You are on page 1of 10

Catalysis Today 302 (2018) 180–189

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Quenching of reactive intermediates during mechanochemical T


depolymerization of lignin
Alex D. Brittaina,b, Natasha J. Chrisandinaa, Rachel E. Coopera, Michael Buchananb,

John R. Cortc, Mariefel V. Olarted, Carsten Sieversa,b,
a
Georgia Institute of Technology, School of Chemical & Biomolecular Engineering, Atlanta, GA 30332, United States
b
Georgia Institute of Technology, Renewable Bioproducts Institute, Atlanta, GA 30332, United States
c
Biological Sciences Division, Pacific Northwest National Laboratory, P.O. Box 999, Richland, WA 99352, United States
d
Chemical and Biological Processes Development Group, Pacific Northwest National Laboratory, P.O. Box 999, Richland, WA 99352, United States

A R T I C L E I N F O A B S T R A C T

Keywords: Mechanochemical reactions are performed to depolymerize organosolv lignin with sodium hydroxide in a mixer
Mechanocatalysis ball mill. GPC analysis reveals that rapid depolymerization into small oligomers occurs within minutes of milling
Biomass time, followed by a slower reduction in average relative molecular mass over the next 8 h of milling. Monomeric
Ball mill products are identified by GC–MS and quantified by GC-FID. The extent of depolymerization appears to be
Organosolv lignin
limited by repolymerization reactions that form bonds between products. Suppression of these repolymerization
Scavenger
Hydrolysis
reactions can be achieved through the addition of methanol as a scavenger or adjustment of the moisture content
of the feedstock. These modifications result in lower average relative molecular masses and higher yields of
monomers. These results are an important step towards designing an efficient pathway for lignin valorization.

1. Introduction difficulty for heterogeneously catalyzed conversion [4]. Pre-treatment


with ionic liquids [8] or ozone [9] increases the susceptibility of lignin
Conversion of lignocellulosic biomass to biofuels and chemicals has to enzymatic hydrolysis. Steam explosion also cleaves βeOe4 bonds,
attracted tremendous interest in recent years [1]. The U.S. Department which are the most abundant linkages in lignin [10]. Various basic
of Agriculture and the Department of Energy have mandated that 20% catalysts have been used to depolymerize lignin in batch reactors, and
of liquid transportation fuels and 25% of chemicals and materials sodium hydroxide has consistently emerged as an effective option
should be derived from biomass by 2022, and this has driven research [6,11]. Unfortunately, depolymerization under alkaline conditions
in the conversion of raw biomass into valuable products [2]. While generally results in the formation of large amounts of aqueous waste,
many promising concepts have been developed for the conversion of high consumption of base, and presents problems for the isolation of
cellulose and hemicellulose, the lignin fraction of lignocellulosic products [4].
biomass remains underutilized [3]. In 2004, the pulp and paper Another issue that often reduces the effectiveness of lignin depoly-
industry alone produced 50 million tons of extracted lignin, but only merization is that many lignin fragments can undergo condensation
2% were used for commercial products rather than low-value fuel [4]. reactions [4]. Repolymerization can occur simultaneously with depo-
Therefore, the effective and efficient valorization of lignin has great lymerization, and reactive intermediates can react to form stable
potential for providing additional revenue [3,5]. carbon–carbon linkages [10]. Such repolymerization reactions can be
Lignin does not have a regular, repeating structure like cellulose but prevented with scavengers, notably phenol, alcohols like methanol and
consists of a three-dimensional amorphous polymer comprised of ethanol, or formaldehyde [11–15]. These scavengers convert reactive
aromatic rings of varying functionalization, primarily hydroxyl, meth- intermediates to stable molecules before these can cross-link with other
oxy, and propyl groups [6]. The three most abundant monomeric lignin fragments.
species are typically guaiacyl, syringyl, and p-hydroxyphenyl units [1]. Many processes for the conversion of lignocellulosic biomass require
Due to the diversity of linkages between these monomeric units, lignin pre-treatment before the depolymerization reaction, which adds an
is more difficult to depolymerize than cellulose [7]. In addition, lignin extra step to the process [4,6,16]. Milling has been used in academia
exhibits poor solubility in many solvents, which creates particular and industry to reduce the crystallinity of cellulose in biomass.


Corresponding author at: Georgia Institute of Technology, School of Chemical & Biomolecular Engineering, Atlanta, GA 30332, United States.
E-mail address: carsten.sievers@chbe.gatech.edu (C. Sievers).

http://dx.doi.org/10.1016/j.cattod.2017.04.066
Received 21 December 2016; Received in revised form 7 April 2017; Accepted 30 April 2017
Available online 10 May 2017
0920-5861/ © 2017 Elsevier B.V. All rights reserved.
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

However, this pre-treatment often requires long milling times and only equation:
marginally reduces the molecular mass of biomass components. As an mtotalsample − mdriedsample
option for more direct depolymerization of biomass, mechanochemical Moisture Content (wt %) = *100%
mtotalsample (1)
depolymerization has gained attention [17]. Mechanochemistry uses
mechanical energy provided by a ball mill to drive a solventless
catalytic reaction at ambient conditions, avoiding issues caused by 2.3. Mechanochemical reactions
the poor solubility of biomass polymers (i.e. cellulose, lignin) in most
common solvents. Mechanochemical ball milling has been used to Ball-milling of lignin was performed in stainless steel vessels
produce water-soluble oligosaccharides from cellulose by hydrolysis (25 mL) using a Retsch MM400 ball mill at room temperature. The
with both solid acid catalysts and after impregnation with liquid acids vessels were equipped with three milling balls each (12 mm diameter,
[18–25] and to cleave βeOe4 bonds in lignin and model compounds stainless steel). The ball-milling was carried out at 800 rpm (13.3 Hz)
with sodium hydroxide catalysts [26]. More limited cleavage of βeOe4 for 0–8 h. To avoid overheating at long milling times, the mill was
bonds in lignin model compounds was also observed through mechan- stopped for 10 min after every 30 min of milling. For all tests, the
ochemical milling in the presence of water [27–32]. Various hypotheses milling vessel was filled with organosolv lignin (1.50 g) and sodium
for the way in which mechanochemistry operates have been proposed, hydroxide pellets (1.50 g). For ion scavenging tests, methanol
including that friction provides energy in the form of high temperature (0.40 mL) was added to the lignin/sodium hydroxide mixture in the
pockets that enable molecules to overcome a solid–solid diffusion milling vessel. To investigate the influence of the moisture content,
barrier [21,26,33–35], or that mechanical impact deforms molecules water was added to the milling vessel to increase the total moisture
resulting in higher, more reactive energy states [34,36]. content in the sample from its original value of 3.7 wt% to the desired
In mechanochemical conversion of cellulose, high yields of oligo- moisture content. Finally, to serve as a control experiment, organosolv
mers are readily obtained, but yields of monomeric products are lignin samples (1.50 g) and sodium hydroxide pellets (1.50 g) were
relatively low [19]. Several theories for this phenomenon have been milled for 1 min at 800 rpm (13.3 Hz) separately, then combined and
put forth [19,22]. Some attention has been paid to the moisture content shaken in the mill without milling balls for 10 s. These samples were
of feedstock, because the dominant reaction is presumed to be hydro- left to sit for 5 min–8 h before being prepared for molecular mass
lysis [19,21]. While the addition of certain quantities of water is analysis.
necessary to provide an otherwise limiting reactant [37], addition of
excess water can result in a plasticization, which dissipates the 2.4. Gas chromatography – mass spectrometry (GC–MS) and gas
mechanical energy from grinding, limiting chemical reactions or chromatography – flame ionization detector (GC-FID)
preventing grinding altogether in extreme cases [38]. Another theory
for the limited depolymerization is the occurrence of repolymerization The milled samples were dissolved in methanol and diluted to a
reactions [19,22], similar to the reactions of intermediates in thermo- concentration of 10 mg/mL. The samples were neutralized by addition
chemical processes [19,39]. Thus, it will be important to explore ways of hydrochloric acid in an amount equivalent to the sodium hydroxide
for preventing repolymerization of reactive intermediates in mechan- in the reaction mixture. 3,5-Dimethoxyphenol was used as an internal
ochemical reactions. standard. Samples were filtered using 0.2 μm polypropylene mem-
In this work, mechanochemical reactions for lignin depolymeriza- branes. GC–MS analysis was performed using a Varian (Agilent) 450-
tion with sodium hydroxide are investigated. Prevention of repolymer- GC with a FactorFour™ VF–35 ms capillary column
ization reactions through the use of methanol as a scavenger is also (30 m × 0.25 mm × 0.25 μm) coupled with a 300-MS Varian
studied. Finally, the effects of the moisture content of the lignin (Bruker) mass spectrometer (EI, 200 °C). The carrier gas was helium
feedstock are explored. Molecular weight analysis of the products is at 1.0 mL/min with an autosampler injection volume of 1.0 μL. The
performed by GPC, monomeric products are identified and quantified temperature of the column was initially held at 70.0 °C for 1.50 min,
by GC analysis, and an analysis of chain length distributions is obtained then raised at a rate of 35.0 °C/min to 100.0 °C. The temperature of the
by LC–MS. In addition, linkage motifs of selected samples are examined column was then raised at a rate of 10 °C/min to 300.0 °C, where it was
by HSQC NMR. held for 12.0 min.
To obtain more quantitative results for the yields of monomers,
2. Experimental section samples were simultaneously injected into the same Varian (Agilent)
450-GC equipped with a flame ionization detector (FID) with a split
2.1. Materials ratio of 20:1. The GC contained the same column and used the same
temperature program as the GC–MS method described previously. The
Organosolv lignin from beech was obtained from Fraunhofer Institut FID was set at 300 °C, with a make-up flow of 29.0 mL/min, an H2 flow
in Leuna, Germany. Sodium hydroxide (ACS reagent, ≥97.0%, pellets), of 30.0 mL/min, and an air flow of 300 mL/min. It is noted even with
methanol (ACS spectrophotometric grade, ≥99.9%), methanol FID the response factors of different compounds can vary notably [40].
(CHROMASOLV® for HPLC, ≥99.9%), tetrahydrofuran (ACS reagent, > Yield of total detected monomeric species was calculated as follows:
99.0%), acetic acid (glacial, ACS reagent, ≥99.7%), and 3,5 dimethox- m volatile monomers
yphenol (99%) were purchased from Sigma-Aldrich. Hydrochloric acid yield (wt %) = *100%
mtotal reacted sample (2)
(ACS reagent, 36.5–38.0%, liquid) and o-xylene (99%) were obtained
from Alfa Aesar. Ethyl acetate (ACS reagent, 99.9%) was purchased Selectivity of each detected monomeric species was determined as
from Fisher Scientific. d6-DMSO was obtained from Cambridge Isotope follows:
Laboratories. m volatile species
selectivity (wt %) = *100%
mtotal volatile monomers (3)
2.2. Moisture content of organosolv lignin sample

Five samples of the original organosolv lignin (2.00 g) were placed 2.5. Gel permeation chromatography (GPC)
in an oven at 105 °C to evaporate the water content of the sample. The
samples were weighed periodically to track changes in mass. After 48 h, The milled samples were dissolved in THF and diluted to a
when the samples had reached constant masses, the moisture content concentration of 4 mg/mL. The samples were neutralized by addition
was calculated as the average of the five samples using the following of hydrochloric acid in an amount equivalent to the sodium hydroxide

181
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

in the reaction mixture. Samples were filtered using 0.2 μm nylon (0.6 mL) was added to a vial and combined with pyridine (0.9 mL) and
membranes. An amount of 1.0 μL of each sample was injected into an the lignin sample (18 mg). Phosphiltylating agent 2-chloro-4,4,5,5-
Agilent PL-GPC-50 with a Refractive Index Detector. THF was used as tetramethyl-1,3,2-dioxaphospholane (TMDP) was added in an amount
an eluent with a flow rate of 1.0 mL/min, and polystyrene standards equivalent to 2 mmol/mmol water + 1 mmol/mmol oxygen content.
31
were used for calibration. The data was analyzed using Cirrus GPC P NMR analysis was carried out as described previously [43].
software (Version 3). Instrument parameters were as follows: number of scans > 128, 90°
pulse width, 1.2 s acquisition time, 25 s relaxation delay (d1), and
2.6. Liquid chromatography – mass spectrometry (LC–MS) inverse-gated decoupling. The regions of interest were 152–145 ppm
(aliphatic OH), 138–145 ppm (phenolic OH), and 134.6–136 ppm
LC–MS analysis was performed using a modified procedure based on (carboxylic OH).
the work of Haupert et al. [41] The milled samples were dissolved in a Calculation of the concentration of hydroxyl groups per gram of
50:50 (v:v) methanol/water solution and diluted to a concentration of sample was determined as follows:
100 μg/mL. o-Xylene was used as an internal standard. Samples were li
filtered using 0.2 μm polypropylene membranes. LC–MS analysis was mmol OHi lTPPO
*[TPPO] * total mass (g) of NMR sample
=
performed on a Shimadzu LCMS-8060 using an ESI source in negative g lignin sample mass (g) lignin sample (4)
ion mode with a spray voltage of 5.0 kV. The nebulizing gas flow of
where i = aliphatic, phenolic, or carboxylic region.
nitrogen was set at 3.0 L/min, and the drying gas flow of nitrogen was
set at 15.0 L/min. An autosampler was used to inject 6 μL of each
sample onto a Phenonenex Luna 5 μ Silica column with dimensions 2.9. Carbonyl titration
150 mm × 4.60 mm × 5 μm. The mobile phase was a 50:50 (v:v)
solution of methanol and water at 0.300 mL/min. Milled samples were dissolved in water and diluted to 50 mg/mL.
Samples were neutralized by addition of hydrochloric acid in an
2.7. Heteronuclear single quantum coherence (HSQC) NMR spectroscopy amount equivalent to the sodium hydroxide in the reaction mixture.
Reaction products were extracted into an organic layer of ethyl acetate,
Milled samples were dissolved in water and diluted to 50 mg/mL. and the solvent was subsequently removed under vacuum in a round-
Samples were neutralized by addition of hydrochloric acid in an bottom flask.
amount equivalent to the sodium hydroxide in the reaction mixture. Carbonyl titration was carried out using the modified Faix method,
Reaction products were extracted into an organic layer of ethyl acetate, as described previously [43,44]. Each lignin sample (100 mg) was
and the solvent was subsequently removed under vacuum in a round- added to DMSO (1 mL), and dissolved in hydroxylamine hydrochloride
bottom flask. (2 mL) and triethanolamine (2 mL). This vial was closed tightly and
Lignin samples (ca. 50 mg) were dissolved in 600 μL d6-DMSO and stirred for 2 h at 80 °C. This solution was then transferred to a titration
transferred to 5 mM Wilmad 535-PP NMR tubes. The solvent contained vessel and diluted to 80% with ethanol and water, and subsequently
0.05% (v/v) tetramethyl silane (TMS) for chemical shift referencing. titrated with 0.1N hydrochloric acid until the endpoint was reached.
One-dimensional 1H and two-dimensional 1H–13C heteronuclear single This procedure was repeated three times without the lignin samples to
quantum coherence (HSQC) spectra were acquired as described pre- constitute blank experiments.
viously[42] at 298 K on a 500 MHz Varian Inova equipped with a The concentration of carbonyl groups (mmol CO/g lignin) in each
Nalorac HCNP triple resonance z-axis pulsed-field gradient probe. 1H lignin sample was determined as follows:
spectra were collected at 293 K in DMSO at 600 MHz with 4 transients, EPBA − EP
[CO] = *[acid ]
8k complex points, 1 s relaxation delay, and processed with 2 x zero mass (g) lignin sample (5)
filling and no apodization. Using the BioPack gChsqc pulse sequence,
separate HSQC spectra were collected for the aliphatic and aromatic where EPBA and EP represent the endpoints of the average of the blank
regions of the 13C chemical shift range. Delay times tCH and lambda for runs and the endpoint of the lignin sample run in mL, respectively.
1/4*JCH, were 1.8 ms and 1.6 ms for aliphatic spectra, and 1.45 ms and
1.3 ms for aromatic spectra. The 1H spectral width was 17 ppm and 13C 3. Results
spectral width was 100 or 60 ppm for the aliphatic or aromatic regions,
respectively. Spectra were collected with 1024 points (Varian para- 3.1. Characterization of the organosolv lignin feedstock
meter np) and 61 ms acquisition time in the 1H dimension and with 128
transients and 192 or 128 complex points (Varian parameter ni in The moisture content of the unmodified organosolv lignin sample
States-TPPI mode) in the indirect 13C dimension, for aliphatic and was 3.7 wt%. Gel Permeation Chromatography (GPC) analysis showed
aromatic spectra, respectively. Adiabatic WURST decoupling was that the average relative molecular mass was approximately 800 g/mol,
applied during acquisition. Data were processed using MestReNova which corresponds to between 5 and 6 monomeric units. NMR spectro-
9.01 with 2X zero filling in both dimensions and matched Gaussian scopy was used to characterize the unmilled lignin feedstock (Fig. 1a).
apodization in both dimensions. Integration of peak areas used identical Using volume integrals of characteristic HSQC cross peaks for α, β, and
elliptical boundaries for all spectra. γ protons in the propyl groups of monolignol residues in lignin [45–48],
the relative abundance of the major linkages in the unmilled lignin was
2.8. Phosphitylation followed by 31
P NMR spectroscopy estimated to be 57:19:24 for βeOe4 (β-aryl ether), β-5 (phenylcou-
maran), and β-β (resinol), respectively. The βeOe4 species comprised
Milled samples were dissolved in water and diluted to 50 mg/mL. several distinct species including a small proportion of linkages to α-
Samples were neutralized by addition of hydrochloric acid in an oxidized (carbonyl) units. The aromatic region of the HSQC spectrum
amount equivalent to the sodium hydroxide in the reaction mixture. was used to determine that the ratio of guaiacyl to syringyl units (G:S
Reaction products were extracted into an organic layer of ethyl acetate, ratio) was 40:60, which is typical of hardwood lignins.
and the solvent was subsequently removed under vacuum in a round-
bottom flask. 3.2. Reactivity studies
A solvent solution was created from a mixture of CDCl3 (2.31 mL),
relaxant chromium (III) acetylacetonate (6.3 mg), and internal standard The mechanochemical conversion of organosolv lignin with NaOH
triphenylphosphine oxide (TPPO) (41.7 mg). A portion of this solution was performed with pure lignin and after addition of methanol or

182
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

Fig. 1. Downfield region of the aliphatic 1H-13C HSQC spectra showing cross peaks characteristic of different types of linked monolignol building blocks in unmilled lignin in DMSO
(panel a), and dried ethyl acetate extracts of mechanocatalytically processed lignin with (panel b) and without (panel c) methanol treatment, in DMSO. Key cross peaks used to determine
relative abundance of linkages are indicated.

water. The following sections describe the analysis of the average with polystyrene standards, as a function of milling time. The lignin
relative molecular mass (GPC), nature and approximate yield of samples milled without the addition of any catalysts or reactants
monomeric species (GC–MS/GC-FID), distribution of chain lengths underwent only a minimal change for the duration of the milling times
(LC–MS), and distribution of linkages (HSQC NMR spectroscopy), tested, as did the lignin samples mixed with sodium hydroxide but left
before the reactions and after the process are discussed. unmilled. This indicates that energy and appropriate reactants or
For studies, in which additional water was supplied to the milling catalysts are required for depolymerization of lignin. When sodium
vessel, attention was paid to the consistency of the resulting mixture hydroxide was added to the milling vessel, the relative number average
because water has a known plasticizing effect on biomass, which molecular mass was reduced from approximately 800 g/mol to 410 g/
impedes its ability to be ground in a mill [38]. For moisture contents mol after 8 h of milling. Most of this depolymerization occurred within
up to 14 wt%, the samples remained as powders, but when the moisture the first 30 min, after which the relative molecular mass was around
content was increased to 20 wt%, the sample possessed physical 470 g/mol. The relative molecular mass was slowly reduced to 410 g/
properties similar to a sludge, and the milling balls stuck to the sides mol after 5 h of milling, and remained at this value up to 8 h. Inclusion
of the container and to each other. of methanol as an additional reactant in the milling vessel enhanced net
depolymerization. The initial reduction in relative molecular mass was
3.3. Gel permeation chromatography (GPC) more rapid, and a value of 400 g/mol was reached in only 5 min of
milling. Afterwards, the decrease in relative molecular mass was much
Fig. 2 displays the number average molecular mass (Mn) and weight slower, but it continued to a final value of under 250 g/mol after 8 h of
average molecular mass (Mw), relative to narrow standard calibration milling. The results for relative weight average molecular mass

183
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

Fig. 2. a) Number average molecular mass (Mn), and b) weight average molecular mass
(Mw), of products from lignin as determined by GPC, plotted as a function of milling time.
Mixtures of 3.00 g uncatalyzed lignin (x), 1.50 g lignin combined with 1.50 g NaOH (○) Fig. 3. a) Number average molecular mass (Mn), and b) weight average molecular mass
and 1.50 g lignin combined with 1.50 g NaOH and 0.4 mL of methanol (△) were milled (Mw) of products from lignin, as determined by GPC, plotted as a function of moisture
in a mixer ball mill at 800 rpm. Unmilled samples of 1.50 g lignin and 1.50 g NaOH (□) content of the milled sample. Lignin samples were milled with sodium hydroxide (50:50,
are included for comparison. w:w) in a 25 mL mixer ball mill for 2 h at 800 rpm. Each test sample contained 3.00 g of
total substrate before addition of water to achieve the necessary moisture content.

followed similar trends as the number average molecular mass data.


GPC analysis was also performed for samples milled for 2 h with Table 1
sodium hydroxide and different moisture content (Fig. 3). The moisture Average apparent selectivity of monomeric species produced by the mechanochemical
reaction between lignin and sodium hydroxide after 5 h.
contents ranged from that of untreated lignin (3.7 wt%) to 20 wt%. The
relative molecular mass of these samples showed a steady decrease up Compound Average Selectivity/wt%
to 18% moisture, where the Mn reached a minimum of 340 g/mol.
However, in the presence of 20 wt% moisture, the samples only reached Propane, 1,1,3,3-tetramethoxy- 2.35
Phenol 29.42
a relative molecular mass of 460 g/mol, larger than the 430 g/mol
Guaiacol 7.91
reached by the unmodified lignin samples with 3.7% moisture. 4-methoxy-3-methylphenol 0.39
Creosol 0.34
Phenol, 4-methoxy-3-methyl- 0.25
3.4. Gas chromatography – mass spectrometry (GC–MS) and gas Unidentified 0.04
chromatography – flame ionization detector (GC-FID) 4-ethylguaiacol 0.05
4-vinylguaiacol 1.88
The products created from the mechanochemical reactions between o-Eugenol 0.12
Syringol 17.85
lignin and sodium hydroxide from 0 to 8 h were studied using GC–MS.
Isoeugenol 0.31
Table 1 shows the monomeric species identified in the product mixture Methyl dodecanoate 0.86
obtained from the mechanochemical reactions. The most abundant of Unidentified 0.16
these compounds included phenol, syringol, and guaiacol, which are 1,2,4-trimethoxybenzene 1.70
among the most common monomeric building blocks found in most 3,4,5-trimethoxytoluene 1.16
Canolol (2,6-dimethoxy-4-vinylphenol) 4.10
types of lignin feedstocks. The other monomers present are comprised
1,2-dimethoxy-4-(2-methoxyethenyl)benzene 1.12
primarily of similarly functionalized aromatic rings containing hydro- 2,4,5-trimethoxybenzaldehyde 18.85
xyl, methoxy, and other groups common to the structure of lignin. 2,6-dimethoxy-4-(2-propenyl)phenol 5.02
Other monomers found represent extractive species typically found in Methyl palmitate 4.12
Octadecanoic acid, methyl ester 2.01
hardwoods, such as methyl dodecanoate and methyl palmitate [49].
Fig. 4 shows the combined apparent yield of these monomeric species
as a function of milling time, as determined by GC-FID. began to decrease once more. The maxima at 5 min of milling
For all of the samples tested, yield of monomeric species increased corresponded to yields of 15% and 13% for samples milled with only
sharply after only 5 min of milling. However, when the milling time sodium hydroxide and samples milled with both sodium hydroxide and
was increased to 10 min, the yield dropped significantly. As milling methanol, respectively. At the maxima after 5 h of milling, the yields
time was increased beyond 10 min, the yield increased again, reaching were 16% and 27%. Fig. 5 shows yields as a function of increasing
a maximum for all samples at 5 h of milling. By 8 h of milling, the yield

184
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

Fig. 4. Yield of monomeric species produced, as determined by GC-FID, plotted as a


function of milling time. For these experiments, 1.50 g lignin combined with 1.50 g NaOH
(○) and 1.50 g lignin combined with 1.50 g NaOH and 0.4 mL of methanol (△) were
milled in a mixer ball mill at 800 rpm.

Fig. 6. Yield of monomers (○), dimers (△), and trimers and larger oligomers (X), as
determined by LC–MS, plotted as a function of milling time. a) 1.50 g lignin combined
with 1.50 g NaOH and b) 1.50 g lignin combined with 1.50 g NaOH and 0.4 mL of
Fig. 5. Yield of monomeric species produced, as determined by GC-FID, plotted as a methanol were milled in a mixer ball mill at 800 rpm.
function of moisture content of the milled sample. Lignin samples were milled with
sodium hydroxide (50:50, w:w) in a 25 mL mixer ball mill for 2 h at 800 rpm. Each test % and 70 wt%) of the sample was comprised of dimers after only five
sample contained 3.00 g of total substrate before addition of water to achieve the
minutes of milling and remained the majority for the duration of the
necessary moisture content.
milling times tested. The one exception to this was the lignin sample
milled with both sodium hydroxide and methanol for five hours, which
moisture content. Monomer yield increased only slightly up to 12%
resulted in the monomer yield slightly surpassing the dimer yield,
moisture, but a significant maximum of 28% was seen with a moisture
which dropped to 41 wt%.
content of 14 wt%. When the moisture content was higher than 14 wt%,
As seen in Fig. 7, modification of the moisture content between the
the yield decreased to a value similar to those seen before the
original 3.7 wt% and 12 wt% produced only slight variation in the
maximum.
distribution of the products formed. With 14 wt% moisture, however, a
large amount of monomeric species was produced and the dimer yield
3.5. Liquid chromatography – mass spectrometry (LC–MS)
dropped significantly. At moisture contents higher than 14 wt%, the
monomeric and dimeric yields decreased and increased again, respec-
LC–MS analysis was performed to determine the distribution of the
products of the mechanochemical reactions based on the size of the
various species present. By taking the average monomer molecular
mass to be 166 g/mol (based on an average molecular formula of
C9H10O3 from the monomers detected by GC–MS), products were
classified as monomers, dimers, and oligomers with 3 or more units
(Fig. 6). Although the majority of product samples were soluble in the
methanol/water solution used as a solvent for this method, it is possible
that small amounts of larger products were insoluble, and thus not
analyzed in the LC–MS.
The monomer yield followed the same trend as the GC–MS results.
Examination of the unmilled lignin sample revealed values of approxi-
mately 3 wt% monomers, 32 wt% dimers, and 65 wt% trimers and
larger oligomers. Analysis of samples milled with sodium hydroxide
(Fig. 6a) and samples milled with both sodium hydroxide and methanol
(Fig. 6b) showed that the fraction of oligomers with a chain length of 3
Fig. 7. Yield of monomers (○), dimers (△), and trimers and larger oligomers (X), as
or more dropped to slightly over 20 wt% within only five minutes of determined by LC–MS, plotted as a function of moisture content of the sample. Lignin
milling. This value increased to 30 wt% after ten minutes of milling, samples were milled with sodium hydroxide (50:50, w:w) in a 25 mL mixer ball mill for
before it reached a maximum after thirty minutes and subsequently 2 h at 800 rpm. Each test sample contained 3.00 g of total substrate before addition of
decreased. For both of these experiments, the majority (between 50 wt water to achieve the necessary moisture content.

185
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

tively. included an extraction of the sodium hydroxide, it is possible that


residual sodium hydroxide may have affected these results.
3.6. Heteronuclear single quantum coherence (HSQC) NMR spectroscopy
3.8. Carbonyl titration
Samples milled for 2 h with NaOH in the presence and absence of
methanol were analyzed by HSQC NMR spectroscopy. Compared with Unmilled lignin as well as lignin milled for 2 h with sodium
the unmilled sample (Fig. 1a), milling with sodium hydroxide changed hydroxide, with and without methanol were also subjected to carbonyl
the relative abundances of βeOe4 (β-aryl ether), β-5 (phenylcoumar- titration, to determine the carbonyl content. For both milled samples,
an), and β-β (resinol) linkages from 57% to 56%, 19% to 16%, and 24% the carbonyl content fell from 2.50 (unmilled) to 2.38 mol C]O/g
to 28%, respectively (Fig. 1b). Increased peak dispersion in regions sample. However, these deviations fall well within the variability of the
containing cross peaks associated with these linkages can be seen as titration method (∼5.4%) [43], so these changes in carbonyl content
well, most notably for the Aβ and Cγ peaks in panels b and c of Fig. 1, are effectively insignificant.
suggesting changes that increase structural heterogeneity of the lignin
have occurred. If one normalized the intensities to that of β-β linkages 4. Discussion
because of the resistance of carbon–carbon bonds to cleavage with basic
catalysts [4], the relative abundance of βeOe4 linkages would 4.1. Depolymerization of lignin with sodium hydroxide
decrease by about 15% during this process. The presence of methanol
in the mechanochemical reaction mixture (Fig. 1c) changed the relative Base-catalyzed depolymerization of lignin in the aqueous phase is a
proportions of the major linkages towards fewer βeOe4 linkages common method for producing simple aromatic monomers. Basic
(45:21:32), corresponding to a 65% decrease relative to a constant catalysts attack ether bonds, which are both the most abundant [6]
number of β-β linkages. Through analysis of the 1-D proton spectra and the weakest linkages between the monomeric building blocks [51].
(Fig. S1), the ratio of aliphatic ether and alcohol protons to aromatic Of these ether linkages, the βeOe4 bond is the most common in
protons was found to be 3.33 for the unmilled sample, 3.44 for the unextracted lignin, comprising between 45% and 65% of all linkages
sample milled with sodium hydroxide, and 3.38 for the sample milled between monomeric units in various lignin sources [52]. Strong bases
with sodium hydroxide and methanol. The change in these ratios is not like sodium hydroxide are particularly effective at hydrolytically
substantial, which indicates that comparing the fractions of linkages is cleaving this bond under various aggressive conditions [11,53,54].
similar to measuring them relative to selected aromatic protons. It is On the other hand, non-ether linkages, such as β-5 or 5e5 bonds, are
important to note that these results may be influenced by any residual typically resistant to base catalyzed depolymerization [4]. These
water in the tested samples. carbon–carbon bonds can make up as much as one third of the linkages
between monomers in unextracted lignin [6]. After lignin is isolated,
3.7. Phosphitylation followed by 31
P NMR spectroscopy the relative abundance of linkages may be altered drastically depending
on the method used [3].
31
P NMR spectroscopic analysis was performed for three phosphi- Kleine et al. showed that the linkages in organosolv lignin and lignin
tylated samples (unmilled lignin, lignin milled with sodium hydroxide model compounds can be cleaved in base-catalyzed mechanochemical
for 2 h, and lignin milled with sodium hydroxide and methanol for 2 h), reactions following similar reaction path as in thermochemical reac-
in order to determine the concentrations of hydroxyl groups for the tions in water [26]. Milling of lignin model dimers containing the
aliphatic, phenolic, and carboxylic regions (Fig. 8). The concentration βeOe4 bond in the presence of a strong base, such as NaOH, KOH, or
of aliphatic hydroxyl groups was reduced from 1.91 mmol OH/g to LiOH, resulted in high yields of monomers in relatively short times.
approximately half this value in the samples milled with and without Milling of organosolv lignin and beech wood with sodium hydroxide led
methanol. Phenolic hydroxyl groups declined slightly from 3.36 mmol to effective cleavage of the βeOe4 linkages. However, only a single
OH/g sample to 2.76 and 3.17 mmol OH/g sample upon milling with time point of 12 h was provided for the analysis of organosolv lignin
and without methanol, respectively. Milling with only sodium hydro- depolymerization, which prevents an interpretation of the dynamic
xide caused a significant increase in carboxylic hydroxyl groups from effects of mechanochemistry, in particular, the simultaneous occur-
0.23 to 1.05 mmol OH/g sample, while milling with both sodium rence of depolymerization and condensation reactions [26]. Further-
hydroxide and methanol only resulted in a marginal increase to more, a milling time of 12 h is too long for a commercially viable
0.37 mmol OH/g sample. It is important to note that the presence of process. Additionally, changes in the average relative molecular mass of
sodium hydroxide can lead to an underestimation of OH groups organosolv lignin and the nature and yield of different products were
[43,50]. Although the preparation of the samples for this analysis not reported.
Mechanochemical reactions of lignin model compounds were also
studied in a series of works by Sumimoto et al. [27–32] These studies
showed that various model compounds undergo chemical reactions
under mechanochemical conditions. These reactions were performed in
the absence of a base, and isomerization reactions and limited homo-
lytic cleavage of bonds were the primary reactions observed [29,31,32].
In instances where reactions were performed using excess water as a
solvent, homolytic bond cleavage was proposed to be more prevalent,
resulting in the creation of %O2H, %OH, and %H radicals formed from a
mechanochemically-induced reaction between water and oxygen
[31,32]. These homolytic radical hydrolysis reactions contrast with
the heterolytic hydrolysis reactions observed by Kleine et al. in the
presence of a base [26]. Additionally, the yields of monomeric reaction
products were greatly increased for similar milling times in the
presence of a base [26,31,32].
Fig. 8. Concentration of aliphatic, phenolic, and carboxylic OH groups of three selected The experiments in the present work clearly show that sodium
samples (unmilled lignin, lignin milled with NaOH for 2 h, and lignin milled with NaOH hydroxide is capable of depolymerizing organosolv lignin under
and MeOH for 2 h). mechanochemical conditions. The average relative molecular mass of

186
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

the sample can be reduced significantly within the first 30 min of condensation reactions has yet to be developed. This may be because
milling, after which depolymerization slows, and Mn eventually the necessity for preventing these reactions is limited because the
reaches a minimum value of approximately 410 g/mol, roughly half oligosaccharides produced from these reactions are sufficiently soluble
of the original average relative molecular mass of the unmilled sample. in water to allow for simple separation of products and subsequent
LC–MS analysis shows that the majority of the resulting product conversion into monosaccharides [23,58–60]. Depolymerization of
consists of dimers after only 5 min of milling. The resistance of the lignin via mechanochemistry presents a greater challenge due to the
samples to fully depolymerize could be due to the presence of CeC lower solubility of lignin oligomers in most solvents compared to
linkages that are resistant to cleavage using basic catalysts [6]. These oligosaccharides [61]. This presents a significant challenge for product
stronger carbon–carbon bonds in lignin can be effectively broken in separation and subsequent downstream processing. Thus, the need for
thermochemical cracking reactions, which typically utilize strong prevention of repolymerization/condensation reactions in mechano-
Brønsted acid catalysts, such as zeolites [6]. Additionally, the yield of chemistry of lignin is essential.
monomers fluctuates between minima and maxima, corresponding to A promising approach to prevent repolymerization reactions is the
maxima and minima in the yield of trimers and larger oligomers, addition of scavengers that quench reactive intermediates [11–14].
respectively. This suggests that repolymerization reactions are taking Proton-donating scavengers include formic acid, phenol, 2-naphthol,
place simultaneously with the depolymerization reaction, presenting and various alcohols. The donation of H+ ions to reactive carbanion
another obstacle to complete depolymerization. intermediates has been shown to reduce char formation by repolymer-
After an initial maximum in the yield of monomeric species after ization reactions in thermochemical reactions [4]. In addition to
5 min of milling, this value drops sharply. The rapid initial production scavenging ionic intermediates, organic solvents, such as methanol,
of monomers in this initial 5 min is likely a result of easily accessible can also act as radical scavengers preventing radical reactions like those
regions in the original lignin structure along with the availability of suggested by Sumimoto et al. [32]
weakly bound water molecules. The rapid depletion of easily accessible To prevent repolymerization reactions, methanol was used as a
water favors repolymerization/condensation reactions after five min- scavenger in the present study, and positive effects on both relative
utes of milling, which consumes a portion of the previously produced molecular mass reduction and yield of monomeric products were
monomeric species. However, as milling progresses from 10 min to 5 h, observed. The initial net rate of depolymerization was more rapid,
more bound water becomes accessible within the structure of lignin and the final average relative molecular mass of the reaction products
[55], leading to a rise in production of monomers. In this period, decreased by an additional 40% compared to the samples milled with
condensation reactions could also occur due to localized depletions of sodium hydroxide in the absence of methanol. Furthermore, the
water molecules, reproducing βeOe4 bonds or forming new CeC maximum monomer yield after 5 h of milling increased by 70%, and
bonds, which are unbreakable by base-catalyzed hydrolysis [6]. Strong the fraction of cleaved βeOe4 bonds detected by 2D NMR after 2 h of
evidence for the formation of new βeOe4 bonds is the limited milling increased from 15% to 65%. The addition of methanol also
reduction in βeOe4 bonds after 2 h of milling as observed by NMR, limited the formation of new carboxylic hydroxyl groups as detected by
31
while GPC showed a decrease of Mn to nearly half of its original value. P NMR spectroscopy after phosphitylation with TMDP. All these
However, the continuous release of bound water ultimately pushes the results suggest a reduction of the extent to which repolymerization
reaction towards hydrolysis rather than the condensation pathway. reactions take place when methanol is added to the reaction mixture.
Between 5 and 8 h of milling, the yield of monomers decreases. This Interestingly, neither the species of monomers detected by GC–MS
could be related to eventual formation of more recalcitrant bond during nor their respective selectivities determined by GC-FID (Table 1) were
repolymerization of lignin fragments. An increasing fraction of the altered significantly when methanol was added to the milling vessel.
remaining linkages appears to be CeC bonds due to their resilience and Analysis of the 2D NMR spectra also showed no evidence of new
the formation of additional ones in repolymerization reactions. Thus, methoxy groups in the sample upon the addition of methanol. At the
the previously produced monomers are again consumed in the forma- same, time the detection of some monomeric products with three
tion of repolymerization products. This is marked by an increase in the methoxy groups implies that transfer of such groups between different
yield of trimers and larger oligomers at 8 h of milling. aromatic rings appears to occur to a certain extent because such
building block are not common in lignin [4]. Further experiments,
4.2. Use of methanol as a scavenger for reactive intermediates such as reactions with isotopically labeled methanol, are needed to gain
certainty regarding the potential incorporation of carbon atoms from
One of the biggest issues for effective depolymerization of lignin is methanol in the products. However, based on the current results, it
the prevention of repolymerization reactions [4,6,56]. In base-cata- appears more likely that methanol is donating a proton to reactive
lyzed hydrolysis of lignin, unstable intermediates may react with each carbanion intermediates that are formed in the reaction, rather than a
other to recombine into larger oligomers rather than completing the methoxy ion to a carbocation species. This donated proton can be
transformation to stable monomers [56]. The char that is formed in replenished by the small amount of water in the system, consuming the
these repolymerization reactions typically contains carbon–carbon water and completing the base-catalyzed hydrolysis reaction. However,
bonds that are resistant to cleavage with basic catalysts [6]. the potential role of radical intermediates also needs to be investigated
The dominant mechanism of cellulose depolymerization in an acid- further.
impregnated, ball-milled mechanochemical system has been suggested It is also possible that methanol may contribute to the depolymer-
to be through hydrolysis rather than homolytic cleavage to radicals that ization reaction through a solvolysis reaction in conjunction with the
are known to appear in mechanical pre-treatments of biomass [18]. In hydrolysis reaction. This solvolysis reaction can continue until the
contrast, the formation of radicals by cleavage of ether bonds has been methanol or NaOH are completely consumed, producing sodium
proposed to occur to a limited extent during mechanical treatment of methoxide from the sodium hydroxide originally in the system. Base
biomass [31,32,57]. Specifically, Sumimoto et al. proposed the genera- catalyzed solvolysis of ether linkages in lignin using small alcohols
tion of radical species by reactions between oxygen and water during using thermochemical methods has been documented in several pub-
mechanochemical reactions of lignin model compounds in aqueous lications and is mechanistically similar to hydrolysis of ethers and
phase. The lower water content during the mechanochemical reactions esters, producing basic alkoxides from hydroxides [54,62]. The metha-
described here likely limits this path for radical generation. nol can be subsequently regenerated upon the addition of water [63].
Repolymerization reactions were shown to reduce the yield of Additionally, solvolysis with methanol has been shown to be less
monosaccharides during mechanochemical depolymerization of cellu- kinetically limited than hydrolysis, which can reduce the chance of
lose [19,22], and a viable strategy for the prevention of these intermediates undergoing repolymerization reactions [62].

187
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

4.3. Adjustment of moisture content nature of the scavenger, the most abundant monomeric products
include phenol, 2,4,5-trimethoxybenzaldehyde, syringol, and guaiacol.
Base-catalyzed depolymerization of lignin is predominantly a These products have similar molecular structures, making them an
hydrolysis reaction [6]. Thus, the amount of water present in the attractive stream for the production of aromatics.
system plays an important role. Generally, water in mechanochemical
reactions is provided through the moisture content of the biomass Acknowledgements
feedstock [21]. Several researchers have suggested that the addition of
more water to the system may be needed to achieve the desired The authors would like to acknowledge the Renewable Bioproducts
reactivity for the depolymerization reactions [19,21]. However, the Institute at the Georgia Institute of Technology for their funding
only reported use of increased moisture in a mechanochemical system through the Paper Science and Engineering Fellowship. Imerys is
resulted in a decrease in depolymerization of cellulose [19]. The thanked for a generous donation. Dr. Christopher Jones and Caroline
tradeoff of excess water is plasticization due to the formation of Hoyt are thanked for usage of and training on GPC equipment, and Dr.
hydrogen-bonds, which lead to a significant reduction of the glass Charles Liotta and Dr. Fiaz Mohammed are thanked for usage of and
transition temperature [64]. This phenomenon prevents efficient mix- training on LC–MS equipment. Brandan Brown and Lucas Ferreira are
ing and grinding. In the present work, well-controlled amounts of water thanked for experimental support. The authors also acknowledge the
were added to the lignin to avoid the detrimental plasticizing effects of financial support of the Bioenergy Technologies Office of the US.
excess water. Department of Energy under contract number DE-AC06-76RLO-1830
The fresh lignin feedstock contained 3.7 wt% water. Because the with Battelle as well as the technical support of Ms. Marie Swita for
hydrolysis reaction requires one water molecule for each linkage carbonyl analysis.
broken, the original moisture content is not large enough to allow for
hydrolysis of all the ether bonds in the lignin. If one assumes the Appendix A. Supplementary data
average molecular mass of a lignin monomer to be 166 g/mol and that
57% of the linkages in unmilled lignin are ether linkages (as indicated Supplementary data associated with this article can be found, in the
by NMR spectroscopy) a moisture content of 6.0 wt% would be required online version, at http://dx.doi.org/10.1016/j.cattod.2017.04.066.
to cleave all the ether linkages. However, the maximum in yield of
monomers was observed at 14% moisture. This represents the optimal References
tradeoff between the availability of water for hydrolysis while avoiding
significant plasticization. The discrepancy between the calculated [1] A.J. Ragauskas, C.K. Williams, B.H. Davison, G. Britovsek, J. Cairney, C.A. Eckert,
stoichiometric amount of water and this result is likely related to W.J. Frederick, J.P. Hallett, D.J. Leak, C.L. Liotta, Science 311 (2006) 484–489.
[2] V. Balan, ISRN Biotechnol. 2014 (2014) 463074.
non-perfect mixing of the solid system. Curiously, this maximum in [3] R. Rinaldi, R. Jastrzebski, M.T. Clough, J. Ralph, M. Kennema, P.C. Bruijnincx,
monomer yield did not correspond to a significant decrease in average B.M. Weckhuysen, Angew. Chem. Int. Ed. 55 (2016) 8164–8215.
relative molecular mass due to a decrease in dimer yield, but an [4] J. Zakzeski, P.C. Bruijnincx, A.L. Jongerius, B.M. Weckhuysen, Chem. Rev. 110
(2010) 3552–3599.
increase in the yield of higher length chains. This suggests the increase [5] S.H. Li, S.Q. Liu, J.C. Colmenares, Y.J. Xu, Green Chem. 18 (2016) 594–607.
in water leads to an increase in reactivity, and also that the dimers are [6] C. Li, X. Zhao, A. Wang, G.W. Huber, T. Zhang, Chem. Rev. 115 (2015)
converted into both stable monomers and unstable products that 11559–11624.
[7] M.P. Pandey, C.S. Kim, Chem. Eng. Technol. 34 (2011) 29–41.
participate in repolymerization reactions. [8] N. Sathitsuksanoh, K.M. Holtman, D.J. Yelle, T. Morgan, V. Stavila, J. Pelton,
By adding more water to the system, the average relative molecular H. Blanch, B.A. Simmons, A. George, Green Chem. 16 (2014) 1236–1247.
mass after milling with sodium hydroxide could be lowered until [9] M.V. Bule, A.H. Gao, B. Hiscox, S. Chen, J. Agric. Food Chem. 61 (2013)
3916–3925.
reaching a moisture content of 20 wt%, where the average relative
[10] J. Li, G. Henriksson, G. Gellerstedt, Bioresour. Technol. 98 (2007) 3061–3068.
molecular mass increased again. Once at 20 wt% moisture, the amount [11] A. Toledano, L. Serrano, J. Labidi, J. Chem. Technol. Biotechnol. 87 (2012)
of water in the system reached a point where the sample resembled a 1593–1599.
sludge rather than a powder, and effective grinding was no longer [12] K. Okuda, M. Umetsu, S. Takami, T. Adschiri, Fuel Process. Technol. 85 (2004)
803–813.
possible. Consequently, addition of excess water in the beginning of the [13] R.J. Gosselink, W. Teunissen, J.E. Van Dam, E. De Jong, G. Gellerstedt, E.L. Scott,
reaction is impractical and can result in an overall reduction of yields. J.P. Sanders, Bioresour. Technol. 106 (2012) 173–177.
However, it is possible that the gradual addition of water to the system, [14] M. Kleinert, J.R. Gasson, T. Barth, J. Anal. Appl. Pyrolysis 85 (2009) 108–117.
[15] L. Shuai, M.T. Amiri, Y.M. Questell-Santiago, F. Héroguel, Y. Li, H. Kim, R. Meilan,
such that the water in the system is never fully consumed, can ensure C. Chapple, J. Ralph, J.S. Luterbacher, Science 354 (2016) 329–333.
that plasticization does not limit the reaction. This will also likely [16] J. Zhu, Sustainable Production of Fuels, Chemicals, and Fibers from Forest Biomass,
alleviate repolymerization reactions that occur when water is depleted Am. Chem. Soc., Washington D.C, 2011, pp. 89–107.
[17] Q. Zhang, F. Jérôme, ChemSusChem 6 (2013) 2042–2044.
in the system. [18] M. Käldström, N. Meine, C. Farès, F. Schüth, R. Rinaldi, Green Chem. 16 (2014)
3528–3538.
5. Conclusions [19] N. Meine, R. Rinaldi, F. Schuth, ChemSusChem 5 (2012) 1449–1454.
[20] J. Hilgert, N. Meine, R. Rinaldi, F. Schüth, Energy Environ. Sci. 6 (2013) 92–96.
[21] S.M. Hick, C. Griebel, D.T. Restrepo, J.H. Truitt, E.J. Buker, C. Bylda, R.G. Blair,
Mechanochemistry offers a promising approach for the depolymer- Green Chem. 12 (2010) 468–474.
ization of lignin in a solvent-free process close to ambient conditions. In [22] R.G. Blair, Mechanical and combined chemical and mechanical treatment of
biomass, Production of Biofuels and Chemicals with Ultrasound, Springer, 2015, pp.
the presence of sodium hydroxide, the molecular mass of lignin
269–288.
fragments can be reduced substantially within 10 min of milling. [23] F. Schüth, R. Rinaldi, N. Meine, J. Hilgert, Method for obtaining sugar alcohols
However, the effectiveness of depolymerization is limited by the having five to six carbon atoms, US Patent 20,150,274,618, (2015).
formation of reactive intermediates that tend to undergo repolymeriza- [24] R. Carrasquillo-Flores, M. Käldström, F. Schüth, J.A. Dumesic, R. Rinaldi, ACS
Catal. 3 (2013) 993–997.
tion reactions (e.g., condensation) that form products, which can be [25] A.K. Deepa, P.L. Dhepe, ACS Catal. 5 (2014) 365–379.
more difficult to depolymerize than the original lignin. The addition of [26] T. Kleine, J. Buendia, C. Bolm, Green Chem. 15 (2013) 160–166.
methanol as a scavenger can prevent these reactions and allow for the [27] D. Lee, S. Tachibana, M. Sumimoto, Cellul. Chem. Technol. 22 (2) (1988) 201–210.
[28] D.-Y. Lee, M. Matsuoka, M. Sumimoto, Holzforschung-Int. J. Biol. Chem. Phys.
formation of stable monomeric products. Moisture can have a similar Technol. Wood 44 (1990) 415–418.
effect, but the addition of excess water results in plasticization, so that [29] D.-Y. Lee, M. Sumimoto, Holzforschung-Int. J. Biol. Chem. Phys. Technol. Wood 44
effective milling becomes impossible. Thus, controlled addition of (1990) 347–350.
[30] Z. Wu, M. Matsuoka, D. Lee, M. Sumimoto, J. Jpn. Wood Res. Soc. 37 (1991)
scavengers could be necessary to achieve the desired efficiency in 164–171.
mechanochemical processes for lignin conversion. Regardless of the

188
A.D. Brittain et al. Catalysis Today 302 (2018) 180–189

[31] Z.-H. Wu, M. Sumimoto, H. Tanaka, Holzforschung-Int. J. Biol. Chem. Phys. Press, 2010.
Technol. Wood 48 (1994) 395–399. [47] R. Samuel, Y. Pu, B. Raman, A.J. Ragauskas, Appl. Biochem. Biotechnol. 162 (2010)
[32] Z.-H. Wu, M. Sumimoto, H. Tanaka, Holzforschung-Int. J. Biol. Chem. Phys. 62–74.
Technol. Wood 48 (1994) 400–404. [48] T.-Q. Yuan, S.-N. Sun, F. Xu, R.-C. Sun, J. Agric. Food Chem. 59 (2011)
[33] S.L. James, C.J. Adams, C. Bolm, D. Braga, P. Collier, T. Friščić, F. Grepioni, 10604–10614.
K.D. Harris, G. Hyett, W. Jones, Chem. Soc. Rev. 41 (2012) 413–447. [49] R. Black, A. Rosen, S. Adams, J. Am. Chem. Soc. 75 (1953) 5344–5346.
[34] P. Baláž, Mechanochemistry in Nanoscience and Minerals Engineering, Springer [50] H. Ben, J.R. Ferrell III, RSC Adv. 6 (2016) 17567–17573.
Verlag, Berlin Heidelberg, 2008. [51] V. Roberts, S. Fendt, A.A. Lemonidou, X. Li, J.A. Lercher, Appl. Catal. B: Environ. 95
[35] P. Fox, J. Mat. Sci. 10 (1975) 340–360. (2010) 71–77.
[36] F. Schuth, R. Rinaldi, N. Meine, M. Kaldstrom, J. Hilgert, M.D.K. Rechulski, Catal. [52] P.C. Rodrigues Pinto, E.A. Borges da Silva, A.R.E.d. Rodrigues, Ind. Eng. Chem. Res.
Today 234 (2014) 24–30. 50 (2010) 741–748.
[37] L. Vanoye, M. Fanselow, J.D. Holbrey, M.P. Atkins, K.R. Seddon, Green Chem. 11 [53] Z. Yuan, S. Cheng, M. Leitch, C.C. Xu, Bioresour. Technol. 101 (2010) 9308–9313.
(2009) 390–396. [54] J. Miller, L. Evans, A. Littlewolf, D. Trudell, Fuel 78 (1999) 1363–1366.
[38] M. Shaw, L. Tabil, Agric. Eng. Int.: CIGR J. 9 (2007). [55] S.L. Zelinka, M.J. Lambrecht, S.V. Glass, A.C. Wiedenhoeft, D.J. Yelle, Thermochim.
[39] A. Shrotri, L.K. Lambert, A. Tanksale, J. Beltramini, Green Chem. 15 (2013) Acta 533 (2012) 39–45.
2761–2768. [56] F.S. Chakar, A.J. Ragauskas, Ind. Crops Prod. 20 (2004) 131–141.
[40] S. Maduskar, A.R. Teixeira, A.D. Paulsen, C. Krumm, T.J. Mountziaris, W. Fan, [57] M. Schwanninger, J. Rodrigues, H. Pereira, B. Hinterstoisser, Vib. Spectrosc. 36
P.J. Dauenhauer, Lab. Chip 15 (2015) 440–447. (2004) 23–40.
[41] L.J. Haupert, B.C. Owen, C.L. Marcum, T.M. Jarrell, C.J. Pulliam, L.M. Amundson, [58] F. Boissou, N. Sayoud, K. De Oliveira Vigier, A. Barakat, S. Marinkovic, B. Estrine,
P. Narra, M.S. Aqueel, T.H. Parsell, M.M. Abu-Omar, Fuel 95 (2012) 634–641. F. Jérôme, ChemSusChem 8 (19) (2015) 3263–3269.
[42] L. Zhang, L. Yan, Z. Wang, D.D. Laskar, M.S. Swita, J.R. Cort, B. Yang, Biotechnol. [59] Y. Dong, L. Schneider, T. Hu, M. Jaakkola, J. Holm, J.M. Leveque, U. Lassi, Biomass
Biofuels 8 (2015) 203. Bioenergy 86 (2016) 36–42.
[43] J.R. Ferrell, M.V. Olarte, E.D. Christensen, A.B. Padmaperuma, R.M. Connatser, [60] P. Dornath, H.J. Cho, A. Paulsen, P. Dauenhauer, W. Fan, Green Chem. 17 (2015)
F. Stankovikj, D. Meier, V. Paasikallio, Biofuels Bioprod. Biorefin. 10 (2016) 769–775.
496–507. [61] J.B.N. Aden, F. Tompkins, World Resour. Inst. (2013), http://pdf.wri.org/energy-
[44] O. Faix, B. Andersons, G. Zakis, Holzforschung-Int. J. Biol. Chem. Phys. Technol. efficiency-in-us-manufacturing-midwest-pulp-and-paper.pdf.
Wood 52 (1998) 268–274. [62] M.L. Bender, W.A. Glasson, J. Am. Chem. Soc. 81 (1959) 1590–1597.
[45] E.A. Capanema, M.Y. Balakshin, J.F. Kadla, J. Agric. Food Chem. 52 (2004) [63] A. Chesson, J. Sci. Food Agric. 32 (1981) 745–758.
1850–1860. [64] V.P. Heljo, A. Nordberg, M. Tenho, T. Virtanen, K. Jouppila, J. Salonen, S.L. Maunu,
[46] C. Heitner, D. Dimmel, J. Schmidt, Lignin and Lignans: Advances in Chemistry, CRC A.M. Juppo, Pharm. Res. 29 (2012) 2684–2697.

189

You might also like