Biochar Thermocatalytic

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Chemical Engineering Journal 356 (2019) 461–471

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Catalytic depolymerization of organosolv lignin from bagasse by T


carbonaceous solid acids derived from hydrothermal of lignocellulosic
compounds
Panuruj Asawaworarita, Pornlada Daorattanachaia, Weerawan Laosiripojanab,

Chularat Sakdaronnarongc, Artiwan Shotiprukd, Navadol Laosiripojanaa,
a
The Joint Graduate School of Energy and Environment, King Mongkut’s University of Technology Thonburi, Prachauthit Road, Bangmod, Bangkok 10140, Thailand
b
Department of Tool and Materials Engineering, Faculty of Engineering, King Mongkut’s University of Technology Thonburi, Prachauthit Road, Bangmod, Bangkok 10140,
Thailand
c
Department of Chemical Engineering, Faculty of Engineering, Mahidol University, 25/25 Phutthamonthon 4 Road, Salaya, Phutthamonthon, Nakorn Pathom 73170,
Thailand
d
Department of Chemical Engineering, Faculty of Engineering, Chulalongkorn University, Thailand

H I GH L IG H T S

• Catalytic depolymerization of lignin from organosolv fractionation of bagasse was studied.


• Carbonaceous solid acid catalysts were synthesized from hydrothermal of sugar, cellulose, lignin.
• Catalyst from hydrothermal of lignin showed the highest catalytic activity and stability.
• Methyl isobutyl ketone (MIBK) is the best solvent for the reaction.
• 4-ethylphenol and guaiacol were the major phenolic products from the reaction.

A R T I C LE I N FO A B S T R A C T

Keywords: In the present work, depolymerization of isolated lignin from organosolv fractionation of bagasse in the presence
Organosolv lignin of several homogeneous and heterogeneous acid catalysts was studied. The catalysts included homogeneous
Depolymerization acids (i.e. H2SO4, H3PO4 and HCl), and heterogeneous carbonaceous solid acids (CSAs) synthesized from the
Hydrothermal carbon hydrothermal treatment of glucose, cellulose and lignin. It was found that the highest phenolic monomers yield
Phenolic monomers
was achieved from CSA derived from lignin (CSAlignin). In addition, this catalyst showed the highest activity and
stability after 5 consecutive cycles, which was related to its high acidity as well as its low acid leaching compared
to CSACellulose and CSAGlucose. The effects of catalyst loading, reaction temperature, reaction time, and solvent
type (i.e. methanol, ethyl acetate, methyl isobutyl ketone (MIBK), acetone) on the depolymerization perfor-
mance were also performed. It was observed that the reaction at 350 °C for 3 h using MIBK as solvent with
CSAlignin loading of 10 wt% gave the maximum total phenolic yield of 32.8%, from which 4-ethylphenol and
guaiacol were the major phenolic products from the reaction. The results in the work suggest that catalytic
depolymerization of organosolv lignin in the presence of CSA is a promising way for production of high value-
added chemicals.

1. Introduction been widely investigated globally. Generally, lignocellulosic biomass


consists mainly of three organic compounds including cellulose
Nowadays, fossil feedstock (i.e. crude oil and natural gas) con- (40–50%), hemicellulose (20–30%), and lignin (10–25%) [1]. Among
sumption has been increasing due to the high demand of petro-based these compounds, lignin represents the most abundant organic mate-
chemicals, fuels, and energies. Therefore, the concept of replacing these rial, which could be efficiently converted to a wide range of industrial-
raw materials with renewable feedstock e.g. lignocellulosic biomass has needed chemicals. Typically, lignin from different types of biomass


Corresponding author.
E-mail address: navadol@jgsee.kmutt.ac.th (N. Laosiripojana).

https://doi.org/10.1016/j.cej.2018.09.048
Received 27 September 2017; Received in revised form 27 August 2018; Accepted 6 September 2018
Available online 07 September 2018
1385-8947/ © 2018 Elsevier B.V. All rights reserved.
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

always comprises of three aromatic alcohols, including p-coumaryl al- were also employed for qualitative and quantitative analysis of the
cohol, coniferyl alcohol, and sinapyl alcohol, linked together by con- isolated lignin compared to the commercial grade organosolv lignin.
densed and ether linkages such as β-O-4, β-β, β-5, 4-O-5, and dihy-
drobenzofuran [2]. The proportion of these aromatic alcohols depends 2. Experimental
considerably from species to species, particularly with regard to the
type and quantity of linkages in polymer and the number of methoxy 2.1. Chemicals and raw materials
groups present on the aromatic rings.
Prior to the conversion of lignin to value-added products, extraction The standard of aromatic monomers (guaiacol, phenol, 2,6-di-
or fractionation of lignocellulosic components is an important initial methoxyphenol, p-cresol, vanillin, 4-ethylphenol, and 4-ethylguaiacol)
step where each biopolymers are separated and utilized for maximized were purchased from Sigma-Aldrich and Tokyo chemical industry Co.
values by conversion to a spectrum of biofuels, commodity and speci- Ltd. For catalyst synthesis, D-Glucose anhydrous was of analytical grade
alty chemicals. Among the biomass fractionation methods, organosolv and was purchased from Ajax Finechem Pty Ltd (Thailand). Sulfuric
process is versatile to types of biomass and offers high selectivity and acid was of analytical grade and was purchased from Fluka (Singapore).
purity to lignocellulose components. Several organic solvents including Cellulose was purchased from Wako Pure Chemical Company (Osaka,
alcohols (e.g. methanol and ethanol), esters, and ketones have been Japan), while organosolv lignin was also supplied from Chemical Point
studied for organosolv fractionation of different lignocellulosic mate- UG (Germany). All chemicals and reagents were used as received.
rials [3,4]. Recently, the clean fractionation (CF) process using ternary Sugarcane bagasse was obtained from PTT Global Chemicals PCL,
mixture of solvents including short-chain alcohol, methylisobutylk- Rayong, Thailand. It was dried at 70 °C for 24 h and cut by Retsch
etone (MIBK), and water in the presence of sulfuric acid as promoter ZM200 to be the average particles size of 0.5–0.85 mm in diameter.
has been reported as an effective approach for separation of lig- According to the standard NREL analysis [23], the biomass contained
nocellulose components [5,6]. The extracted lignin from this fractio- 38.3 wt% cellulose, 20.7 wt% hemicelluloses, 24.6 wt% lignin, and
nation process is called organosolv lignin, which has been known as a 4.3 wt% ash.
promising lignin source for depolymerization to bio-based valuable
chemicals because its structure is highly pure, unaltered, and sulfur-free 2.2. Lignin isolation procedure
which is greatly benefit for utilizing in several catalytic applications.
Regarding the current lignin utilization approaches, most lignin has Isolated lignin was obtained by our developed fractionation proce-
been used mainly as a low-grade fuel for heat and power generation in dure. In details, 10 g of bagasse was mixed with 100 mL of the single-
pulp and paper industries. Only a few proportion of lignin has been phase solvent mixture containing 70–90% by volume of solvent mixture
employed as binders, dispersant or emulsifier, carbon fiber, phenolic [water/ethanol/ethyl acetate (50%:25%:25% as the starting ratio)]
resins, and bio-fuels [7,8]. Recently, chemical reaction approaches for with 10–30% by volume of formic acid. Nitrogen was purged into the
lignin conversion to low molecular weight phenolic compounds have reactor and adjusted the initial pressure to 20 bar. The system was
been widely reported including hydrolysis, hydrogenolysis (or hydro- heated to the desirable temperature (160 °C) for a specified residence
cracking), pyrolysis, solvolysis, and oxidation [9]. Among these pro- time (30 min). Then, the reactor was quenched in a water bath to stop
cesses, lignin depolymerization is mostly dominated by thermochemical the reaction. The solid cellulose-enriched fraction was separated by
transformation which requires harsh conditions with and/or without filtration and washed with the ethyl acetate and water (1:2 v/v) mix-
catalysts [10]. Additional of suitable catalyst to promote lignin depo- ture. The liquid fraction was combined with the rinsate and placed into
lymerization is an attractive strategy to achieve high yields and high a separatory funnel. Water was added to the aqueous-organic fraction
selectivity of desired phenolic products. It is known that strong acids until phase separation was obtained. The mixture was stirred and then
(i.e. H2SO4) can efficiently catalyze the depolymerization reactions. placed at room temperature for 20 min for complete phase separation.
Nevertheless, the main concern for the use of strong mineral acids are The aqueous phase containing hemicelluloses and soluble products was
the corrosive nature of acids, formation of possible unwanted by-pro- recovered. The separated organic phase was dried at 105 °C to obtain
ducts, and the requirement for an acid-containing waste treatment isolated lignin.
[11,12]. Alternatively, the use of heterogeneous acid catalysts would The sample was characterized by several techniques. Fourier
efficiently eliminate or minimize these drawbacks. Previously, several transform infrared spectroscopy (FT-IR) analysis was performed using
solid acid catalysts such as zeolites (H-USY, H-MOR, H-BEA, and H- Perkin-Elmer System 2000 and the samples were prepared by using KBr
ZSM-5) and amorphous silica-alumina (SiO2-Al2O3) and carbonaceous pellet method. The measurement resolution was set at 4 cm−1 with 64
solid acids (CSA) have been widely reported for upgrading bio-oil via scans in the range from 4000 to 400 cm−1. The purity of lignin was
liquefaction in hydrocracking process [13–16]. Among them, CSA have determined by Klason lignin technique, from which 0.3 of lignin sample
several favorable characteristics such as high catalytic activity and se- was added with 3 mL of 72% H2SO4 in pressure tubes. The tubes were
lectivity, long catalyst life, and ease in recovery. CSA catalysts with then immersed in a water bath at 30 °C for 1–2 h. The hydrolyzed lignin
SO3H, COOH and OH groups were also reported to active for the hy- samples were then diluted to 4% through the addition of water and
drolysis of cellulose and its derived polysaccharides [17–22]. Further- autoclaved at 121 °C for 1 h. The samples were filtered with porcelain
more, CSA can be easily synthesized by the low temperature hydro- crucibles to remove any solids and the liquid fraction was analyzed by
thermal of renewable hydrocarbon feedstocks e.g. cellulose, lignin, high performance liquid chromatography (HPLC) for glucose, xylose
sugars. The purpose of this study is to examine catalytic performance of and arabinose. The solid fraction comprised of acid-insoluble lignin and
CSA for lignin depolymerization into phenolic monomers. The pre- the acid-insoluble ash was placed in a muffle furnace at 575 °C for 3 h.
paration method of CSA and the precursor type were intensively stu- After cooling, the remaining acid-insoluble ash was weighted and cal-
died. In details, hydrothermal synthesis was applied while the types of culated as a percentage of the original dry weight of sample [24]. The
precursor for CSA included glucose, cellulose and lignin. For the de- elemental composition of organosolv lignin was estimated by an Ele-
polymerization test, the effects of catalyst loading, solvent system, mental Analyzer CHN-S 628 (LECO Corp.). Lignin samples were dried at
temperature and hydrogen donor on the product yield were performed 60 °C in vacuum evaporator with 20 mbar to remove moisture. Then,
over lignin isolated from bagasse by organosolv fractionation process. 0.1 g of lignin sample was encapsulated in the container to determine
The liquid and solid products i.e. oil, residual lignin and char/un- carbon, hydrogen, and nitrogen in sample. For sulfur analysis, 0.2 g of
converted lignin were intensively analyzed. In addition, different lignin sample was placed into ceramic boat furnace. This was in-
characterization techniques including Klason lignin determination, cineration at 1350 °C using sulfur IR cells detect the amount of sulfur.
elemental analysis, FT-IR, and Gel permeation chromatography (GPC) Lastly, the molecular weight of isolated lignin was determined by GPC,

462
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

from which the sample was prepared in eluent of Tetrahydrofuran 280 °C and kept at this temperature for 10 min. The final temperature
(THF) at a concentration of 0.2 mg/ml. Prior to analysis, the slurry was was held at 300 °C for 10 min at the same heating rate. The main
filtered through a 0.45 µm syringe filter to remove any solid residues. phenolic monomers showing a relatively high concentration in the li-
HPLC was used with a Shodex Asahipak GS-320 HQ column and a quid product were further quantitated by using external standard
Water 2487 UV detector (280 nm). The mobile phase was THF with the method on GC-FID (Shimadzu GC-2014, Japan). The quantitative ana-
flow rate of 0.5 mL/min. The temperature of column was maintained at lysis was carried out by a GC-FID (Shimadzu GC-2014, Japan) equipped
30 °C. Sodium polystyrene sulphonate standards with molecular with a PetrocolTM DH 50.2 fused silica capillary column
weights 1530, 4950, 16600, and 34700 g/mol were used to prepare a (50 m × 0.2 mm × 0.5 μm; Supelco, USA) with flame ionization de-
standard calibration curve. Lignin weight average molecular weight tector (FID). The column was initially kept at 50 °C for 5 min, then was
(Mw) and number average molecular weight (Mn) were calculated after heated at a rate of 10 °C·min−1 to 120 °C, and maintained for 5 min.
comparison with standards. Then, it was heated up to 280 °C at 10 °C·min−1 and kept at this tem-
perature for 15 min. The phenolic products were quantified using the
2.3. Catalyst preparation and characterization calibration curves of a number of lignin depolymerization products (i.e.
guaiacol, phenol, 2,6-dimethoxyphenol, p-cresol, vanillin, 4-ethyl-
CSA was initially prepared in stainless steel closed batch reactor by phenol, and 4-ethylguaiacol).
hydrothermally carbonizing of 1 g carbon precursor (i.e. glucose, cel- The yields of phenolic monomers, residual lignin (RL), liquid acid
lulose and lignin) in 5 mL of deionized water. The closed batch reactor soluble, and char/unconverted lignin, and lignin conversion were cal-
containing the mixture was heated with an electric heater to the desired culated using the following equations:
temperature, 220 °C. After a holding period of 3 h, the reactor was
Phenolic monomers yield (\%)
cooled to room temperature by submerging it into a water bath. The
solid product was separated from the liquid carbonization portion by weight of phenolic monomers (g) ⎞
= 100 ⎜⎛ ⎟
filtering and then washed with deionized water until the pH of washed ⎝ weight of initial lignin (g) ⎠
water was neutral. After drying overnight at 110 °C, the CSA was pre-
pared by heating 5 g of the hydrothermal solid in 50 mL of concentrated weight of RL in acid insoluble (g) ⎞
Residual lignin yield (\%) = 100 ⎛⎜ ⎟
H2SO4 (> 96%) using a 3-neck round bottom flask at 150 °C for 15 h. ⎝ weight of initial lignin (g) ⎠
The obtained CSA was washed with deionized water (1000 mL) before
being separated with a vacuum filter. The solid was repeatedly washed Liquid acid soluble yield (\%)
with boiling distilled water until no change in water pH was observed weight of liquid fraction (g) − weight of RL (g) ⎞
and eventually dried at 110 °C. It is noted that the glucose, cellulose and = 100 ⎛⎜ ⎟

⎝ weight of initial lignin (g) ⎠


lignin derived CSAs are called CSAglucose, CSAcellulose and CSAlignin, re-
spectively. The elemental compositions (C, H, and N) of CSAs were Char/unconverted lignin yield (\%)
determined by a CHN analyzer (LECO CHN628 elemental analyzer).
weight of solid fraction (g)− weight of catalyst (g) ⎞
The oxygen (O) was calculated by sub-traction of C, H, and N content = 100 ⎜⎛ ⎟

⎝ weight of initial lignin (g) ⎠


from 100%. The specific area, pore volume and pore size diameter of
samples were determined by N2 physisorption technique using the Lignin converion (\%)
Brunauer-Emmet-Teller (BET) method with a Belsorp-mini (BEL Japan,
weight of initial lignin (g)− weight of solid fraction (g)− weight of RL (g) ⎞
Tokyo, Japan); the samples were pre-treated to remove moisture at = 100 ⎜⎛ ⎟

⎝ weight of initial lignin (g) ⎠


150 °C for 3 h prior to measurement. The total acidity of the samples
was analyzed by temperature-programmed desorption (TPD) with am-
monia. 3. Results

2.4. Lignin depolymerization test 3.1. Isolated lignin characterizations

The mixture of 1.75% (w/v) lignin in selected solvents with and Prior to the conversion test, the characteristics of isolated lignin in
without the presence of catalyst was filled in a stainless steel tubular the present work were examined and compared to commercial orga-
reactor (1/2 in. O.D. and 10 cm, length) before heating to the desired nosolv lignin by several techniques including FT-IR, Klason lignin,
temperatures. Various acid catalysts used in this study were loaded at CHNS and GPC analyses. FT-IR of lignin was employed in order to
1% (w/w) into the reactor. Initially, the reactor was purged with ni- identify the presence of chemical functional groups. As presented in
trogen and shake in vertical dimension at 40 rpm. After the reaction, Fig. 2, it can be seen that isolated lignin and commercial organosolv
the reactor was quenched in a water bath to stop the reaction im- lignin samples contain a broad band at 3437 cm−1, attributed to the
mediately. The liquid fraction was separated from solid fraction (in- hydroxyl (OeH) groups in phenolic and aliphatic structures, and the
clude char, unconverted lignin, and catalyst) by centrifugation at bands centered around 2934 and 2839 cm−1 predominantly arising
5000 rpm for 10–15 min described in Fig. 1. Subsequently, the liquid from C-H stretching in aromatic methoxyl groups and in methyl and
fraction was acidified to a pH of 2.0 with sulfuric acid (1 M). In this methylene groups of side chains. The bands at 1609, 1511, and
step, residual lignin (RL) fraction was precipitated out as a solid form 1422 cm−1 were absorbed due to aromatic skeleton vibrations. The
and was dried overnight at 60 °C. Apart from that, the liquid fraction characteristics of adsorption of lignin at 834 cm−1 were related to the
(containing phenolic monomers) was filtered through a 0.45 μm syringe out of plane CeH vibration in guaiacyl (G) units. In addition, the band
filter and further characterized by GC-MS and GC-FID, respectively. The at 1115 cm−1 assigned to the characteristic absorption of p-hydro-
yields of the all dried fractions (liquid acid soluble fraction, residual xypheny (H) units. The peak at 1710 cm−1 assigned to C]O stretching
lignin fraction, and char/unconverted lignin fraction) were measured vibration.
gravimetrically and the yields were calculated in comparison to initial To determine the chemical compositions as well as impurities in-
weight of lignin. cluding insoluble lignin (AIL), acid soluble lignin (ASL), sugars, and ash
Phenolic products were identified by GC-MS. The GC-MS system content, the samples was analyzed by Klason lignin technique. As
was equipped with an Agilent CP-Sil 5 CB capillary column shown in Table 1, the AIL proportion of isolated lignin in the present
(60 m × 0.32 mm × 1.00 μm; Agilent, USA). The temperature program study (92.3%) was relatively in the same range as that of commercial
started at 50 °C to 120 °C for 5 min. Then, the reaction was heated up to organosolv lignin (91.7%). In addition, the compositions of C6 sugars

463
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

Fig. 1. Diagram of product separation after depolymerization reactions.

that the isolated lignin composes of 61–63% C and 5–6% H with sig-
nificantly low sulphur content (< 0.4%). The general formula is CxHyOz
with x of 7.9–9.0; y of 10–10.62; and z of 2.89–4.5, which correlates
well those reported in the literature [10].
GPC analysis was also applied to identify the molecular weight of
isolated lignin. As shown in Fig. 3 and Table 3, the isolated lignin
contained Mn of 1719 gmol−1, Mw of 3144 gmol−1 with D (Mw/Mn) of
1.8, while the commercial organosolv lignin exhibited Mn of
1133 gmol−1, Mw of 1424 gmol−1 with D (Mw/Mn) of 1.3. These
characterization results implied that the isolated lignin in this study has
higher heterogeneous distribution of fractions with different chain
lengths than the commercial organosolv lignin.

3.2. Catalytic activities toward lignin depolymerization

Lignin depolymerization reaction was initially carried out at 300 °C


for 1 h under atmospheric pressure in MIBK with and without the
presence of catalysts. The phenolic yields from the reaction are illu-
Fig. 2. FT-IR spectra of commercial organosolv lignin and bagasse lignin strated in Fig. 4. Although the conversion of lignin was always higher
samples. than 90%, it was found that without the presence of catalyst only 11.6%
of total phenolic yield was observed. Importantly, the additional of
catalyst clearly promote the phenolic yield production, from which the
and ash contents of isolated lignin were much lower than that of reaction with H2SO4 gained higher activity than that with H3PO4 and
commercial organosolv lignin, whereas the content of C5 sugars espe- HCl. Interestingly, the reaction with CSAs can produce higher phenolic
cially arabinose was relatively higher. The organic and inorganic ele- yield than that with homogeneous acids. Among the CSA catalysts,
ments (C, H, N, and S) as well as the molecular weight of C9-formulae in CSAlignin provided the highest yield of phenolic monomers of 17.5%
isolated lignin were given in Table 2. CHNS elemental analysis revealed with high selective productions of guaiacol and 4-ethylphenol. This

464
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

Table 1
Comparison of lignin composition in this work and other recent publications.
Lignin samplea Klason lignin (%) Sugar (%) Ash (%) Moisture (%) Ref.

Glucose Xylose Arabinose

COL 91.7 0.92 0.12 0.12 0.86 < 0.1 This study
IL 92.3 0.16 0.52 0.27 0.22 < 0.1 This study
FL 79.0 1.00 0.75 0.14 2.71 N/A Ercodia et al. [44]
AL 69.1 2.61 2.26 0.14 2.57 N/A Ercodia et al. [44]
ACL ∼90 ————————Sugars: 0.50———————— < 0.1 <4 Bozell et al. [5]

a
COL: commercial organosolv lignin, IL: isolated lignin, FL: formasolv lignin, AL: acetosolv lignin, ACL: alcell lignin.

Table 2 higher porosity than CSAGlucose. The trends of these physical and che-
Elemental composition of lignin samples. mical characteristics are in good agreement with the catalytic perfor-
Lignin sample Elemental analysis (wt%) mance toward lignin depolymerization reaction.
It should be noted that the depolymerization reaction over CSA
C H O N S synthesized from commercial grade organosolv lignin (CSACOL) was
also tested for comparison. As shown in Fig. 4, the reaction perfor-
COL 61.43 5.61 30.99 0.81 0.30
IL 62.41 5.91 30.67 0.56 0.19
mance catalyzed by CSACOL was slightly less than that catalyzed by
CSAlignin. This could be due to the higher molecular weight of sugarcane
bagasse lignin compared to commercial grade organosolv lignin, which
leads to the better catalyst surface properties and acidity value (as also
presented in Table 4). The catalyst reusability was then carried out as
shown in Fig. 6(a)–(c). The post-reaction solid sample was rinsed with
acetone for several times to remove lignin fraction. The spent catalyst
and char were then tested without separation. After 5 consecutive cy-
cles, the reaction with CSAlignin also provided the lowest deactivation
percentage compared to CSACellulose and CSAGlucose. From the post-re-
action TPD, the acidity of CSAlignin decreased by 8.14%, whereas that of
CSAGlucose and CSACellulose decreased by 15.7% and 21.5%, respectively.
These results indicated the relatively lower acid leaching from CSAlignin,
which would be greatly benefited for the long-term reaction operation.
The elemental analysis of CSAlignin after 5 consecutive cycles was also
analyzed. It was found that the carbon content increased from C:H:O of
75.39:7.48:16.89 to 79.28:7.35:13.19, which could be due to the char
deposition on the surface of the spent catalyst.
The effect of catalyst loading was also studied as shown in Fig. 7.
Increasing CSAlignin loading up to 10 wt% at temperature of 300 °C for
1 h resulted in a maximum increase in the yield of phenolic monomers.
By loading more than 10%, the liquid yield in acid soluble fraction
Fig. 3. GPC curves of commercial organosolv lignin and bagasse lignin.
trends to decrease while residual lignin and solid char proportions
oppositely increased significantly. This result corresponded with Klein
Table 3 et al. [26] who reported that 10 wt% catalyst loading for depolymer-
Molecular weight of lignin samples. ization of lignin from woody biomass gave the highest yield of lignin-
Molecular weight (g/mol) derived products. It is noted that the increases of guaiacol and 4-
ethylphenol in the acid soluble fraction was accompanied by the de-
Lignin sample Mw Mn D (Mw/Mn) creases of phenol and p-cresol. This implied that an increased catalyst
loading efficiently depolymerized and suppressed recondensation be-
COL 1424 1133 1.257
IL 3144 1719 1.828 tween lignin fragments, but overwhelming catalyst amount led to fur-
ther decomposition to the side reactions.

may be due to unique characteristics of catalysts (shape selectivity,


3.3. Effect of solvent type
acidity and strength of the acid sites, surface area, pore diameter, pore
volume) as well as the complexity of the reactions in terms of which
As the next step, the reactions using different types of organic sol-
bonds in lignin were interacting with catalytically active sites under the
vents including methanol, acetone and ethyl acetate were tested and
reaction conditions [25]. According to the TPD and BET measurements
compared to the use of MIBK. It is noted that the reaction was carried
(Table 4), among all CSAs, CSAlignin showed the highest acidity
out at 300 °C for 1 h in the presence of 10 wt% CSAlignin. As shown in
(14.13 mmol g−1 compared to 12.57 mmol g−1, and 10.69 mmol g−1
Fig. 8, the reaction with acetone provided the highest lignin conversion
for CSAGlucose and CSACellulose respectively) as well as specific surface
with relatively low proportions of solid char and residual lignin gen-
area (14.7 m2 g−1 compared to 4.2 m2 g−1 and 8.8 m2 g−1 for
erated from the reactions, whereas the use of ethyl acetate gained the
CSAGlucose and CSACellulose respectively). In addition, a scanning elec-
lowest lignin conversion. These results implied that acetone can effi-
tron microscope (SEM, JEOL, JSM-6610 LV) connected with an Energy
ciently break the functional linkages in lignin compounds, whereas
Dispersive X-Ray Spectrometer (EDS) analyzer was also used for the
ethyl acetate shows limit ability for breaking down lignin to monomeric
morphology observation. As shown in Fig. 5, SEM micrograph of
phenols. Nevertheless, in term of total phenolic production yield, MIBK
CSAlignin revealed a homogeneous surface characteristic with relatively
can comparatively gain the highest yield of total phenolic products.

465
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

Fig. 4. Effect of catalyst type on lignin conversion and products yield from depolymerization of bagasse lignin at 300 °C in MIBK for 1 h.

Table 4
Surface area and acid properties of CSA catalysts.
Catalysts BET Surface Pore volume Pore size Total acidity
Areaa (m2 g−1) (cm3 g−1) (nm) (mmol g−1)

CSAGlucose 4.2 0.005 10.39 12.57


CSACellulose 8.8 0.006 7.92 10.69
CSALignin 14.7 0.011 4.28 14.13
CSACOL 11.3 0.009 5.73 13.06

This could be due to the fact that MIBK solvent is not inert under the
depolymerization; it undergoes condensation reactions. Therefore, it
can be clarified from this study that acetone is appropriate for lignin
decomposition, whereas the most suitable solvent for maximizing the
phenolic compound production from lignin depolymerization is MIBK.
In addition, it is clear that the reaction medium strongly affects the
degree of lignin conversion as well as the distribution of each phenolic
compound in lignin depolymerization reactions.

3.4. Effects of reaction temperature and time

Lastly, the effects of reaction temperature and time on product yield


of phenolic monomers in the presence of CSAlignin using MIBK as solvent
was investigated in order to determine the conditions where the yield of
phenolic monomers can be maximized. As illustrated in Fig. 9, the re-
action temperature showed strong impact on lignin conversion and
product yield. It was found that the lignin conversion considerably in-
creased at higher temperature, while the proportions of char and re-
sidual lignin contents significantly decreased; nevertheless, above
350 °C, the char contents oppositely increased, which could be due to
the formation of organic acids in the side chains of phenolic propanoid
units [27–29]. In term of liquid product, the amount of liquid acid
soluble greatly increased from 9.1 to 24.0% with increasing the tem-
perature from 250 to 350 °C, which could be due to the fact that in-
creasing the reaction temperature likely promotes the rates of depoly-
merization reaction especially the cleavage of β-O-4 ether bonds in
lignin to monomeric phenols as well as the increased lignin solubility.
This result is consistent with Ren et al. [30], in which the yield of
phenol and alkyl phenols elevated with increasing temperature at Fig. 5. Surface morphologies of CSAGlucose (a) and CSAlignin (b) analyzed by
SEM.
350 °C due to occurred secondary reactions via demethylation and de-
methoxylation. Above 350 °C, the yield of phenolic monomers tended to

466
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

Fig. 6. Catalyst reusability test for (a) CSALignin, (b) CSACellulose, and (c) CSAGlucose.

decrease, which could be due to the decomposition of these compounds time, the positive effect became less pronounce. Therefore, it can be
to several smaller molecule and gaseous compounds. revealed from this study that the optimizing conditions for depoly-
The reaction time also affected the lignin conversion and product merizing isolated lignin from organosolv fractionation of bagasse were
yield (Fig. 10), from which the conversion and phenolic yield increased using MIBK as solvent in the presence of CSAlignin as catalyst at 350 °C
with increasing the reaction time from 0.5 h to 3 h. At longer reaction for 3 h, from which as high as 32.8% phenolic monomers can be gained.

467
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

Fig. 7. Effect of catalyst loading on lignin conversion and product distribution from depolymerization of bagasse lignin at 300 °C in MIBK for 1 h in the presence of
CSAlignin.

4. Discussion from the hydrothermal process as carbon-based acid catalyst since it


has several favorable characteristics compared to activated carbon
The selective depolymerization of lignin to phenolics is one of the when functionalized with acid promoters (e.g. SO3H) such as longer
current challenging research areas. Due to the heterogeneity of lignin catalyst life, lower deactivation after recycling and ease in recovery.
structures and the diversity of ether and condensed linkages, the major Furthermore, CSA can be easily synthesized at moderate temperature
difficulties of this reaction include char formation, repolymerization of (180–250 °C) under self-generated pressure. The major bottleneck of
products, low product yield and selectivity [31]. It is generally known activated carbon is the use of high carbonization/pyrolysis tempera-
that the depolymerization and repolymerization reactions compete tures (400–800 °C). The energy requirement becomes even more con-
with each other during the lignin conversion [32–34]. To maximize the suming, particularly when additional heat is needed to remove water
lignin depolymerization while restrain lignin repolymerization, the from biomass with high moisture content. It is noted that, for com-
operating conditions need to be optimized. Recently, several studies parison, the reaction tests over acid-functionalized activated carbon
have been attempting to investigate the selections of catalyst, solvent, using commercial grade activated carbon (total surface area of
and reaction condition where lignin can be effectively cleaved to reduce 1343 m2 g−1 and average pore size of 0.67 nm) supplied from Carbo-
the multitude of products and increase product yield as well as se- karn (Thailand) was also performed. It found that although the con-
lectivity. Carbon-based acid catalysts have been reported to possess version of lignin was slightly higher than the use of CSA, the total
acid sites that can efficiently catalyze several reactions including the phenolic yield from the reaction is relatively lower (15.6% compared to
upgrading bio-oil via liquefaction in hydrocracking process and hy- 17.5%). This could be due to the too high acidity of acid-functionalized
drolysis of cellulose and its derived polysaccharides. The sources of activated carbon, which leads to the unwanted side reactions of phe-
carbon can come from the carbonization and/or hydrothermal of hy- nolic compounds. Moreover, after 5 consecutive cycles, the reaction
drocarbon compounds. In the present work, we decided to use the CSA with acid-functionalized activated carbon provided relatively higher

Fig. 8. Effect of solvent system on lignin conversion and product distribution from depolymerization of bagasse lignin at 300 °C for 1 h in the presence of CSAlignin.

468
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

Fig. 9. Effect of reaction temperature on product distribution and lignin conversion from depolymerization of bagasse lignin in MIBK for 1 h in the presence of
CSAlignin.

Fig. 10. Effect of reaction time on product distribution and lignin conversion from depolymerization of bagasse lignin in MIBK at 350 °C in the presence of CSAlignin.

deactivation percentage compared to CSA (15.2% compared to 8.1% Therefore, the suitable amount of acid catalyst loading needs to be
deactivation) due to the high acid leaching percentage. optimized to suppress or prevent recondensation reaction. From our
For the novelty of this study, we found that physical and chemical experiment results, we found that at 10 wt% acid loading gave the
characteristics of CSA strongly depends on the source of carbon pre- highest yield of aromatic products. It should also be noted that the
cursors, from which CSA synthesized from lignin contained sig- presence of phenol as product from the reaction could also suppress
nificantly higher acidity and porosity compared to that synthesized char formation as reported by Toledano et al. [36].
from sugar and cellulose. In term of reaction, CSAlignin also showed the Apart from the study of catalyst, the type of solvent also strongly
best performance in terms of phenolic monomer yield and potential affects the performance of lignin depolymerization. The use of good
reusability. In addition, the characteristics of lignin (i.e. molecular solvent could reduce char formation, and minimize repolymerization
weight, purity) also affects the properties of CSA. Our study also in- reactions. According to our previous studies [37,38], lignin solubility
dicated that the presence of CSAlignin led to a significant formation of was tested in a wide range of organic solvents including acetone, me-
aromatic compounds e.g. guaiacol, which is due to the increase in de- thanol, ethanol, THF, DMSO, MIBK, ethyl acetate, methanol, and
gree of acid-catalyzed cleavage of ether linkages. Similar result was acetone. We found that the degree of lignin solubility in solvent
reported by Shao et al. [35], who studied the lignin depolymerization strongly affects the lignin depolymerization performance. Among these
using formic acid as catalyst. They found that bio-oil yield increased solvents, MIBK, ethyl acetate, methanol, and acetone are the promising
dramatically, accompanied by a decrease in residue with increasing solvents in terms of lignin solubility and depolymerization. Further-
catalyst loading. However, further increase of catalyst loading resulted more, these solvents can be easily reused because they are stable under
in a decrease of bio-oil yield and led to an increase in residue yield since the reaction conditions in this study. Importantly, in term of the prac-
the recondensation reaction was preferred at a higher acid loading. tical application, these solvents are cost effective and commonly used in

469
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

the industries. From the literature, the depolymerization of lignin in the versicolor, Bioresources 6 (2011) 1273–1287.
presence of several solvents i.e. water, methanol, dioxane, tetra- [2] X. Huang, T.I. Korányi, M.D. Boot, E.J.M. Hensen, Catalytic depolymerization of
lignin in supercritical ethanol, ChemSusChem 7 (2014) 2276–2288.
hydrofuran, ethanol was also compared [39]. It was found that the use [3] J. Wildschut, A.T. Smit, J.H. Reith, W.J.J. Huijgen, Ethanol-based organosolv
of 1:1 (v/v) of methanol–water shows the best performance among the fractionation of wheat straw for the production of lignin and enzymatically diges-
uses of water, methanol, dioxane, tetrahydrofuran, ethanol, from which tible cellulose, Bioresour. Technol. 135 (2013) 58–66.
[4] C.G. Boeriu, F.I. Fitigau, R.J.A. Gosselink, A.E. Frissen, J. Stoutjesdijk, F. Peter,
17.9% yield of monomeric compounds was achieved at 150 °C with Fractionation of five technical lignins by selective extraction in green solvents and
60 min of reaction time. Commonly, alcohols are good choice for lignin characterization of isolated fractions, Ind Crops Prod. 62 (2014) 481–490.
depolymerization since they act as hydrogen-donor solvents [34,40]. [5] J.J. Bozell, S.K. Black, M. Myers, D. Cahill, W.P. Miller, S. Park, Solvent fractio-
nation of renewable woody feedstocks: Organosolv generation of biorefinary pro-
Wu et al. [41] reported the depolymerization of lignin catalyzed by cess steams for the production of biobased chemicals, Biomass Bioenergy 35 (2011)
SiO2-Al2O3 using methanol, ethanol, isopropanol, and THF as solvent. 4197–4208.
Among these solvents, ethanol was the best solvent in terms of phenolic [6] G. Brudecki, I. Cybulska, K. Rosentrater, Integration of extrusion and clean frac-
tionation processes as a pre-treatment technology for prairie cordgrass, Bioresour.
yield achievement and char suppression. Chen et al. [42] also reported
Technol. 135 (2013) 672–682.
that the use of ethanol can reduce char formation and provided high [7] R.J.A. Gosselink, E. de Jong, B. Guran, A. Abächerli, Co-ordination network for
yield of monomers from the lignin depolymerization reaction. In con- lignin standardisation, production and applications adapted to market requirements
trast, Wang and Rinaldi [43] indicated that the reaction using methanol (EUROLIGNIN), Ind. Crops Prod. 20 (2004) 121–129.
[8] J. Lora, Chapter 10 – Industrial commercial lignins: Sources, properties and ap-
as solvent showed better performance in term of phenol production plications A2 – Belgacem, in: A. Gandini (Ed.), Mohamed Naceur monomers,
than the uses of ethanol and butanol. Furthermore, Ercodia et al. [44] polymers and composites from renewable resources, Elsevier, Amsterdam, 2008,
reported that acetone is the better solvent (compared to methanol and pp. 225–241.
[9] M.P. Pandey, C.S. Kim, Lignin depolymerization and conversion: a review of ther-
ethanol) for depolymerizing lignin to phenolic compounds. Sameni mochemical methods, Chem. Eng. Technol. 34 (2011) 29–41.
et al. [45] and Gosselink et al. [33] revealed that organosolv lignin [10] J. Zakzeski, P.C. Bruijnincx, A.L. Jongerius, B.M. Weckhuysen, The catalytic va-
effectively dissolve in acetone compared to alcohols and could lead to lorization of lignin for the production of renewable chemicals, Chem. Rev. 110
(2010) 3552–3599.
the good depolymerization performance. For the role of acetone in the [11] K. Karimi, S. Kheradmandinia, M.J. Taherzadeh, Conversion of rice straw to sugars
depolymerization of lignin, Hwang et al. [46] studied the degradation by dilute-acid hydrolysis, Biomass Bioenergy 30 (2006) 247–253.
of polystyrene in supercritical acetone. They proposed that polystyrene [12] P. Lenihan, A. Orozco, E. O’Neill, M.N.M. Ahmad, D.W. Rooney, G.M. Walker,
Dilute acid hydrolysis of lignocellulosic biomass, Chem. Eng. J. 156 (2010)
can be degraded to monomer and dimers by partial cleavage of ether 395–403.
linkage and carbon-carbon linkage under acetone media. For MIBK, this [13] J.A. Bootsma, B.H. Shanks, Cellobiose hydrolysis using organic-inorganic hybrid
solvent is one of the well-known organic solvents for biomass pre- mesoporous silica catalysts, Appl. Catal. A: Gen. 327 (2007) 44–51.
[14] H. Kobayashi, T. Komanoya, K. Hara, A. Fukuoka, Water-tolerant mesoporous-
treatment and fractionation. Until now, the role of MIBK for the de-
carbon-supported ruthenium catalysts for the hydrolysis of cellulose to glucose,
polymerization of lignin to low molecular weight products is not clearly ChemSusChem 3 (2010) 440–443.
understood. Teng et al. [47] reported that MIBK is one of the preferred [15] A. Takagaki, C. Tagusagawa, K. Domen, Glucose production from saccharides using
candidates for the biomass conversion process due to its stability to- layered transition metal oxide and exfoliated nanosheets as a water tolerant solid
acid catalyst, Chem. Commun. 42 (2008) 536–545.
ward acid, medium polarity, low toxicity, medium boiling point, poor [16] L. Zhou, Z. Liu, M. Shi, S. Du, Y. Su, X. Yang, J. Xu, Sulfonated hierarchical H-USY
miscibility with water, and good extraction ability for polar products. zeolite for efficient hydrolysis of hemicellulose/cellulose, Carbohydr. Polym. 98
From these observations, it can be revealed that the effectiveness of (2013) 146–151.
[17] K. Fukuhara, K. Nakajima, M. Kitano, H. Kato, S. Hayashi, M. Hara, Structure and
solvent to breakdown lignin depends on several parameters e.g. oper- catalysis of cellulose-derived amorphous carbon bearing SO3H groups,
ating conditions, source and type of lignin, and lignin solubility. ChemSusChem 4 (2011) 778–784.
[18] M. Kitano, D. Yamaguchi, S. Suganuma, K. Nakajima, H. Kato, S. Hayashi, M. Hara,
Adsorption enhanced hydrolysis of beta-1,4-glucan on graphene-based amorphous
5. Conclusions carbon bearing SO3H COOH, and OH groups, Langmuir 25 (2009) 5068–5075.
[19] A. Onda, T. Ochi, K. Yanagisawa, Selective hydrolysis of cellulose into glucose over
The catalytic activities of several homogeneous and heterogeneous solid acid catalysts, Green Chem. 10 (2008) 1033–1037.
[20] J.F. Pang, A.Q. Wang, M.Y. Zheng, T. Zhang, Hydrolysis of cellulose into glucose
acid catalysts toward the depolymerization of organosolv lignin to
over carbons sulfonated at elevated temperatures, Chem. Commun. 46 (2010)
phenolic monomers were comparatively studied under several oper- 6935–6937.
ating conditions. Among all catalysts studied, CSAlignin showed the best [21] S. Suganuma, K. Nakajima, M. Kitano, D. Yamaguchi, H. Kato, S. Hayashi,
H. Michikazu, Hydrolysis of cellulose by amorphous carbon bearing SO3H, COOH,
performance in terms of phenolic monomer yield and potential reusa-
and OH groups, J. Am. Chem. Soc. 130 (2008) 12787–12793.
bility. Increasing the catalyst dosage to 10 wt% positively promoted the [22] S. Suganuma, K. Nakajima, M. Kitano, D. Yamaguchi, H. Kato, S. Hayashi, Synthesis
lignin conversion and the total phenolic monomers (mainly 4-ethyl- and acid catalysis of cellulose-derived carbon-based solid acid, Solid State Sci. 12
phenol and guaiacol) yield. Furthermore, the study revealed that sol- (2010) 1029–1034.
[23] A. Sluiter, B. Hames, R. Ruiz, C. Scarlata, J. Sluiter, D. Templeton, D. Crocker,
vent system, reaction temperature and time also play an important role Determination of Structural Carbohydrates and Lignin in Biomass, Laboratory
on the lignin conversion and the total phenolic monomer production, Analytical Procedure (LAP), National Renewable Energy Laboratory, 2011 NREL/
from which the reaction using MIBK as solvent at 350 °C for 3 h can gain TP-510-42618.
[24] A. Sluiter, B. Hames, R. Ruiz, C. Scarlata, J. Sluiter, D. Templeton, Determination of
the highest phenolic monomers of 32.8%. Sugars, Byproducts, and Degradation Products in Liquid Fraction Process Samples, ,
Laboratory Analytical Procedure (LAP), National Renewable Energy Laboratory,
Acknowledgements 2008 NREL/TP-510-42623.
[25] A.K. Deepa, P.L. Dhepe, Lignin depolymerization into aromatic monomers over
solid acid catalysts, ACS Catal. 5 (2014) 365–379.
The author (NL) was supported by the Thailand Research Fund [26] I. Klein, B. Saha, M.M. Abu-Omar, Lignin depolymerization over Ni/C catalyst in
(Grant number RTA5980006) and National Science and Technology methanol, a continuation: effect of substrate and catalyst loading, Catal. Sci.
Technol. 5 (2015) 3242–3245.
Development Agency (Grant number P-15-50457). The author (PD) was
[27] V.M. Roberts, V. Stein, T. Reiner, A. Lemonidou, X. Li, J.A. Lercher, Towards
supported by the Thailand Research Fund (Grant number quantitative catalytic lignin depolymerization, Chem. Eur. J. 17 (2011) 5939–5948.
TRG6080013). The author (AS) thank the Thailand Research Fund [28] J. Barbier, N. Charon, N. Dupassieux, A. Loppinet-Serani, L. Mahé, J. Ponthus,
Hydrothermal conversion of lignin compounds A detailed study of fragmentation
(Grant numbers RSA5880047 and IRG5780014) and e-ASIA JRP project
and condensation reaction pathways, Biomass Bioenergy 46 (2012) 479–491.
for financial support. [29] S. Kang, X. Li, J. Fan, J. Chang, Hydrothermal conversion of lignin: a review,
Renew. Sust. Energ. Rev. 27 (2013) 546–558.
References [30] M.-M. Ren, H.-S. Lü, M.-H. Zhang, G.-Q. Wang, Y.-P. Sun, Research progress on the
application of lignin, Polym. Bull. 8 (2012) 44–49.
[31] C. Li, X. Zhao, A. Wang, G.W. Huber, T. Zhang, Catalytic transformation of lignin for
[1] H.M.N. Iqbal, M. Asgher, H.N. Bhatti, Optimization of physical and nutritional the production of chemicals and fuels, Chem. Rev. 115 (2015) 11559–11624.
factors for synthesis of lignin degrading enzymes by a novel strain of Trametes [32] J. Li, G. Henriksson, G. Gellerstedt, Lignin depolymerization/repolymerization and

470
P. Asawaworarit et al. Chemical Engineering Journal 356 (2019) 461–471

its critical role for delignification of aspen wood by steam explosion, Bioresour. Technol. 144 (2016) 181–185.
Technol. 98 (2007) 3061–3068. [40] X. Huang, C. Atay, J. Zhu, S.W.L. Palstra, T.I. Korányi, M.D. Boot, E.J.M. Hensen,
[33] R.J.A. Gosselink, W. Teunissen, J.E.G. van Dam, E. de Jong, G. Gellerstedt, Catalytic depolymerization of lignin and woody biomass in supercritical ethanol:
E.L. Scott, J.P.M. Sanders, Lignin depolymerisation in supercritical carbon dioxide/ Influence of reaction temperature and feedstock, ACS Sustain. Chem. Eng. 5 (2017)
acetone/water fluid for the production of aromatic chemicals, Bioresour. Technol. 10864–10874.
106 (2012) 173–177. [41] Q. Wu, L. Ma, J. Long, R. Shu, Q. Zhang, T. Wang, Y. Xu, Depolymerization of
[34] B. Güvenatam, E.H.J. Heeres, E.A. Pidko, E.J.M. Hensen, Lewis-acid catalyzed de- organosolv lignin over silica-alumina catalysts, Chinese J. Chem. Phys. 29 (2016)
polymerization of Protobind lignin insupercritical water and ethanol, Catal. Today 474–480.
259 (2016) 460–466. [42] P. Chen, Q. Zhang, R. Shu, Y. Xu, L. Ma, T. Wang, Catalytic depolymerization of the
[35] L. Shao, Q. Zhang, T. You, X. Zhang, F. Xu, Microwave-assisted efficient depoly- hydrolyzed lignin over mesoporous catalysts, Bioresour. Technol. 226 (2017)
merization of alkaline lignin in methanol/formic acid media, Bioresour. Technol. 125–131.
264 (2018) 238–243. [43] X. Wang, R. Rinaldi, Solvent effects on the hydrogenolysis of diphenyl ether with
[36] A. Toledano, L. Serrano, J. Labidi, Improving base catalyzed lignin depolymeriza- Raney nickel and their implications for the conversion of lignin, ChemSusChem 5
tion by avoiding lignin repolymerization, Fuel 116 (2014) 617–624. (2012) 1455–1466.
[37] T. Klamrassamee, T. Tana, N. Laosiripojana, L. Moghaddam, Z. Zhang, J. Rencoret, [44] X. Erdocia, R. Prado, J. Fernández-Rodríguez, J. Labidi, Depolymerization of dif-
A. Gutierrez, J.C. del Rio, W.O.S. Doherty, Effects of an alkali-acid purification ferent organosolv lignins in supercritical methanol, ethanol, and acetone to produce
process on the characteristics of eucalyptus lignin fractionated from a MIBK-based phenolic monomers, ACS Sustain. Chem. Eng. 4 (2016) 1373–1380.
organosolv process, RSC Adv. 6 (2016) 92638–92647. [45] J. Sameni, S. Krigstin, M. Sain, Solubility of lignin and acetylated lignin in organic
[38] W. Wanmolee, N. Laosiripojana, P. Daorattanachai, L. Moghaddam, J. Rencoret, solvents, Bioresources 12 (2017) 1548–1565.
J.C. del Río, W.O.S. Doherty, Catalytic conversion of organosolv lignins to phenolic [46] G.C. Hwang, B.K. Kim, S.Y. Bae, S.C. Yi, H. Kumazawa, Degradation of polystyrene
monomers in different organic solvents and effect of operating conditions on yield in supercritical acetone, J. Ind. Eng. Chem. 5 (1999) 150–154.
with methyl isobutyl ketone, ACS Sustain. Chem. Eng. 6 (2018) 3010–3018. [47] J. Teng, H. Ma, F. Wang, L. Wang, X. Li, Catalytic fractionation of raw biomass to
[39] X. Ouyang, T. Ruan, X. Qiu, Effect of solvent on hydrothermal oxidation depoly- biochemicals and organosolv lignin in a methyl isobutyl ketone/H2O biphasic
merization of lignin for the production of monophenolic compounds, Fuel Proc. system, ACS Sustain. Chem. Eng. 4 (2016) 2020–2026.

471

You might also like