bioresour.
Technol. 106, 173177.
Heitner, C., Dimmel, D., Schmidt, J., 2010. Lignin and Lignans: Advances in
Chemistry. CRC Taylor & Franci, Boca Raton.
Huang, Y.F., Chiueh, P.T., Kuan, W.H., Lo, S.L., 2013. Pyrolysis kinetics of biomass
from product information. Appl. Energy 110, 18.
Keiluweit, M., Nico, P.S., Johnson, M.G., Kleber, M., 2010. Dynamic molecular
structure of plant biomass-derived black carbon (biochar). Environ. Sci. Technol.
44, 12471253.
Lavoie, J.-M., Bare, W., Bilodeau, M., 2011. Depolymerization of steam-treated lignin
for the production of green chemicals. Bioresour. Technol. 102, 49174920.
Long, J., Guo, B., Teng, J., Yu, Y., Wang, L., Li, X., 2011. SO3H-functionalized ionic
liquid: efcient catalyst for bagasse liquefaction. Bioresour. Technol. 102,
1011410123.
Long, J., Li, X., Guo, B., Wang, L., Zhang, N., 2013. Catalytic delignication of
sugarcane bagasse in the presence of acidic ionic liquids. Catal. Today 200, 99
105.
Mayo, D.W., Miller, F.A., Hannah, R.W., 2004. Course Notes on the Interpretation of
Infrared and Raman Spectra. John Wiley & Sons, New Jersey.
McFarland, E.W., Metiu, H., 2013. Catalysis by doped oxides. Chem. Rev. 113, 4391
4427.
Morrison, R.T., Boyd, R.N., 1972. Organic Chemistry, second ed. Allyn and Bacon,
Boston.
Shabtai, J.S., Zmierczak, W.W., Chornet, E., Johnson, D., 2003a. Conversion of
biomass into blending component for petroleum-derived fuel by
depolymerizing lignin feed material in aqueous solvent to provide rst
composition, and hydroprocessing rst composition to provide second
composition, America Pat., US2003115792-A1, 2003.
Shabtai, J.S., Zmierczak, W.W., Chornet, E., Johnson, D., 2003b. Conversion of
biomass into blending component for petroleum-derived fuel, comprises

J. Long et al. / Bioresource Technology 154 (2014) 1017


extracting lignin-containing fraction in reaction medium from biomass to
provide lignin feed material. America Pat., US2003100807-A1, 2003.
Song, Q., Wang, F., Cai, J., Wang, Y., Zhang, J., Yu, W., Xu, J., 2013. Lignin
depolymerization (LDP) in alcohol over nickel-based catalysts via a
fragmentation-hydrogenolysis process. Energy Environ. Sci. 6, 9941007.
Thring, R.W., 1994. Alkaline degradation of Alcell(R) lignin. Biomass Bioenergy 7,
125130.

17

Toledano, A., Serrano, L., Labidi, J., 2012. Organosolv lignin depolymerization with
different base catalysts. J. Chem. Technol. Biotechnol. 87, 15931599.
Zakzeski, J., Bruijnincx, P.C.A., Jongerius, A.L., Weckhuysen, B.M., 2010. The catalytic
valorization of lignin for the production of renewable chemicals. Chem. Rev.
110, 35523599.

You might also like