Sunteți pe pagina 1din 10

Chemical Engineering Journal 309 (2017) 187196

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Production of phenolic hydrocarbons using catalytic depolymerization of


empty fruit bunch (EFB)-derived organosolv lignin on Hb-supported Ru
Minsun Kim a,b, Deokwon Son a,b, Jae-Wook Choi a, Jungho Jae a,c, Dong Jin Suh a,d,e, Jeong-Myeong Ha a,c,e,,
Kwan-Young Lee b,e
a
Clean Energy Research Center, Korea Institute of Science and Technology, Seoul 02792, Republic of Korea
b
Department of Chemical and Biological Engineering, Korea University, Seoul 02841, Republic of Korea
c
Department of Clean Energy and Chemical Engineering, Korea University of Science and Technology, Daejeon 34113, Republic of Korea
d
Department of Green Process and System Engineering, Korea University of Science and Technology, Daejeon 34113, Republic of Korea
e
Green School (Graduate School of Energy and Environment), Korea University, Seoul 02841, Republic of Korea

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

 Catalytic depolymerization of lignin


to monophenyl hydrocarbons.
 Ru/Hb exhibited the largest yields of
monophenyls and the smallest
molecular weight.
 Acid sites improve the
depolymerization with the help of
hydrogen-adsorbing metals.

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

Article history: Catalytic depolymerization of empty fruit bunch (EFB)-derived organosolv lignin, which is free of ash and
Received 30 May 2016 catalyst-poisoning sulfur, is performed using supported metal catalysts. Without the improved reaction
Received in revised form 1 October 2016 activity in the presence of supercritical solvents, water and ethanol in this study, at the subcritical con-
Accepted 4 October 2016
ditions, Hb-supported Ru catalyst achieved 16.5% yield of small molecule phenolic hydrocarbons, includ-
Available online 4 October 2016
ing guaiacol, 4-methylguaiacol, 4-ethylguaiacol, cerulignol, and iso-eugenol, and produced degraded
lignin polymer with 63% decreased weight-average molecular weight based on polystyrene-calibrated
Keywords:
gel permeation chromatography results. The catalytic depolymerization activity was determined by
Organosolv lignin
Catalyst
the quantity of acid sites on the catalysts; however, the presence of metal nanoparticles is required to
Depolymerization supply hydrogen atoms to the reaction system, which particles improve the catalytic depolymerization
Ru/Hb activity. The structure of organosolv lignin prepared by extraction using a mixture of ethanol and water
Solid acid was also studied using GPC and NMR results, which indicated that a possibly linear polymer composed of
phenolic monomers with two or fewer linking functionalities.
2016 Elsevier B.V. All rights reserved.

1. Introduction

Lignin has been a byproduct of traditional pulping processes


and recently developed bioalcohol production; it is used as a
Corresponding author at: Clean Energy Research Center, Korea Institute of
Science and Technology, Seoul 02792, Republic of Korea.
low-energy solid fuel for the heating units of pulping plants [1
E-mail address: jmha@kist.re.kr (J.-M. Ha). 3] and as an additive for the construction materials [4,5]. With

http://dx.doi.org/10.1016/j.cej.2016.10.011
1385-8947/ 2016 Elsevier B.V. All rights reserved.
188 M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196

increasing efforts to improve the economic feasibility of biofuels combined investigation o small-molecule aromatic hydrocarbon
and develop more valuable chemicals from biomass, aromatic products and lignin residue, the depolymerization of lignin on
monomeric units of lignin have attracted significant interest the catalysts can be explained in detail.
because they can be converted to valuable aromatic fine chemicals
that are not easy to obtain from other carbon sources including
2. Experimental
recently booming natural or shale gas and biomass-derived
carbohydrates.
2.1. Materials
Lignocellulose consists of cellulose, hemicellulose, and lignin.
Lignin is composed of 1530% by weight and 40% by energy of
Ruthenium (III) chloride hydrate (RuCl3xH2O), tetraamine plat-
the lignocellulosic biomass; it is the only sustainable resource in
inum(II) chloride hydrate (Pt(NH3)4Cl2xH2O), carbon-supported
nature that is composed of aromatics [2,68]. While cellulose
5 wt% Ru/C (5 wt% Ru/C), toluene (99%), 4-ethylguaiacol (98%),
and hemicellulose, which are important in fermentation and other
cerulignol (2-methoxy-4-propenylphenol, 99%), eugenol (2-
chemical processes to produce fuels and fine chemicals, are natural
methoxy-4-(2-propenyl)phenol, 99%), ethanol (anhydrous, 99.5%),
polymers made of well-defined sugars, lignin is a natural amor-
methyl homovanillate (methyl 4-hydroxy-3-methoxyphenylace
phous and three-dimensional complex branched polymer com-
tate, 97%), 2-chloro-4,4,5,5-tetramethyl-1,3,2-dioxaphospholane
posed of phenolic monomers, including p-coumaryl, coniferyl,
(95% TMDP), chromium (III) acetyl acetonate (99.9%), deuterated
and synapyl alcohol subunits [1]. These subunits are mainly
chloroform (99.96% CDCl3), deuterated dimethyl sulfoxide (DMSO,
cross-linked by COC ether bonds and CC bonds, while cellulose
99.9%), cyclohexanol (99%), pyridine (anhydrous, 99.8%), acetic
and hemicellulose are made of glucosidic bonds between sugars.
anhydride (99%), magnesium oxide (light), and silica-alumina
Because of its unique structure, lignin is highly touted as a poten-
powder (SiO2-Al2O3, grade 135) were purchased from Aldrich (Mil-
tial renewable resource for producing both aromatic fuels and
waukee, WI, USA). Sulfuric acid and tetrahydrofuran (HPLC grade,
chemicals; the depolymerization of lignin into its monophenyl
stabilized with butylated hydroxytoluene) were purchased from
units is required to use lignin as a feedstock for further processes.
J.T. Baker (Phillipsburg, NJ, USA). Zeolite b (ammonium form, BET
Several methods have been suggested to depolymerize lignin
surface area = 680 m2/g and Si/Al = 25 (atom/atom)) and zeolite Y
for value added products. Thermochemical methods, including
(hydrogen form, BET surface area = 760 m2/g and Si/Al = 60
pyrolysis and hydrothermal/solvothermal liquefaction, have been
(atom/atom)) were purchased from Zeolyst (Conshohocken, PA,
developed to degrade lignin [916]; however, these methods are
USA). Alumina (Al2O3, c-phase, 99.97%), silica gel (SiO2, 60325
difficult to control and have been found to lead to the formation
mesh), and ZSM-5 (ammonium form, BET surface area = 420 m2/g
of large amounts of gases, tar, and char. The significant carboniza-
and Si/Al = 50(atom/atom)) were purchased from Alfa Aesar (Ward
tion of lignin melt at high temperature has also been found to be a
Hill, MA, USA). Nickel (II) nitrate hexahydrate (Ni(NO3)26H2O) was
critical issue making the thermochemical methods difficult to scale
purchased from Junsei (Tokyo, Japan). Copper (II) nitrate trihydrate
up and commercialize. Compared to the thermochemical methods,
(Cu(NO3)3H2O) was purchased from Daejung (Seoul, Korea).
the chemical depolymerization of lignin to distillable small mole-
Hydrogen gas (H2, 99.999%), helium gas (He, 99.999%), 5% (v/v)
cules is less active for the conversion of biomass but possibly more
H2/Ar gas, 0.5% (v/v) O2/N2 gas, and 1% CO gas were purchased from
selective for valuable aromatic monomers [2].
Shinyang Sanso (Seoul, Korea). Deionized water was prepared
Because lignin can be obtained as a byproduct of pulping pro-
using the aquaMAX Ultra 370 water purification system from
cesses, a lot of lignin products have been suggested and their
YoungLin Instrument (Anyang, Korea).
chemical properties will certainly affect the operation of catalytic
depolymerization [17]. Kraft lignin, mass-produced by the Kraft
pulping process [5] and containing sulfur atoms in its structure, 2.2. Isolation of organosolv lignin from empty fruit bunch
has a complex structure and is frequently used for the depolymer-
ization research [1,4,1820]. Organosolv lignin, free of sulfur and Lignin was extracted from oil palm empty fruit bunch (EFB from
ash [21], has also been depolymerized [2224]. Ethanol and Malaysia) using the modified organosolv process [28,30,35]. The
methanol can be used for extracting lignin from lignocellulose reaction was performed in a 10 L stainless autoclave type batch
under relatively mild conditions [2527]. reactor. A mixture of biomass (1000 g), ethanol/water (5000 mL,
The objectives of this study are to produce lignin-derived small 66 wt% ethanol), and aqueous sulfuric acid solution (15 mM) was
molecule aromatic hydrocarbons and more processible lower stirred at 170 C for 1 h. The mixture was cooled to 40 C or lower.
molecular weight lignin. Organosolv lignin is selected as a feed- The obtained suspension was filtered using a Whatman type 3
stock because its sulfur-free structure cannot poison the metal cat- paper filter. The filtrate (organosolv liquor) was stored in a refrig-
alysts and because the less complex structure, compared to that of erator for lignin recovery. Organosolv lignin was precipitated in the
Kraft lignin, makes it easier for researchers to understand the organosolv liquor and washed using refrigerated water (water/so-
depolymerization behavior of lignin [2830]. Noble or transition lution = 3 (w/w)). The particles were further centrifuged and
metal catalysts, including Pt, Ru, Ni, and Cu, supported on solid freeze-dried to obtain the lignin.
acids, including metal oxides and zeolites, are studied to depoly-
merize organosolv lignin. Supercritical reaction conditions of tem- 2.3. Compositional analysis of empty fruit bunch-derived organosolv
perature above 374 C and pressure of 221 bar for water and lignin
temperature of 241 C and pressure of 63 bar for ethanol were
avoided in order to focus on the screening of optimum catalysts. The fractions of cellulose, hemicellulose, and lignin were mea-
Both lignin-derived small molecule aromatic hydrocarbon prod- sured using a method reported in the literature [3638]. The sam-
ucts and remaining lignin polymer are analyzed. The analysis of ple (0.3 g, dry basis) was treated with sulfuric acid (72 wt%) at
small-molecule aromatic hydrocarbon products determines the 30 C for 1 h and stirred every 15 min. Deionized water
most active catalysts; lignin residue was found to provide a win- (18.2 MXcm, 84 mL) was added to the mixture to obtain 4 wt%
dow into how catalytic depolymerization occurs, which knowledge sulfuric acid solution. The mixture was then heated to 121 C in
may improve understanding of the catalytic depolymerization pro- a pressure tube glass (8648-30 ACE glass) with screw-on Teflon
cess. To understand the depolymerization pathway, nuclear mag- caps and o-ring seals (5845-47 plug) and left for 1 h at 100 C.
netic resonance (NMR) spectroscopy is used [3134]. With the The mixture was cooled to room temperature and further filtered.
M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196 189

Quantities of acid-insoluble-lignin (AIL), acids, water-free solid GC/MS and GC-FID equipment to identify and quantify the prod-
residue, and ash were measured based on the methods of NREL/ ucts dissolved in toluene. Second, the ethanol-water layer, remain-
TP-510-42621 and 42622 [3638]. ing after removing the toluene layer, was dried at 50 C using a
rotavapor; this was followed by further drying in a vacuum at
2.4. Preparation of catalysts 50 C for 16 h. The remaining solid was assigned as lignin residue
in this study. The collected dry lignin residue was weighed and fur-
Commercially available supports of Al2O3, SiO2, silica-alumina ther characterized.
(SiAl), MgO, and HY were directly used without any further treat-
ment. Hb and HZSM-5 were prepared by calcining NH3-zeolite b 2.7. GC/MS and GC-FID for aromatic monomer products
and NH3-ZSM-5 at 550 C for 3 h. The support powder was sieved
to obtain particles smaller than 150 lm prior to impregnation. Aromatic monomer products dissolved in toluene were identi-
5 wt% Ru was impregnated onto each support using an incipient- fied using a GC/MS (Agilent 5975C inert XL MSD with a triple-
wetness method. All impregnated catalysts were calcined in air axis detector) equipped with a HP-5ms capillary column
at 400 C for 2 h. The calcined catalysts were reduced in a 5% (60 m  0.25 mm  0.25 lm) and quantified using a GC-FID
(v/v) H2/Ar flow at 400 C for 4 h and passivated in a 0.5% (v/v) (YoungLin 6500 series gas chromatograph system) equipped with
O2/N2 flow at room temperature for 30 min. The prepared catalysts a HP-5 capillary column (60 m  0.25 mm  0.25 lm) and a flame
were stored under ambient conditions prior to use. ionization detector (FID).

2.5. Characterizations of catalysts


2.8. Gel-permeation chromatography for polymer products

The acidity of the catalysts was measured using ammonia-


The number-average molecular weight (Mn), weight-average
temperature programmed desorption (NH3-TPD) using BELCAT-B
molecular weight (Mw), and polydispersity index (PDI) of lignin resi-
(Bel Japan, Osaka, Japan) equipped with a thermal conductivity
due were determined by GPC after acetylation. In brief, a lignin resi-
detector (TCD) and a quadrupole mass spectrometer BELMass
due sample (0.050 g) was dissolved in an acetic anhydride/pyridine
(Bel Japan, Osaka, Japan). The catalyst powder (3050 mg) was out-
(5 mL, 1:1 v/v) mixture and stirred for 24 h at room temperature.
gassed in a He flow (50 mL/min) at 400 C for 60 min and passi-
Ethanol was added to the mixture, which was rotoevaporated and
vated in a 7.5% NH3/He at 100 C for 30 min. After stabilizing TCD
further dried in a vacuum at 50 C for 16 h to obtain the acetylated
at 100 C for 60 min in a He environment (30 mL/min), desorbed
lignin. The acetylated lignin residue was dissolved in tetrahydrofu-
NH3 was measured using a TCD and a mass spectrometer at 100
ran (THF) (1.0 g/L). The GPC analysis was performed at 30 C using
900 C at a ramping rate of 10 C/min in a He flow (30 mL/min).
a THF eluent at a flow rate of 1.00 mL/min with an Agilent HPLC
The quantity of desorbed NH3 was measured in a range of
gel-permeation chromatography (GPC) system equipped with two
100500 C. CO-chemisorption was performed using a BELCAT-B
columns (Shodex LF-804) and UV detector (observed at 270 nm).
catalyst analyzer (BEL Japan, Inc.) equipped with a thermal conduc-
The calibration for the GPC was performed using the plystyrene
tivity detector (TCD). The catalyst powder (50 mg), placed in a
standard ReadyCal set (Aldrich, 2502,000,000 g/mol) [4042].
U-shaped quartz reactor, were heated to 300 C with 10 C/min
of heating rate in a He flow (50 mL/min), and reduced in a H2 flow 31
(50 mL/min) at 300 C for 1 h. The treated catalyst powder was 2.9. P, 1H, 13
C, and HSQC-NMR
cooled to 50 C in a He flow (50 mL/min). After treating the catalyst 31
powder using a He flow (50 mL/min) at 50 C for 30 min, the pulses P NMR was performed using an Agilent spectrometer at
of 1% CO in He were injected to measure the active metal surface 600 MHz with a 90 pulse angle, a 25-s pulse delay, an inverse gated
area. H2-chemisorption was performed using a Autochem 2920 proton decoupling pulse, a temperature of 298 K, and 256 scans, as
(Micromeritics) equipped with a thermal conductivity detector described in the literature [31]. For 31P NMR, lignin powder (60 mg)
(TCD). The catalyst powder (50 mg), placed in a quartz reactor, containing a relaxation agent (2.5 mg of chromium(III) acetylaceto-
were heated to 300 C with 10 C/min of heating rate in an Ar flow nate) was dried in a vacuum at 35 C for 16 h, after which N,N-
(50 mL/min), and reduced in a H2 flow (50 mL/min) at 300 C for dimethylformamide (150 lL), anhydrous pyridine (150 lL), and
1 h. The treated catalyst powder was cooled to 30 C in an Ar flow cyclohexanol (20 lL, internal standard) were added. The
(50 mL/min). After treating the catalyst powder using an Ar flow suspension was stirred for 12 h at room temperature to obtain a
(50 mL/min) at 30 C for 30 min, the pulses of 1% H2 in Ar were homogenous solution. 2-Chloro-4,4,5,5-tetramethyl-1,3,2-dioxa
injected to measure the active metal surface area. phospholane (TMDP) (200 lL) mixed with CDCl3 (400 lL) was
added to the lignin mixture and stirred for 23 min at room tem-
2.6. Catalytic depolymerization of empty fruit bunch-derived perature. The mixture was transferred to a 5 mm NMR tube for a
31
organosolv lignin P NMR analysis. 1H and 13C NMR were performed using an Agilent
spectrometer at 600 MHz. HSQC analysis was performed using a
The catalytic depolymerization of lignin was performed in a Bruker Avance 600 MHz equipped with cryoporbe installed at the
100-mL stainless steel batch autoclave reactor [20,39]. Organosolv National Center for Inter-university Research Facilities (NCIRF) at
lignin prepared in this study was dried in a vacuum at 35 C for Seoul National University (Seoul, Korea). The lignin was dried at
16 h prior to use. In a typical process, dry organosolv lignin 80 C in vacuum for 16 h and dissolved (100 g/L) in deuterated
(0.3 g) was dissolved in a mixture of ethanol and water (30 mL, dimethyl sulfoxide (DMSO) prior to the 1H, 13C, and HSQC-NMR
65% (v/v)) at 50 C for 60 min. The catalyst powder (0.1 g) was analyses.
added to the reactor along with the lignin solution. The reactor
was purged three times using H2 and pressurized to 40 bar. The 3. Results and discussion
reaction mixture was heated to 225 C and agitated at 500 rpm.
After 6 h, the reaction mixture was slowly cooled to room temper- 3.1. Composition of empty fruit bunch-derived organosolv lignin
ature. The mixture was extracted with toluene (10 mL) and frac-
tionated into two products [20,39]. First, the toluene layer The composition of organosolv lignin prepared in our lab was
containing lignin monomers was collected and observed using measured using a method reported in the literature [37], which
190 M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196

Table 1
Composition and elemental analysis results of organosolv lignin.

Composition analysis (wt%) Elemental analysis (atom%)


Cellulose Hemicellulose Lignin C H N S O
2.53.5 3.14.4 92.094.1 39.4 43.1 0 0 13.2

revealed that the fraction of lignin including acid insoluble and sol- small molecules dissolved in toluene (measurements depicted in
uble lignin components was in the range of 92.094.1 wt% Table 2), (ii) aromatic small molecules dissolved in ethanol/water
(Table 1). The amount of sugar impurities derived from cellulose mixture (not measured), (iii) lignin residue precipitated in the mix-
and hemicellulose was in the range of 5.67.9 wt%; these seem ture of toluene, ethanol, and water (not measured), (iv) lignin resi-
to have been obtained from organosolv liquor [35]. Atomic compo- due dissolved in the toluene layer (not measured), (v) lignin
sition measurement indicated that the organosolv lignin prepared residue dissolved in the ethanol/water mixture (measured), and
in this study was composed of 42.3% (atom/atom) C, 45.5% (atom/ (vi) light hydrocarbons vaporized to hydrogen-rich gas phase
atom) H, and 13.8% (atom/atom) O, with no detectable N or S (not measured). Among these, (i) was analyzed to measure yields
(Table 1). Sulfur-free organosolv lignin, compared to sulfur- of aromatic small molecules, (v) was analyzed to measure the
containing ligninsulfonates and Kraft lignin, is a suitable feedstock quantity of lignin residue, (ii) was not measured because the solu-
for a reaction using metal catalysts because sulfur-free reactants bility of aromatic compounds is low in aqueous solvents, (iii) was
do not poison the metal catalyst surface [43]. not measured because the isolation of lignin residue from catalyst
powder was difficult, (iv) was not measured because we assumed
3.2. Catalytic depolymerization of empty fruit bunch-derived that only small aromatic molecules could be extracted in toluene,
organosolv lignin and (vi) was difficult to quantify because of the dilution by large
amount of H2 gas. In this study, the small molecule products
Two isolated products, including aromatic monomers extracted extracted in the toluene layer (i) were used to measure the yield
in the toluene phase and polymeric lignin residue dissolved in the of reaction products and the efficiency of the catalytic reaction.
aqueous (ethanol-water) phase after the toluene-extraction, were The lignin residue dissolved in the ethanol/water mixture (v) was
analyzed; this analysis measured the quantity of depolymerized also analyzed to measure the degree of depolymerization. The
lignin and the degree of depolymerization, respectively. It should complete mass balance was not measured because some fractions
be noted that the lignin reactant can be converted into (i) aromatic of products were difficult to quantify.

Table 2
Concentrations and yields of major aromatic monomer products.a

Catalyst Concentration (mM in 10 mL toluene)


Guaiacol 4-methylguaiacol 4-ethylguaiacol Cerulignol Iso-eugenol Combined concentration (mM)b Yield (%)c
No catalyst 0 0 1.18 1.83 0 3.01 (0.48) 1.6
5 wt% Ru/MgO 0 0.10 1.11 2.72 0 3.93 (0.64) 2.1
5 wt% Ru/SiO2 0 0.06 0.53 2.51 0.21 3.31 (0.54) 1.7
5 wt% Ru/SiO2 (with aqueous HNO3)d 0 0 0.90 5.58 2.52 8.99 (1.48) 4.9
5 wt% Ru/Al2O3 0 0.08 1.00 4.82 0.42 6.31 (1.03) 3.4
5 wt% Ru/SiAl 0 0.15 1.24 4.28 2.92 8.60 (1.40) 4.6
5 wt% Ru/HY 0.43 0.33 2.01 11.09 1.57 15.42 (2.50) 8.2
5 wt% Ru/HZSM-5 0.56 0.67 2.94 12.21 1.08 17.45 (2.81) 9.3
5 wt% Ru/Hb 0.62 0.51 2.29 18.99 7.44 29.85 (5.02) 16.5
5 wt% Ru/Hb (spent) 0 0 0.84 6.91 0 7.75 (1.28) 4.2
5 wt% Pt/MgO 0 0.19 1.07 1.80 0.36 3.42 (0.55) 1.7
5 wt% Pt/SiO2 0 0.08 0.85 5.76 0.25 6.95 (1.14) 3.7
5 wt% Pt/Al2O3 0 0.06 1.53 3.53 0.74 5.87 (0.95) 3.2
5 wt% Pt/SiAl 0 0.12 1.04 5.90 1.36 8.43 (1.38) 4.5
5 wt% Pt/HY 0.17 0.67 1.68 15.47 1.57 19.56 (3.20) 10.6
5 wt% Pt/HZSM-5 0.46 0.43 2.05 20.76 0.27 23.98 (3.92) 13.0
5 wt% Pt/Hb 0.36 0.71 1.46 18.65 2.15 23.33 (3.82) 12.5
20 wt% Cu/MgO 0 0.06 1.10 1.22 0.26 2.63 (0.42) 1.4
20 wt% Cu/SiO2 0 0.03 0.83 1.60 0.37 2.83 (0.46) 1.5
20 wt% Cu/Al2O3 0 0.08 0.85 2.30 0.11 3.34 (0.54) 1.8
20 wt% Cu/SiAl 0 0.08 0.82 2.26 0.09 3.25 (0.53) 1.7
20 wt% Cu/HY 0 0.15 1.19 8.24 0.20 9.79 (1.61) 5.2
20 wt% Cu/HZSM-5 0.58 0.41 1.90 9.41 1.21 13.51 (2.18) 7.1
20 wt% Cu/Hb 0.29 0.17 1.09 13.78 1.96 17.30 (2.84) 9.4
20 wt% Ni/MgO 0 0.08 1.06 1.40 0.27 2.81 (0.45) 1.5
20 wt% Ni/SiO2 0 0.12 0.40 1.64 0.53 2.69 (0.44) 1.4
20 wt% Ni/Al2O3 0 0.09 0.99 4.00 2.04 7.11 (1.16) 3.8
20 wt% Ni/SiAl 0 0.18 0.84 1.92 1.21 4.16 (0.67) 2.2
20 wt% Ni/HY 0 0.10 0.37 8.95 0.05 9.46 (1.56) 5.0
20 wt% Ni/HZSM-5 0.17 0.25 1.01 9.28 3.75 14.47(2.37) 7.7
20 wt% Ni/Hb 0.13 0.31 0.95 13.17 1.13 15.88(2.61) 8.4
a
Reaction conditions: 0.1 g catalysts, 0.3 g organosolv lignin dissolved in a mixture of ethanol/water (30 mL, 65% (v/v)), 6 h reaction time, 40 bar H2, 225 C, and agitation
rate of 500 rpm.
b
Numbers in parentheses are concentrations in g/L.
c
Yield (%) = (Weight of produced guaiacol, 4-methylguaiacol, 4-ethylguaiacol, cerulignol, and iso-eugenol)/(Weight of dry lignin reactant)  100.
d
Aqueous nitric acid (66 wt%, 0.03 mL, approximately 3 equivalent of acidity of Hb) was added to the reaction mixture.
M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196 191

Table 3 The small molecule aromatic hydrocarbon products (i)


Molecular weights (Mw and Mn), and polydispersity index (PDI = Mw/Mn) of lignin and extracted with toluene were identified and quantified using GC/
lignin residue.a
MS and GC-FID to calculate the aromatic monomer yields and
Catalysts Mn (g/mol) Mw (g/mol) PDI determine the most active catalysts. The GC/MS results for the aro-
Lignin, no reaction 1333 3899 2.92 matic monomers exhibited that there were five major phenolic
No catalyst 1097 2787 2.54 compounds: guaiacol, 4-methylguaiacol, 4-ethylguaiacol, cerulig-
5 wt% Ru/MgO 1044 2465 2.36 nol (2-methoxy-4-propylphenol), and iso-eugenol (Table 2,
5 wt% Ru/SiO2 1007 2230 2.31
5 wt% Ru/Al2O3 942 2052 2.18
Figs. S1, and S2); these results are similar to results obtained using
5 wt% Ru/SiAl 980 1995 2.04 transition metals [44], Pt/C with formic acid [24], and supported
5 wt% Ru/HY 877 1662 1.90 noble metals [45,46]. Other products were ignored in this study
5 wt% Ru/HZSM-5 883 1610 1.87 because of their negligible quantities; no production of phenol or
5 wt% Ru/Hb 698 1431 2.05
catechol was observed as their formation has been reported for
5 wt% Pt/MgO 1239 2797 2.26
5 wt% Pt/SiO2 1161 2674 2.30 reactions using metal oxides [19]. Hydrodeoxygenated products
5 wt% Pt/Al2O3 1119 2551 2.28 of deoxygenated aromatic or cycloalyl molecules were not
5 wt% Pt/SiAl 1018 2303 2.26 observed although solid-acid-supported metals are known to be
5 wt% Pt/HY 1039 2247 2.16 good hydrodeoxygenation catalysts [4751]. Phenol and catechol
5 wt% Pt/HZSM-5 1120 2199 1.96
were the major products of the noncatalytic supercritical-phase
5 wt% Pt/Hb 961 1474 1.53
20 wt% Cu/MgO 1105 2775 2.51 reaction of Kraft lignin [15], but only guaiacol and alkyl guaiacol
20 wt% Cu/SiO2 1041 2709 2.60 derivatives were observed in this study. The major products of gua-
20 wt% Cu/Al2O3 859 2670 3.11 iacol and 4-alykyl-substituted guaiacols were derived from the
20 wt% Cu/SiAl 1090 2623 2.41
coniferyl alcohol units, which are the major monomeric compo-
20 wt% Cu/HY 978 2098 2.41
20 wt% Cu/HZSM-5 863 1687 1.95 nent of softwood lignin [52]. Syringol derivatives and vanillin,
20 wt% Cu/Hb 757 1536 2.03 which were observed as the products of the depolymerization of
20 wt% Ni/MgO 1186 2709 2.28 Miscanthus giganteus-derived lignin, were not observed [53].
20 wt% Ni/SiO2 1100 2698 2.45 Among the four metal particle catalysts, 5 wt% Ru and Pt exhib-
20 wt% Ni/Al2O3 1125 2605 2.31
ited better activity than 20 wt% Cu and Ni. For the Ru catalysts,
20 wt% Ni/SiAl 1118 2486 2.22
20 wt% Ni/HY 1003 2266 2.26 5 wt% Ru/Hb exhibited the largest yield of aromatic monomers,
20 wt% Ni/HZSM-5 995 2162 2.17 producing 0.62 mM guaiacol, 0.51 mM 4-methylguaiacol,
20 wt% Ni/Hb 989 2032 2.05 2.29 mM 4-ethylguaiacol, 18.99 mM cerulignol, and 7.44 mM iso-
a
Reaction conditions: 0.1 g catalyst, 0.3 g organosolv lignin dissolved in a mix- eugenol; these results demonstrated a value of 29.85 mM of com-
ture of ethanol/water (30 mL, 65% (v/v)), 6 h reaction time, 40 bar H2, 225 C, and bined concentration of five major products dissolved in 10 mL
agitation rate of 500 rpm.

Fig. 1. Representative molecular weight distributions of lignin and lignin residue after the reaction on (a) 5 wt% Ru, (b) 5 wt% Pt, (c) 20 wt% Cu, and (d) 20 wt% Ni depending
on supports.
192 M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196

Table 4 a value of Mw = 2465 g/mol. 5 wt% Ru/Hb also exhibited a sharp


NH3-TPD and CO-chemisorption results of catalysts. peak at 300 g/mol indicating the formation of lignin oligomer
Catalyst NH3-TPD CO-chemisorption H2-chemisorption (Fig. 1).
Quantity of acid sites [CO]/[Ru] [H]/[Ru] The most active catalyst, Ru/Hb, was recycled after the reaction,
(lmol/g) (mol/mol) (mol/mol) and the spent Ru/Hb, when calcined and reduced, exhibited
5 wt% Ru/MgO 0 0.0364 0.00123
decreased activity with a combined yield of 4.2%, which is only
5 wt% Ru/SiO2 0 0.0505 0.00332 25% of the yield observed for the fresh Ru/Hb (Table 2). The
5 wt% Ru/Al2O3 60 0.1251 0.00509 decreasing activity can be attributed to the loss of active sites on
5 wt% Ru/SiAl 636 0.0778 0.0118 metals and to the loss of solid acid support during the reaction
5 wt% Ru/HY 827 0.00650 0.00243
as the BET surface area of Ru/Hb decreased from 575 to 401 m2/g.
5 wt% Ru/HZSM-5 1113 0.0182 0.00133
5 wt% Ru/Hb 1375 0.0264 0.0746
3.3. Characterization of catalysts

NH3-temperature programmed desorption (NH3-TPD) and CO/


H2-chemisorption of catalysts were performed to understand the
observed catalytic activity based on acidity and metal dispersion
(Table 4, Figs. S4, and S5). Supported 5 wt% Ru catalysts were
selected for these measurements. For the acidity, zeolite-
supported Ru exhibited large quantities of acid sites. The acidity
was not well correlated with the BET surface area (Fig. S6). For
the dispersion of Ru, Ru/Al2O3 and Ru/SiAl exhibited the large ratio
of [CO]/[Ru], but zeolite (HY, HZSM-5, and Hb)-supported Ru cata-
lysts were poorly dispersed, exhibiting a ratio of [CO]/[Ru] < 0.03.
Ru/Hb exhibited the largest [H]/[Ru] ratio based on the H2-
chemisorption results. It should be noted that the catalysts with
low dispersions exhibited significantly different dispersions in
CO- and H2-chemisorption methods because of the high uncer-
tainty. Based on the results depicted in Table 4, we attempted to
correlate the dispersions of Ru and the acidities of the catalysts
with the catalytic activities, as represented by the yields of small
molecule hydrocarbons, weight-averaged molecular weights of lig-
nin residue, and quantities of remaining lignin residue. We were
not able to observe any reliable correlation between the dispersion
Fig. 2. Correlation of quantities of acid sites of Ru-based catalysts with (a) yield of
of Ru and the catalytic activity. In comparison, the acidity exhibited
small molecule hydrocarbons and (b) weight-average molecular weight (Mw). good correlation with the catalytic activity (Fig. 2). These results
indicate that the catalytic depolymerization activity is determined
by the surface acidity of the catalysts. 5 wt% Ru/Hb, which pro-
toluene from the 30 mL reaction mixture, indicating that 16.5% of duced the highest yield of major aromatic products and the lowest
the initial lignin reactants were converted to phenolic monomers. value of Mw of the lignin residues exhibited the highest total acid-
In comparison, the least active catalysts among Ru and Pt catalysts, ity (1375 lmol/g). With increasing acidity of the catalysts, the
Ru/SiO2, produced a combined concentration of 3.31 mM or 1.7% yield of the major aromatic monomer products increased, while
yield. GPC measurements of acetylated lignin and lignin residue the quantity of small molecule hydrocarbons and the weight-
were used to determine the molecular weight distribution along average molecular weight (Mw) decreased; these results indicate
with the number-average molecular weight (Mn), weight-average better depolymerization activity on more acidic catalysts. The con-
molecular weight (Mw), and polydispersity index (PDI = Mw/Mn), trol experiment using Ru/SiO2 with an addition of aqueous nitric
which depicts the depolymerization activity (Table 3, Figs. 1, and acid (approximately 3 equivalents of acidity of Hb) increased the
S3). With the initial value of Mw = 3899 g/mol of the lignin reac- formation of small molecule products 2.7 times, indicating that
tant, 5 wt% Ru/Hb, which exhibited the lowest amount of lignin the acidity improved the depolymerization of lignin (Table 2).
residue and the highest concentration of aromatic monomer We also note that, according to our previous observations, the Hb
products, produced lignin residue with a value of Mw = 1431 g/mol, used in this study is a nanostructured material with a particle size
exhibiting a 63% decrease of Mw. In contrast, 5 wt% Ru/MgO, which smaller than 50 nm [54].
exhibited the largest amount of lignin residue and the lowest Although the depolymerization activity is better correlated with
concentration of aromatic monomers, produced lignin residue with the quantity of acids, the presence of metal particles on the solid

Table 5
Effects of metals and H2 on the catalysis results.a

Catalyst Concentration (mM in 10 mL toluene)


Guaiacol 4-methylguaiacol 4-ethylguaiacol Cerulignol Iso-eugenol Combined concentration (mM)b Yield (%)c
Hb 0 0.93 2.11 1.77 2.64 7.45 (1.29) 4.25
5 wt% Ru/Hb 0 0 3.15 20.30 4.42 27.87 (4.82) 15.89
5 wt% Ru/Hb (No H2, in 40 bar He) 0 2.49 2.25 2.75 4.18 11.67 (2.02) 6.65
a
Reaction conditions: 0.1 g catalyst, 0.3 g organosolv lignin dissolved in a mixture of ethanol/water (30 mL, 65% (v/v)), 6 h reaction time, 40 bar H2, 250 C, and agitation
rate of 500 rpm.
b
Numbers in parentheses are concentrations in g/L.
c
Yield (%) = (Weight of produced guaiacol, 4-methylguaiacol, 4-ethylguaiacol, cerulignol, and iso-eugenol)/(Weight of dry lignin reactant)  100.
M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196 193

Table 6
1
H NMR results of lignin and lignin residue.

Catalyst Aromatic H Hc of b-O-4 Methoxy H Aliphatic H


(6.17.9 ppm) (4.24.8 ppm) (3.44.0 ppm) (0.62.9 ppm)
No reaction 1 0.23 0.74 0.60
No catalyst 1 0.06 1.61 0.55
5 wt% Ru/SiO2 1 0.06 1.89 0.58
5 wt% Ru/Hb 1 0.06 1.65 0.53

Table 7
31
P NMR results of lignin and lignin residue.

Catalyst Aliphatic C5-substituted Guaiacyl OH p-


used OH (144.5 guaiacyl OH (137.5.0 Hydroxyphenyl
148.0 ppm) (140.0 140.0 ppm) OH
143.5 ppm) (137.0 ppm)
No reaction 0.68 (1)a 0.03 (0.06) 1.86 (2.74) 0.02 (0.03)
No catalyst 1.79 (1) 0.03 (0.02) 1.91 (1.07) 0.03 (0.02)
5 wt% Ru/SiO2 1.91 (1) 0.08 (0.04) 1.98 (1.04) 0.07 (0.04)
5 wt% Ru/Hb 1.90 (1) 0.12 (0.06) 2.85 (1.50) 0.05 (0.03)
Fig. 3. 1H NMR results of (a) lignin, (b) lignin with no catalyst, 5 wt% Ru supported
a
on (c) SiO2, and (d) Hb. (1: Aromatic H, 2: Hc of b-O-4, 3: methoxy H, 4: ethoxy H. 5: Numbers in parentheses are relative areas of peaks regarding that of aliphatic
aliphatic H.). OH.

amount of Hc of b-O-4 decreased significantly on the lignin residue,


indicating a cleavage of b-O-4 during the catalytic depolymeriza-
tion of lignin. The formation of ethoxy H was observed because
of possible etherification between ethanol solvent and hydroxyls
of lignin.
31
P NMR was performed to observe the hydroxyl groups in lig-
nin and lignin residue; these groups may be produced by ether
cleavage during the depolymerization of lignin (Fig. 4 and Table 7)
[31]. Observed 31P NMR results indicate that organosolv lignin in
this study contained aliphatic OH, C5-substituted guaiacyl OH,
guaiacyl OH, and p-hydroxyphenyl OH. Syringyl OH at 142
144 ppm was not observed, but guaiacyl OH at 137.5140.0 ppm
was clearly observed, indicating that the organosolv lignin used
in this study was produced from softwood, actually palm empty
fruit bunch (EFB) [52]. Carboxylic groups were also observed but
ignored in this study because of their negligible amount. The
amounts of C5-substituted guaiacyl OH and guaiacyl OH of lignin
Fig. 4. 31P NMR results of (a) lignin, (b) lignin with no catalyst, 5 wt% Ru supported residue increased after the catalytic depolymerization of lignin,
on (c) SiO2, and (d) Hb. (1: aliphatic OH. 2: C5 substituted guaiacyl OH. 3: guaiacyl
which indicates ether cleavage to produce a polymeric lignin resi-
OH. 4: p-hydroxyphenyl OH.)
due with a lower molecular weight. For lignin and lignin residue,
the amount of p-hydroxyphenyl OH was negligible compared with
acids is required for depolymerization, which was confirmed dur- other hydroxyl functionalities, but its amount slightly increased
ing H2-free and metal-free control experiments (Table 5). When after the depolymerization of lignin, indicating the cleavage of
Hb without deposited metal catalysts was used, the yield of small the corresponding ether bonds. For the most active catalyst, Ru/
molecule hydrocarbons was smaller than 1/3 of that of the Ru- Hb, the combined amounts of C5-substituted guaiacyl OH and gua-
deposited catalyst, indicating that the presence of metal on the iacyl OH significantly increased indicating the greater cleavage of
solid acid supports improves the reaction. When H2 was replaced the guaiacyl ether linkages.
with He, the catalytic activity was also poor, indicating that the 2D HSQC NMR results of lignin and lignin residue. We per-
supply of hydrogen is important to initiate the reaction. From these formed two-dimensional HSQC NMR analysis to characterize the
observations, we suggest that the depolymerization occurred on changes of lignin and lignin residue structures during reactions,
the acid sites of solid acid support: the metal particles allow hydro- thus clarifying the depolymerization pathway (Figs. 5 and 6)
gen atoms to adsorb to the metal catalyst surface and react with [29,56,57]. In the aliphatic side-chain region, methoxy, b-O-4 link-
lignins ether bonds to produce cleaved hydrocarbons. Thus, H2 age, b-5 linkage, and a-O-4 linkage can be observed clearly in
seems to improve the hydrogenolysis of the ether bonds by supply- organosolv lignin. The presence of b-O-4 linkage was observed by
ing hydrogen atoms to the reactants, lignin in this study, and CH correlation for the a, b, and c positions. The CH correlation
improving the cleavage of ether bonds. in phenylcoumaran was shown for the a-O-4 and b-5 positions.
In the aromatic region of the spectrum, only the guaiacyl unit
was found for the C2/H2, C5/H5, and C6/H6 correlations. These
3.4. NMR results for empty fruit bunch-derived organosolv lignin and results confirm that the reactant used in this study was softwood
lignin residue lignin (from EFB in this study) [52]. After the reaction, cross peaks
for the b-O-4, b-5, and a-O-4 linkages decreased or even disap-
1
H NMR was performed to identify and quantify the functional- peared in the lignin residue with 5 wt% Ru/Hb. In addition, cross
ities of lignin and lignin residue (Fig. 3 and Table 6) [55]. The peaks for guaiacyl units decreased significantly in the lignin
194 M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196

Fig. 5. HSQC NMR results of organosolv lignin (A and B) and lignin residue (C and D, after the reaction using 5 wt% Ru/Hb).

residue with 5 wt% Ru/Hb. These results indicate that the cleavage
of the b-O-4, b-5, and a-O-4 linkages in the guaiacyl unit is the
main path to produce small-molecule hydrocarbons during
depolymerization of lignin.
Based on the above observations, we propose the model struc-
ture of organosolv lignin from EFB used in this study (Fig. 7); this
structure may consist of coniferyl alcohols, which are guaiacyl
units linked with b-O-4, a-O-4, and b-5 bonds. Instead of a
complex networked structure [52,58], a linear polymer composed
of guaiacols and 4-alkyl-substituted guaiacols is suggested [59]
Fig. 6. Major units in organosolv lignin. (A, b-O-4 linkage; B, b-5/a-O-4 phenyl- based on the following reasons: (i) Monomeric units obtained
coumaran, G, guaiacyl). after the depolymerization exhibited only two possible cleaved

Fig. 7. Schematic structure of organosolv lignin extracted from EFB.


M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196 195

functionalities, hydroxyl and alkyl, while the networked polymer References


can be obtained in the presence of monomers containing three or
more functionalities. (ii) The networked polymer can be obtained [1] C.R. Kumar, N. Anand, A. Kloekhorst, C. Cannilla, G. Bonura, F. Frusteri, K. Barta,
H.J. Heeres, Solvent free depolymerization of Kraft lignin to alkyl-phenolics
by the condensation of monomers at the condensation centers of using supported NiMo and CoMo catalysts, Green Chem. 17 (2015) 49214930.
ternary or quaternary carbons, which may produce branched- [2] H. Wang, M. Tucker, Y. Ji, Recent development in chemical depolymerization of
alkyl-substituted phenolic monomers by depolymerization; lignin: a review, J. Appl. Chem. 2013 (2013) 838645.
[3] X. Huang, T.I. Kornyi, M.D. Boot, E.J. Hensen, Catalytic depolymerization of
however, branched-alkyl-substituted phenolic monomers, for lignin in supercritical ethanol, ChemSusChem 7 (2014) 22762288.
example iso-propyl- or iso-butyl-substituted phenolic monomers, [4] J. Zakzeski, A.L. Jongerius, P.C.A. Bruijnincx, B.M. Weckhuysen, Catalytic lignin
were not observed among the final small molecule hydrocarbon valorization process for the production of aromatic chemicals and hydrogen,
ChemSusChem 5 (2012) 16021609.
products. It should be noted that the suggested structure of
[5] M.P. Pandey, C.S. Kim, Lignin depolymerization and conversion: a review of
organosolv lignin in this study appears to be limited to the thermochemical methods, Chem. Eng. Technol. 34 (2011) 2941.
organosolv lignin extracted by the mixed solvent of ethanol and [6] J. Long, Y. Xu, T. Wang, Z. Yuan, R. Shu, Q. Zhang, L. Ma, Efficient base-catalyzed
water mixture. The large fraction of lignin, not extracted by the decomposition and in situ hydrogenolysis process for lignin depolymerization
and char elimination, Appl. Energy 141 (2015) 7079.
ethanol/water mixture, may be different from the schematic [7] J. Zhang, J. Teo, X. Chen, H. Asakura, T. Tanaka, K. Teramura, N. Yan, A series of
structure suggested in Fig. 7. NiM (M = Ru, Rh, and Pd) bimetallic catalysts for effective lignin
hydrogenolysis in water, ACS Catal. 4 (2014) 15741583.
[8] A.J. Ragauskas, G.T. Beckham, M.J. Biddy, R. Chandra, F. Chen, M.F. Davis, B.H.
Davison, R.A. Dixon, P. Gilna, M. Keller, Lignin valorization: improving lignin
4. Conclusions processing in the biorefinery, Science 344 (2014) 1246843.
[9] D.J. Nowakowski, A.V. Bridgwater, D.C. Elliott, D. Meier, P. de Wild, Lignin fast
The catalytic depolymerization of organosolv lignin was per- pyrolysis: results from an international collaboration, J. Anal. Appl. Pyrolysis
88 (2010) 5372.
formed using supported metal catalysts. The small molecule aro- [10] A.A. Peterson, F. Abild-Pedersen, F. Studt, J. Rossmeisl, J.K. Nrskov, How
matic hydrocarbon products extracted with toluene were copper catalyzes the electroreduction of carbon dioxide into hydrocarbon
identified and quantified using GC/MS and GC-FID to calculate fuels, Energy Environ. Sci. 3 (2010) 13111315.
[11] Y. Muranaka, R. Murata, I. Hasegawa, K. Mae, Production of depolymerized
the aromatic monomer yields and to determine the most active lignin resin material from lignocellulosic biomass using acetonewater binary
catalysts. Compared to the literature on the use of supported metal solution, Chem. Eng. J. 274 (2015) 265273.
catalysts, the use of acidic zeolites as supports of the metal cata- [12] K. Kang, R. Azargohar, A.K. Dalai, H. Wang, Systematic screening and
modification of Ni based catalysts for hydrogen generation from
lysts in this study improved the production of small aromatic supercritical water gasification of lignin, Chem. Eng. J. 283 (2016) 10191032.
molecules from lignin. The solid acid-supported metal catalysts [13] D. Shen, N. Liu, C. Dong, R. Xiao, S. Gu, Catalytic solvolysis of lignin with the
work as bifunctional catalysts to depolymerize lignin by the syner- modified HUSYs in formic acid assisted by microwave heating, Chem. Eng. J.
270 (2015) 641647.
gic effect of the solid acid and metal. Ru/Hb, exhibiting the largest
[14] S. Brand, R.F. Susanti, S.K. Kim, H.-S. Lee, J. Kim, B.-I. Sang, Supercritical ethanol
quantity of acid sites, produced the largest yield (16.5%) of small as an enhanced medium for lignocellulosic biomass liquefaction: influence of
molecule monophenyl hydrocarbons and the polymeric lignin resi- physical process parameters, Energy 59 (2013) 173182.
due with the smallest weight-average molecular weight. Com- [15] H.-S. Lee, J. Jae, J.-M. Ha, D.J. Suh, Hydro- and solvothermolysis of kraft lignin
for maximizing production of monomeric aromatic chemicals, Bioresour.
pared with previous works [15,16], in this study supercritical Technol. 203 (2016) 142149.
conditions for water or alcohol were not used and the activity of [16] H.S. Lee, H. Lee, J.M. Ha, J. Kim, D.J. Suh, Production of aromatic compounds
the supported metal catalysts determined the catalytic activity; it from oil palm empty fruit bunches by hydro- and solvothermolysis, Ind. Crops
Prod. 76 (2015) 104111.
was found that Ru/Hb exhibited good depolymerization activity [17] J. Lora, W. Glasser, Recent industrial applications of lignin: a sustainable
without the support of supercritical conditions. The depolymeriza- alternative to nonrenewable materials, J. Polym. Environ. 10 (2002) 3948.
tion activity of the catalysts was best described by the quantity of [18] H.X. Ben, A.J. Ragauskas, One step thermal conversion of lignin to the gasoline
range liquid products by using zeolites as additives, RSC Adv. 2 (2012) 12892
surface acid sites, but the presence of deposited metals along with 12898.
an H2 environment also improved the activity. The NMR results [19] T. Yoshikawa, T. Yagi, S. Shinohara, T. Fukunaga, Y. Nakasaka, T. Tago, T.
indicate that the cleavage of b-O-4, a-O-4, and b-5 ether linkages Masuda, Production of phenols from lignin via depolymerization and catalytic
cracking, Fuel Process. Technol. 108 (2013) 6975.
in the guaiacyl units occurred; this was the major pathway of the [20] B. Sanyoto, A.A. Dwiatmoko, J.-W. Choi, J.-M. Ha, D.J. Suh, C.S. Kim, J.-C. Lim,
depolymerization of organosolv lignin. The NMR results along with Catalytic depolymerization of alkali lignin using supported Pt nanoparticle
the catalysis results suggest that the organosolv lignin prepared in catalysts, J. Nanosci. Nanotechnol. 16 (2016) 45704575.
[21] J. Quesada-Medina, F.J. Lpez-Cremades, P. Olivares-Carrillo, Organosolv
our lab may be a linear polymer with one or two linking function-
extraction of lignin from hydrolyzed almond shells and application of the d-
alities in its monomer units, supporting the suggestion in one value theory, Bioresour. Technol. 101 (2010) 82528260.
recent report [59], and not a networked complex polymer with [22] Y. Ye, Y. Zhang, J. Fan, J. Chang, Novel method for production of phenolics by
three or more linking functionalities in its monomer units [52,58]. combining lignin extraction with lignin depolymerization in aqueous ethanol,
Ind. Eng. Chem. Res. 51 (2011) 103110.
[23] A.K. Deepa, P.L. Dhepe, Lignin depolymerization into aromatic monomers over
solid acid catalysts, ACS Catal. 5 (2014) 365379.
Acknowledgements [24] W. Xu, S.J. Miller, P.K. Agrawal, C.W. Jones, Depolymerization and
hydrodeoxygenation of switchgrass lignin with formic acid, ChemSusChem 5
(2012) 667675.
This work was supported by the National Research Council of
[25] Z. Yuan, S. Cheng, M. Leitch, C. Xu, Hydrolytic degradation of alkaline lignin in
Science & Technology (NST) grant by the Korea government (MSIP) hot-compressed water and ethanol, Bioresour. Technol. 101 (2010) 93089313.

(No. CAP-11-04-KIST). This work was also supported by the New & [26] R. Thring, Alkaline degradation of ALCELL lignin, Biomass Bioenergy 7 (1994)
125130.
Renewable Energy Core Technology Program of the Korea Institute
[27] Z. Liu, F.-S. Zhang, Effects of various solvents on the liquefaction of biomass to
of Energy Technology Evaluation and Planning (KETEP) granted produce fuels and chemical feedstocks, Energy Convers. Manage. 49 (2008)
financial resource from the Ministry of Trade, Industry & Energy, 34983504.
Republic of Korea (No. 20143010091790). [28] M.J. Diaz, W.J.J. Huijgen, R.R. van der Laan, J.H. Reith, C. Cara, E. Castro,
Organosolv pretreatment of olive tree biomass for fermentable sugars,
Holzforschung 65 (2011) 177183.
[29] G. Hu, C. Cateto, Y. Pu, R. Samuel, A.J. Ragauskas, Structural characterization of
Appendix A. Supplementary data switchgrass lignin after ethanol organosolv pretreatment, Energy Fuels 26
(2011) 740745.
[30] W.J.J. Huijgen, J.H. Reith, H. den Uil, Pretreatment and fractionation of wheat
Supplementary data associated with this article can be found, in straw by an acetone-based organosolv process, Ind. Eng. Chem. Res. 49 (2010)
the online version, at http://dx.doi.org/10.1016/j.cej.2016.10.011. 1013210140.
196 M. Kim et al. / Chemical Engineering Journal 309 (2017) 187196

[31] Y. Pu, S. Cao, A.J. Ragauskas, Application of quantitative 31P NMR in biomass [46] A. Toledano, L. Serrano, J. Labidi, Organosolv lignin depolymerization with
lignin and biofuel precursors characterization, Energy Environ. Sci. 4 (2011) different base catalysts, J. Chem. Technol. Biotechnol. 87 (2012) 15931599.
31543166. [47] A.A. Dwiatmoko, L. Zhou, I. Kim, J.-W. Choi, D.J. Suh, J.-M. Ha,
[32] Z.C. Xia, L.G. Akim, D.S. Argyropoulos, Quantitative 13C NMR analysis of lignins Hydrodeoxygenation of lignin-derived monomers and lignocellulose
with internal standards, J. Agric. Food Chem. 49 (2001) 35733578. pyrolysis oil on the carbon-supported Ru catalysts, Catal. Today 265 (2016)
[33] D.S. Argyropoulos, 31P NMR in wood chemistry: a review of recent progress, 192198.
Res. Chem. Intermed. 21 (1995) 373395. [48] A.A. Dwiatmoko, S. Lee, H.C. Ham, J.-W. Choi, D.J. Suh, J.-M. Ha, Effects of
[34] L.G. Akim, D.S. Argyropoulos, L. Jouanin, J.C. Leple, G. Pilate, B. Pollet, C. carbohydrates on the hydrodeoxygenation of lignin-derived phenolic
Lapierre, Quantitative 31P NMR spectroscopy of lignins from transgenic compounds, ACS Catal. 5 (2015) 433437.
poplars, Holzforschung 55 (2001) 386390. [49] J.S. Yoon, J.-W. Choi, D.J. Suh, K. Lee, H. Lee, J.-M. Ha, Water-assisted selective
[35] W.J.J. Huijgen, G. Telysheva, A. Arshanitsa, R.J.A. Gosselink, P.J. de Wild, hydrodeoxygenation of lignin-derived guaiacol to monooxygenates,
Characteristics of wheat straw lignins from ethanol-based organosolv ChemCatChem 7 (2015) 26692674.
treatment, Ind. Crops Prod. 59 (2014) 8595. [50] J.S. Yoon, Y. Lee, J. Ryu, Y.-A. Kim, E.D. Park, J.-W. Choi, J.-M. Ha, D.J. Suh, H. Lee,
[36] A. Sluiter, National Renewable Energy Laboratory (U.S.), Determination of Production of high carbon number hydrocarbon fuels from a lignin-derived a-
structural carbohydrates and lignin in biomass laboratory analytical procedure O-4 phenolic dimer, benzyl phenyl ether, via isomerization of ether to alcohols
(LAP): issue date, 4/25/2008, Technical report NREL/TP-510-42618, National on high-surface-area silica-alumina aerogel catalysts, Appl. Catal. B 142143
Renewable Energy Laboratory, Golden, Colo., 2008, pp. 16 p. digital, PDF file. (2013) 668676.
[37] J.B. Sluiter, R.O. Ruiz, C.J. Scarlata, A.D. Sluiter, D.W. Templeton, Compositional [51] C.R. Lee, J.S. Yoon, Y.-W. Suh, J.-W. Choi, J.-M. Ha, D.J. Suh, Y.-K. Park, Catalytic
analysis of lignocellulosic feedstocks. 1. Review and description of methods, J. roles of metals and supports on hydrodeoxygenation of lignin monomer
Agric. Food Chem. 58 (2010) 90439053. guaiacol, Catal. Commun. 17 (2012) 5458.
[38] D.W. Templeton, C.J. Scarlata, J.B. Sluiter, E.J. Wolfrum, Compositional analysis [52] E. Adler, Lignin chemistrypast, present and future, Wood Sci. Technol. 11
of lignocellulosic feedstocks. 2. Method uncertainties, J. Agric. Food Chem. 58 (1977) 169218.
(2010) 90549062. [53] J.M.W. Chan, S. Bauer, H. Sorek, S. Sreekumar, K. Wang, F.D. Toste, Studies on
[39] B. Sanyoto, A.A. Dwiatmoko, J.-W. Choi, J.-M. Ha, D.J. Suh, C.S. Kim, J.-C. Lim, the vanadium-catalyzed nonoxidative depolymerization of Miscanthus
Highly dispersed Pt nanoparticles for the production of aromatic hydrocarbons giganteus-derived lignin, ACS Catal. 3 (2013) 13691377.
by the catalytic degrading of alkali lignin, J. Nanosci. Nanotechnol. 16 (2016) [54] J.K. Jung, Y. Lee, J.-W. Choi, J. Jae, J.-M. Ha, D.J. Suh, J. Choi, K.-Y. Lee, Production
45654569. of high-energy-density fuels by catalytic b-pinene dimerization: effects of the
[40] R.J. Stoklosa, J. Velez, S. Kelkar, C.M. Saffron, M.C. Thies, D.B. Hodge, Correlating catalyst surface acidity and pore width on selective dimer production, Energy
lignin structural features to phase partitioning behavior in a novel aqueous Convers. Manage. 116 (2016) 7279.
fractionation of softwood Kraft black liquor, Green Chem. 15 (2013) 2904 [55] J.-Y. Kim, E.-J. Shin, I.-Y. Eom, K. Won, Y.H. Kim, D. Choi, I.-G. Choi, J.W. Choi,
2912. Structural features of lignin macromolecules extracted with ionic liquid from
[41] M. Helander, H. Theliander, M. Lawoko, G. Henriksson, L. Zhang, M.E. poplar wood, Bioresour. Technol. 102 (2011) 90209025.
Lindstrm, Fractionation of technical lignin: molecular mass and pH effects, [56] J. Rencoret, G. Marques, A. Gutirrez, L. Nieto, J.I. Santos, J. Jimnez-Barbero, .
BioResources 8 (2013) 22702282. T. Martnez, J.C. del Ro, HSQC-NMR analysis of lignin in woody (Eucalyptus
[42] S. Baumberger, A. Abaecherli, M. Fasching, G. Gellerstedt, R. Gosselink, B. globulus and Picea abies) and non-woody (Agave sisalana) ball-milled plant
Hortling, J. Li, B. Saake, E. de Jong, Molar mass determination of lignins by size- materials at the gel state, Holzforschung 63 (2009) 691698.
exclusion chromatography: towards standardisation of the method, [57] A. Ragauskas, Y. Pu, R. Samuel, N. Jiang, C. Fu, Z.-Y. Wang, Structural
Holzforschung 61 (2007) 459468. characterization of lignin in wild-type versus COMT down-regulated
[43] F.G. Calvo-Flores, J.A. Dobado, Lignin as renewable raw material, switchgrass, Front. Energy Res. 1 (2014) 14.
ChemSusChem 3 (2010) 12271235. [58] W.G. Glasser, Potential role of lignin in tomorrows wood utilization
[44] Q. Song, F. Wang, J. Cai, Y. Wang, J. Zhang, W. Yu, J. Xu, Lignin depolymerization technologies, Forest Prod. J. 31 (1981) 2429.
(LDP) in alcohol over nickel-based catalysts via a fragmentation- [59] C. Crestini, F. Melone, M. Sette, R. Saladino, Milled wood lignin: a linear
hydrogenolysis process, Energy Environ. Sci. 6 (2013) 9941007. oligomer, Biomacromolecules 12 (2011) 39283935.
[45] Y. Ye, Y. Zhang, J. Fan, J. Chang, Novel method for production of phenolics by
combining lignin extraction with lignin depolymerization in aqueous ethanol,
Ind. Eng. Chem. Res. 51 (2012) 103110.

S-ar putea să vă placă și