You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/333156936

Cellulose Dissolution and Biomass Pretreatment Using Quaternary


Ammonium Ionic Liquids Prepared from H-, G-, and S-Type Lignin-Derived
Benzaldehydes and Dimethyl Carbonate

Article  in  Industrial & Engineering Chemistry Research · May 2019


DOI: 10.1021/acs.iecr.9b00640

CITATIONS READS

22 208

5 authors, including:

Victoria Diez Roland Kalb


Queens University of Charlotte proionic GmbH
1 PUBLICATION   22 CITATIONS    45 PUBLICATIONS   1,027 CITATIONS   

SEE PROFILE SEE PROFILE

David N Blauch Aaron M Socha


Davidson College Queens University of Charlotte
32 PUBLICATIONS   880 CITATIONS    21 PUBLICATIONS   800 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Illium Technologies, LLC View project

Next Generation Virucides View project

All content following this page was uploaded by Aaron M Socha on 23 June 2019.

The user has requested enhancement of the downloaded file.


Article

Cite This: Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX pubs.acs.org/IECR

Cellulose Dissolution and Biomass Pretreatment Using Quaternary


Ammonium Ionic Liquids Prepared from H-, G-, and S‑Type Lignin-
Derived Benzaldehydes and Dimethyl Carbonate
Victoria Diez,† Andrew DeWeese,† Roland S. Kalb,‡ David N. Blauch,§ and Aaron M. Socha*,†

Department of Environmental Science and Chemistry, Queens University of Charlotte, 1900 Selwyn Avenue, Charlotte, North
Carolina 28207, United States

Proionic, GmbH, Parkring 18, A-8074 Grambach, Austria
§
Department of Chemistry, Davidson College, 405 North Main Street, Davidson, North Carolina 28035, United States
Downloaded by LAWRENCE BERKELEY NATL LABORATORY at 13:16:19:405 on May 31, 2019

*
S Supporting Information

ABSTRACT: The synthesis of novel benzyl alkylammonium


ionic liquids (IL) from starting materials derived from grass,
softwood, and hardwood lignin was performed using batch
reactor systems followed by facile methyl carbonate ion
from https://pubs.acs.org/doi/10.1021/acs.iecr.9b00640.

exchange chemistry. Approximately 90% yields of the ILs were


achieved over three synthetic steps, and all novel materials
were fully characterized by 1H and 13C nuclear magnetic
resonance spectroscopy and high-resolution mass spectrom-
etry. The structure−activity relationships of the lignin-derived
ILs were evaluated for their ability to dissolve microcrystalline
cellulose and convert switchgrass to fermentable sugars
(glucose and xylose). All ILs tested, including a 1:1:1 mixture
of ILs prepared from vanillin, syringaldehyde, and p-
anisaldehyde yielded 52−71% of the glucose and 50−71% of the xylose from the total amount of glucan and xylan available
in the raw feedstock. These data compared well to the yields obtained from switchgrass pretreatment with the IL, 1-ethyl-3-
methyl imidazolium acetate, which gave 71% of the glucose and 63% of the xylose from the total glucan and xylan in the raw
feedstock, respectively.

1. INTRODUCTION Because of lignin’s potential to provide diverse and versatile


Found in all vascular plant cell walls, lignin is a structurally feedstocks for materials science10 and organic synthesis,11
complex natural product representing the world’s most chemical and biological lignin depolymerization have been
abundant renewable source of aromatic carbon. There are studied extensively.12−15 Several oxidative lignin depolymeriza-
several major types of oxidatively generated linkages in the tion reviews16,17 and protocols specific to benzaldehyde
lignin macromolecule,1 including β-aryl and phenolic ethers, production18,19 include reactor design20 and purification
carbon−carbon bonds, and benzyl esters, phenylglycoside, and strategies.19−21 For example, yields of 2.4 mol % benzaldehyde
acetal bonds between lignin and polysaccharides.2,3 Lignin from alkali lignin,22 7.4 wt % 4-hydroxybenaldehyde, 14.8 wt %
accounts for approximately 24−33, 19−28, and 15−25% of the 3-methoxy-4-hydroxy benzaldehyde (vanillin), 8.9 wt % 3,5-
dry weights of softwood, hardwood, and grasses, respectively.4 dimethoxy-4-hydroxybenzaldehyde (syringaldehyde) from
Within these plant types, the phenylpropanoid monomeric hardwood organosolv lignin,20 and 3.3 wt % syringaldehyde23
units comprising lignin macromolecules also vary in abundance from eucalyptus kraft lignin have been reported. Because of the
(Figure 1). Gymnosperms (e.g., softwoods such as pine and commercial importance of vanillin in food and fragrances,
spruce) have the most homogeneous lignin, with over 90% reports of lignin-derived vanillin have also appeared in
being made of coniferyl alcohol (G-type lignin).5 Angiosperm experimental reports from pulp mills24 and the Journal of
hardwood lignin, such as that of poplar, is made of Chemical Education.25
approximately 60% sinapyl alcohol (S-type lignin) and 40%
coniferyl alcohol.5 In addition to S- and G-type lignins, wild-
Special Issue: Biorenewable Energy and Chemicals
type grasses can contain 20−50% lignin as p-coumaryl alcohol
units (H-type lignin).6,7 The S/G/H lignin ratio of bioenergy Received: January 31, 2019
crops can now be genetically tuned, and modified lignin strains Revised: May 8, 2019
often require lower pretreatment severities for cellulose Accepted: May 16, 2019
separation and subsequent conversion to fermentable sugars.8,9 Published: May 16, 2019

© XXXX American Chemical Society A DOI: 10.1021/acs.iecr.9b00640


Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 1. Monomers comprising lignin found in all grasses, softwoods, and hardwoods (left) and benzaldehyde oxidative depolymerization
products (right) used to prepare ILs in this study.

Scheme 1. Three-Step Synthesis of Lignin-Derived Ionic Liquids (10−12) from Methylated Lignin-Derived Benzaldehydes

To integrate all products of a biorefinery, we sought to bond from cellulose hydroxyl groups, leading to its
develop a “closed-loop” process similar to the γ-valerolactone solubilization46 and decrystallization.47 To a lesser extent,
pretreatment strategy reported by Luterbacher et al.26 Instead cellulose dissolution depends on its degree of polymerization,
of γ-valerolactone as the pretreatment solvent, an alternative residence time, temperature, and concentration within the IL.
strategy is to use low-cost, biobased ionic liquids (IL) derived When holding the cation constant, it has been shown that
from lignin, hemicellulose, and/or amino acids.27−29 Ionic coordinating anions with greater hydrogen bond basicity have
liquids are defined as salts with loosely coordinated ions, a greater ability to dissolve cellulose, with a general trend of
leading to solvents with melting points less than 100 °C and diethylphosphate ≈ acetate > thioacetate > formate > chloride
negligible vapor pressures. Because of its favorable solvent > bromide.48 It is hypothesized that lignin-derived aprotic ILs
properties,30,31 1-ethyl-3-methyl-imidazolium acetate (EMIM with acetate anions will dissolve cellulose, hemicellulose, and
OAc, 1) can dissolve lignocellulosic biomass and is a well- lignin.
studied ionic liquid for biorefinery applications. Though To more thoroughly assess the cellulose dissolution and
effective, IL pretreatment is resource-intensive.32 At a cost of biomass pretreatment abilities of quaternary ammonium ILs
U.S. $20/kg, imidazolium ILs must be efficiently recycled and prepared from lignin-derived benzaldehydes, a structure−
dried for biomass processes. The cost estimate range of lignin- activity study is presented using three prototypical products
derived ILs is U.S. $5−15/kg.27 of lignin oxidation: p-anisaldehyde, vanillin, and syringalde-
IL pretreatment of lignocellulosic biomass typically dissolves hyde. These feedstocks underwent reductive amination as
hemicellulose and lignin so that a purified cellulose stream can previously reported and were subsequently N-methylated using
be recovered and easily hydrolyzed to monomeric sugars by dimethyl carbonate (Scheme 1). The carbonate-based ionic
cellulase and hemicellulase enzymes. Biomass applications liquid synthesis (CBILS) process was pioneered at Proionic,
using ionic liquids have recently been identified as a highly GmbH in 2005 and today is used to produce ILs on the 100
active research area,33 and commercial processes involving ton scale annually.49 As compared to traditional ionic liquid
ionic liquids and biomass have evolved beyond lignocellulosic syntheses, carbonate-based IL chemistry runs in closed loops
feedstocks.34 The first lignin-derived ionic liquids were tertiary and produces no waste or corrosive halogens in the final
amines prepared from the reductive amination of benzalde- product, and the ion exchange step proceeds with quantitative
hydes and furfurals, which were subsequently protonated with yields due to the strong thermodynamic equilibrium resulting
phosphoric acid.27,35 These protic ILs showed over 40 and from the loss of CO2.50 With lignin-derived quaternary
50% lignin and xylan removal, respectively,27 but because of ammonium methyl carbonate “platform ILs” in hand, facile
ion exchange reactions were used to provide the corresponding
their acidic nature, they did not dissolve cellulose. While protic
acetate ILs. We herein report the results of cellulose
ILs are useful in biorefining36,37 and are recoverable via
dissolution and switchgrass pretreatment using these novel,
distillation,38 they are not typically able to dissolve cellulose.
lignin-based, quaternary benzylammonium acetate ionic liquids
Protic ILs formed from superbases, such as those containing
in direct comparison to EMIM OAc.
guanidine and amidine cations, can dissolve approximately 10
wt % cellulose.39
Our work is motivated by the utility of biobased ILs as 2. EXPERIMENTAL SECTION
solvents for biopolymers, such as cellulose,40,41 lignin,42 2.1. Materials. All chemicals were purchased from Sigma-
chitin,43 and silk,44 and biopolymer applications, including Aldrich (97% pure or greater) and used without additional
fiber welding.45 Most of these applications require the IL to purification unless otherwise stated. N,N-Dimethylbenzylamine
dissolve the biopolymer so that the latter can be regenerated in (≥99% pure) was used to prepare the trimethyl benzylammo-
a new form. ILs with high Kamlet−Taft β values are nium methyl carbonate (5). Dimethyl carbonate was ≥99%
considered to be Brønsted bases and can accept a hydrogen pure. Acetonitrile (Acros, 99.8% pure) was dried with 4 Å
B DOI: 10.1021/acs.iecr.9b00640
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

molecular sieves. Avicel PH-101 cellulose (50 μm particle size) to prepare enzyme cocktails and dilute enzymatic saccha-
was used for cellulose dissolution experiments. 1-Ethyl-3- rification reactions.
methyl-imidazolium acetate was prepared from 1-ethyl-3- 2.3. Reductive Amination of Lignin-Derived Benzal-
methyl-imidazolium methyl carbonate (30 wt % in methanol, dehydes to Dimethyl Benzylamines (2−4). Example
Proionic, GmbH) by adding 1 mol equiv of acetic acid and protocol: 4-methoxybenzaldehyde (p-anisaldehyde, 12.00 g,
allowing to stir for 30 min before removing methanol in vacuo. 88.14 mmol, 1.0 equiv) and sodium triacetoxyborohydride
A 600 mL Parr reactor vessel (Parr Instrument Company, (26.15 g, 123.39 mmol, 1.4 equiv) were dissolved in 294 mL of
Moline, IL) was used to perform reductive amination anhydrous MeCN and added to the reactor vessel with stirring
reactions. CBILS reactions, IL melting-point analysis, cellulose at 200 rpm. At 33 °C, the vessel was evacuated to a pressure of
dissolution, and biomass pretreatment experiments were 400 Torr and then filled to atmospheric pressure (760 Torr)
separately performed in 15 mL (25.4 mm OD x 11.9 cm with dimethylamine. This cycle was repeated four additional
long) #15 Ace-thread pressure tubes with 4845 front-seal plugs times. The cylinder of dimethylamine was situated on a scale
(Ace Glass). The tubes were temperature controlled using a during this procedure to ensure MeCN and the reactor
Fisher Isotemp stirring hot plate by placing in 5 cm × 5 cm × headspace contained at least 1.2 equiv of dimethylamine (4.77
15 cm aluminum blocks drilled with three 2.5 cm deep × 2.6 g, 105.77 mmol). The reactor was heated to 50 °C and allowed
cm diameter holes for the pressure tube and a 2.5 cm deep × 5 to stir for 20 h. The reaction was quenched with approximately
mm diameter hole for the thermoresistor. 200 mL of DI water, acidified to pH 1.8 with 4 M HCl, and
2.2. Instrumental Methods. Nuclear magnetic resonance placed under reduced pressure to remove MeCN. The
(NMR) spectroscopy was performed on a Bruker AVANCE III remaining aqueous mixture was extracted with 3 × 75 mL of
HD spectrometer at frequencies of 400 MHz for 1H and 100 methyl-tert-butyl ether (MTBE) before the pH was increased
MHz for 13C. Samples dissolved in D2O contained acetone as to 11.1 with 4 M KOH. The latter was extracted with 2 × 75
an internal standard. 1H and 13C chemical shifts were mL of MTBE and dried over MgSO4. Solvent was removed
referenced to the acetone CH3 signals at 2.22 and 30.89 under reduced pressure to yield 1-(4-methoxyphenyl)-N,N-
ppm, respectively. For samples dissolved in CDCl3, 1H dimethylmethanamine (2) as an amber, nonviscous oil (13.30
chemical shifts were referenced to the residual proton signal g, 80.49 mmol, 91.3% yield). TLC: Rf of starting material/
at 7.26 ppm and 13C chemical shifts were referenced to the product = 0.6/0.2.
solvent signal at 77.16 ppm.51 High-resolution mass spectrom- 2.4. CBILS of Trimethyl Benzylammonium Methyl
etry (HR-MS) was performed on an Agilent accurate-mass Carbonates (5−8) from Dimethyl Benzylamines (2−4)
6520B Q-TOF mass spectrometer operating in positive-ion and Dimethyl Benzylamines. Example protocol: Com-
mode with 3500 V capillary and 120 V fragmenter voltages. pound 2 (1-(4-methoxyphenyl)-N,N-dimethylmethanamine,
The dual-spray electrospray ionization source was operated at 5.00 g, 30.26 mmol, 1.0 equiv) was dissolved in 4.68 g of
a nebulizer pressure of 40 psi and a 12 L/min flow of N2 drying methanol, and dimethyl carbonate (4.36 g, 48.42 mmol, 1.6
gas at 340 °C. The Agilent ESI-L tuning mix was used for equiv) was added. The reaction mixture was sparged with N2
calibration, and the ESI source continuously introduced a and then heated in an aluminum block to 140 °C for 16−24 h
reference solution containing ammonium trifluoroacetate, in a pressure tube. The reaction was monitored by HPLC, and
purine ([M + H]+ 121.0509), and hexakis (1H,1H,3H- upon complete disappearance of the starting material,
tetrafluoropropoxy)phosphazine ([M + H]+ 922.0098) during methanol and residual dimethyl carbonate were removed
measurements. Samples were dissolved in methanol or water under reduced pressure to yield 1-(4-methoxyphenyl)-N,N,N-
and introduced into the ion source by flow injection analysis trimethylmethanaminium (6) as a white crystalline solid (7.61
using a 50/50 mixture of methanol and 0.1% formic acid g, 29.81 mol, 98.5%). TLC: Rf of starting material/product =
flowing at 0.300 mL/min. GC/MS was performed on dimethyl 0.9/0.1.
benzylamines using a Thermo Trace 1300/ITQ 900 instru- 2.5. Ion Exchange of Trimethyl Benzylammonium
ment with a Thermo TG-SQC column (30 m, 25 mm, 0.25 Methyl Carbonates to Trimethyl Benzylammonium
μm) and a 15 °C/min temperature ramp from 75 °C (hold for Acetate Ionic Liquids. Example protocol: Compound 6 (1-
3 min) to 190 °C (hold for 2 min). The retention time of (4-methoxyphenyl)-N,N,N-trimethylmethanaminium, 5.00 g,
starting benzaldehydes was approximately 11.0 min, and that of 19.58 mmol, 1.00 equiv) was dissolved in 5 mL of methanol,
dimethyl benzylamine products was approximately 11.4 min. and glacial acetic acid (99.9% pure, 1.177 g, 19.58 mmol, 1.00
HPLC was performed on CBILS reactions using a Shimadzu equiv) was added dropwise while stirring at room temperature.
UFLC instrument with an SPD-M20A diode array detector After 1 h or the complete evolution of CO2, methanol was
using a mobile phase (1 mL/min) gradient of 15−65% removed under reduced pressure (0.3 Torr) at 75 °C to yield
methanol in water + 0.1% TFA in 12.0 min using a Restek 10 as a light-pink crystalline solid (4.68 g, 19.58 mmol, 99.9%).
Ultra C18 column (5 μm, 250 mm × 4.6 mm) with the 2.6. Cellulose Dissolution. Example protocol: Compound
column oven set at 40 °C. The retention time of starting 10 (536 mg) was added to a 15 mL pressure tube and heated
dimethyl benzylamines was approximately 6.5 min, and that of in a temperature-controlled aluminum block. Once the melting
trimethyl benzylammonium products was approximately 6.3 point was reached (130 °C), 20 mg of Avicel cellulose was
min. Thin layer chromatography (TLC) was performed on added. Constant temperature was maintained throughout the
Grace Reveleris aluminum-backed normal phase TLC plates experiment, and 10 mg aliquots of cellulose were added every
using a mobile phase of 2:1 hexane/EtOAc (+1% dihexylamine 5−10 min until saturation was observed. Table 1 summarizes
buffer) for dimethyl benzylamines and 4:2:1 hexane/EtOAc/ the melting points of each IL, which is also the temperature
methanol (+1% dihexylamine buffer) as the mobile phase for studied for cellulose dissolution, and the observed cellulose
dimethyl benzylammonium methyl carbonates. Citrate buffer solubility limits. Experiments were conducted in triplicate, and
(0.1 M ) was prepared to pH 5.2 in a 1.0 L volumetric flask average weight percentages and standard deviations are shown.
with 18.0 g of sodium citrate and 3.7 g of citric acid and used Compound 1 is a liquid at room temperature, but cellulose
C DOI: 10.1021/acs.iecr.9b00640
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Table 1. Cellulose Dissolution Experiments with Acetate ILs from nonpretreated switchgrass, and the results are presented
in Table 2.
IL temp (°C) cellulose solubility (wt %) standard deviation
2.9. Enzymatic Saccharification. Enzymatic saccharifica-
1 100 15 1.50 tion reactions were conducted on the individual pretreated
9 115 9 0.87 samples and water negative control following NREL LAP 9
10 130 10 0.35 “Enzymatic Saccharification of Lignocellulosic Biomass”. After
11 100 4 0.38 the final centrifugation cycle, the washed and decanted
12 110 3 0.75
biomass was suspended in 4.0 mL of citrate buffer
(approximately 10% solids loading for the original 400 mg of
biomass sample) and transferred to 15 mL falcon tubes. The
dissolution experiments were conducted at 100 °C to approach CTec2 cellulase enzyme blend (1.5 g, Sigma-Aldrich product
the temperatures used to test cellulose dissolution with lignin- number SAE0020) and 225 mg of hemicellulase enzymes from
derived ILs. Aspergillus niger (Sigma-Aldrich product number H2125) were
2.7. Biomass Pretreatment. Switchgrass (SG) was diluted in 0.1 M citrate buffer to a final volume of 10.0 mL.
obtained from Lawrence Berkeley National Laboratory and From the enzyme solution described above, 100 μL was added
found to be 33.5% glucan, 24.1% xylan, 5.3% arabinan, and to each saccharification experiment to achieve approximate
13.8% lignin as determined by the NREL method for cellulase and hemicellulase concentrations of 125 mg/g of
determining structural carbohydrates in biomass.52 Dry glucan and 25 mg/g of xylan, respectively. Saccharifications
switchgrass (400 mg) was mixed with 3.6 g of IL and 0.4 were placed horizontally on a Thermo Scientific Compact
mL of water to give 10 wt % biomass loading in 15 mL glass Digital Mini Rotator inside a Fisher Scientific Isotherm
pressure tubes. All pretreatment reactions were run in incubator and allowed to incubate (50 °C) with gentle mixing
duplicate, and an EMIM OAc (1) positive control was run (100 rpm). Aliquots of supernatant (1.0 mL) were sampled at
in triplicate. Tubular reactors were placed vertically in 48 and 70 h, centrifuged (14 700 rpm, 5 min, 23 °C), and
aluminum blocks and held at 160 °C for 3 h in a Thermo directly analyzed by HPLC for monosaccharide content as
Heratherm laboratory oven without stirring. After pretreat- described previously.53 Briefly, sugars were detected by the
ment, the reactors were allowed to cool to room temperature. refractive index using a Biorad Aminex 87-P column (300 ×
The mixture of IL, water, and pretreated biomass was 7.8 mm) with 100% water as the mobile phase. Five-point
transferred to a 50 mL falcon tube using DI water to a final calibration curves of glucose and xylose each had an R2 value of
volume of 45 mL and then centrifuged (22 °C, 5000 rpm, 15 0.99. The glucose and xylose yields presented in Figure 3 were
min) to separate the solid and liquid phases. The solid fraction calculated from the maximum possible glucan (e.g., 134.0 mg)
was sequentially washed and centrifuged with 40 mL of and xylan (e.g., 96.4 mg) in 400 mg of raw (untreated)
methanol, 3 × 40 mL of water (70 °C), and 40 mL of 0.1 M biomass.
citrate buffer (pH 5.2) to remove any residual IL and/or
sugars. Washed solids were either stored at 4 °C for 48 h or 3. RESULTS AND DISCUSSION
used directly in enzymatic saccharification. The water negative
control used 400 mg of SG and 4.0 mL of water and was held 3.1. Synthesis, NMR, and Mass Spectra of ILs. The
at 160 °C for 3 h. The 1:1:1 combined IL sample used 400 mg quaternization of amines 2−4 to yield methyl carbonates 6−8
of SG with 1.27 g of 10, 1.20 g of 11, 1.23 g of 12, and 0.4 mL results in the expected downfield shifts in the 1H NMR spectra
of water. owing to the strong diamagnetic effect of creating a positive
2.8. Mass Balance. Following pretreatment with com- charge on the amine nitrogen, while losing the nitrogen lone
pounds 1, 9, and 11, washed solids were dried in the laboratory pair. This downfield shift is strongest (Δδ 1.07 and 0.86 ppm)
oven (105 °C, 18 h) and weighed to determine the percent for the α methylene protons and α methyl protons,
solid recovery. A mass balance was then performed using respectively. NMR spectra of the free amines (2−4) and
standard analytical methods.52 To determine the extent of ionic liquids (5−12) were obtained in CDCl3 and D2O,
cellulose, hemicellulose, and lignin removal after IL pretreat- respectively, owing to solubility considerations. For a wide
ments, the recovered biomass was subjected to 72% H2SO4 range of compounds, changing solvents from CDCl3 to D2O
hydrolysis (30 °C, 60 min, 200 rpm stirring) followed by 4% results in less than a ±0.2 ppm change in the chemical shift.51
H2SO4 hydrolysis (121 °C, 60 min, autoclave). The samples The observed downfield shift of the α CH2 and CH3 sites
were filtered to quantify acid-insoluble lignin content and then resulting from quaternization is larger than the expected
subjected to HPLC (as described in Section 2.9) to quantify solvent effect. Downfield shifts for the benzene ring and
glucan and xylan contents. Lignin, glucan, and xylan methoxy groups are smaller and cannot be reliably
concentrations were compared to compositional analysis values distinguished from the solvent effects.

Table 2. Compositional Analysis of Switchgrass after Pretreatment with Selected Ionic Liquids As Compared to the
Nonpretreated Switchgrass (SG) Controla
IL pretreatment solid recovery, % glucan, % xylan, % arabinan % lignin, % xylan removal, % lignin removal, %
1 51.2 ± 1.5 46.0 ± 1.8 20.8 ± 2.0 4.7 ± 0.4 14.8 ± 0.1 51.3 ± 0.8 44.5 ± 1.1
9 46.4 ± 1.2 49.5 ± 0.6 25.3 ± 0.1 3.2 ± 0.1 13.6 ± 0.4 51.0 ± 0.1 57.5 ± 1.3
11 60.5 ± 3.5 42.0 ± 0.2 27.9 ± 0.3 4.5 ± 0.1 17.3 ± 1.4 23.3 ± 0.5 24.1 ± 6.6
untreated SG N/A 33.5 ± 0.5 24.1 ± 1.5 5.3 ± 0.2 13.8 ± 0.4 N/A N/A

a
± = standard deviation.

D DOI: 10.1021/acs.iecr.9b00640
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

1
H NMR spectra for 5−8 show no signals for the anion, 3.2. Cellulose Dissolution. Cellulose dissolution experi-
while the only anion signal in the 13C spectra appears at 161.0 ments showed that lignin-derived aprotic ILs, at their melting
ppm. NMR spectra of potassium methyl carbonate in D2O are points, can dissolve microcrystalline cellulose (Avicel) between
reported to show 1H peaks at 2.86 ppm and 13C peaks at 53.2 3 and 10 wt %. Compounds 9 and 10, derived from
and 160.2 ppm.54 For 5−8, the chemical shifts for the benzaldehyde and 4-methoxybenzaldehyde, respectively,
N(CH3)3 13C peaks are nearly identical to the expected showed the highest melting points and highest cellulose
location of the methyl carbonate CH3 signal. However, the 1H dissolution. Compound 10 dissolved 10 wt % cellulose, and
spectra for 5−8 show no peaks near 2.86 ppm that would be saturation was observed upon increasing the cellulose
consistent for the methyl carbonate CH3 group (Supporting concentration to 11.6 wt %. As compared to 9 and 10,
Information). The 13C chemical shift for NH4HCO3 in D2O compounds 11 and 12 showed the swelling of the cellulose
has been reported at 160.87 ppm55 and that for NaHCO3 has particles but less dissolution. When raising the temperature to
been reported at 161.20 ppm56 and 160.4 ppm.54 The 130 °C in cellulose dissolution experiments with 11, the
observed peak at 161.0 ppm for 5−8 is consistent with cellulose solubility did not increase beyond 5%. Therefore, it
hydrogen carbonate, but the methyl carbonate and carbonate seems that cation structure is more critical than temperature
ions also show signals close to this position. On the basis of the for cellulose dissolution in this aprotic, lignin-derived IL. It
NMR data, it seems possible that reaction or workup appears that increasing the degree of methoxylation decreases
conditions for 5−8 led to the hydrolysis of the methyl the cellulose solubility in these lignin-derived IL cations.
carbonate ion to the hydrogen carbonate ion. In contrast to previously reported biomass pretreatment
Ion-exchanged ionic liquids (9−12) show the expected 1H studies using protic ILs with the basic, coordinative acetate
and 13C signals for the acetate ion at 1.91 (CH3), 23.90 (CH3), anion38 and protic lignin-derived ILs with the acidic,
and 181.93 (COO) ppm as well as the absence of the 13C peak coordinative H2PO4− anion,27 the quaternary ammonium
at 161.0 ppm. All ionic liquids showed good purity by NMR compounds reported here carry the OAc− anion, and some
with the exception of 12, which contains residual acetic acid have the ability to dissolve cellulose. Protic ILs transfer a
from the ion-exchange process. For this compound, the proton from a Brønsted acid to a Brønsted base58 but typically
“acetate” CH3 1H signal integrates to 4H instead of the lack the requisite basicity to dissolve cellulose.59 However,
expected 3H, indicating 25 mol % or 6% (w/w) acetic acid in many protic ILs effectively pretreat biomass through their
the sample. As a solute in D2O, the acetate ion and acetic acid ability to dissolve and remove lignin,38,27 which renders the
will rapidly exchange hydrogen ions, leading to a single NMR cellulose more accessible for enzymatic hydrolysis to
signal for the CH3 site. The acetate signal is observed at 1.91 monomeric sugars.
ppm (9−11), and the reported chemical shift for acetic acid is 3.3. Mass Balance. A mass balance was performed with
2.08 ppm.51 For 25 mol % acetic acid, the methyl 1H signal is ILs 1, 9, and 11 using the aforementioned pretreatment
predicted to be 1.95 ppm; the observed chemical shift for 12 is conditions and showed 51.2, 46.4, and 60.5% solids recovery,
1.95 ppm. respectively (Table 2). The data for 1 and the untreated
ESI mass spectra of amines 2−4 displayed signals for the switchgrass controls were consistent with those previously
protonated amine, [M + H]+, while quaternary amines 5−8 reported.27 Glucan represented 49.5% of the solids recovered
and 9−12 gave signals for the M+ cation. The m/z values and from pretreatment with 9 and 42.0% of the solids recovered
isotopic distributions were in good agreement with the from pretreatment with 11. Compared to 33.5% glucan in the
expected chemical formulas. Attempts to prepare longer- untreated switchgrass, this represents 47.8 and 25.3% increases
chain amines (i.e., N,N-diethyl, N,N-dipropyl, and N,N- in glucan for 9 and 11, respectively. As anticipated from their
dihexyl) by reductive amination were successful, but these solubilization in the ILs, the amounts of lignin and xylan
tertiary benzylamines decomposed by Hoffman elimination decreased in the pretreated biomass. Compounds 9 and 11
reactions upon attempting N-methylation with dimethyl show 57.5 and 24.1% lignin removal, respectively. Xylan
removal was 51.0% for 9 and 23.3% for 11. Notably,
carbonate at 120−140 °C (MS data not shown). In our
pretreatment with IL 11 provided slightly less glucan in the
hands, N-ethylation with diethyl carbonate was also
recovered biomass and removed significantly less lignin and
unsuccessful because of a lack of reactivity. Our previous
xylan, as compared to 1 and 9. However, the higher solid
results showed the importance of phenol methylation on
recovery after pretreatment with 11 led to glucose and xylose
protic, lignin-derived cations.27 While the free phenols were
yields comparable to those of 1 and 9 after enzymatic
not tested in this study of quaternary alkylammonium acetate
saccharification (Figure 3).
salts, it is noteworthy that dimethyl carbonate has been
3.4. Biomass Pretreatment and Enzymatic Saccha-
reported as a “greener” alternative to dimethyl sulfate for the
rification. Overall, the lignin-derived ILs (9−12) performed
O-methylation of eugenol and vanillin.57 well, as compared to EMIM OAc (1), in terms of both glucose
and xylose yields (Figure 3). After pretreatment at 160 °C for 3
h followed by enzymatic saccharification, EMIM acetate
pretreatment provided 71% of the total possible glucose and
63% of the total possible xylose from raw switchgrass.
Trimethyl benzylammonium acetate (9) and the trimethyl
benzylammonium IL derived from vanillin (11) gave
comparable results with glucose yields of 69 and 70% and
xylose yields of 71 and 67%, respectively. The IL derived from
p-anisaldehyde (10) gave a glucose yield of 64% and a xylose
yield of 62%, and the IL derived from syringaldehyde (12)
Figure 2. Ionic liquids evaluated in this study. gave the lowest glucose and xylose yields of any of the ILs
E DOI: 10.1021/acs.iecr.9b00640
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

4. CONCLUSIONS
A series of novel quaternary ammonium ILs were synthesized
in high purity and high yield over three steps from
benzaldehydes derived from H-, G-, and S-type lignin. The
lignin-derived ILs prepared in this study showed the ability to
dissolve cellulose up to 10 wt % and pretreat biomass as
effectively as 1-ethyl-3-methyl imidazolium acetate for the
production of fermentable sugars (glucose and xylose). A
structure−activity relationship was observed whereby increas-
ing the methoxylation of the cation led to decreased
biopolymer solubility. This phenomenon was studied in
cellulose dissolution experiments of ILs 9−12, and observed
through lignin and xylan removal in mass balance experiments
with 9 and 11. The trend in cellulose dissolution did not
accurately predict the pretreatment efficacy, with the most
notable differences being with IL 11, derived from vanillin.
This compound showed only 4% cellulose dissolution but
provided the highest glucose yields of any lignin-derived ILs
(70%), as compared to that of 1 (71%). This finding is likely
Figure 3. Glucose and xylose yields obtained after pretreatment and due to the fact that 11 gave higher solids recovery and
70 h of enzymatic saccharification of switchgrass using individual comparable glucan recovery as compared to the other ILs
lignin-derived ILs (9-12), a 1:1:1 mixture of 10:11:12, and EMIM examined in mass balance experiments.
OAc (1). It is well known that IL viscosities can be lowered by
optimizing the N-alkyl chain length, branching, degree of
unsaturation,60 and incorporation of the ether functionality.
Future studies will be aimed at preparing lignin-derived ILs
tested. The decreased ability to dissolve cellulose and pretreat with lower viscosities by generating asymmetry via N-alkyl
switchgrass, as observed in 12, is most likely due to the 25% substituents. To fully investigate the properties, pretreatment
molar excess of acetic acid, as described in Section 3.1. mechanisms of action, and the overall utility of these new
Interestingly, lignin-derived ILs 9 and 11 had higher xylose lignin-derived ILs, Kamlet−Taft measurements, spectroscopic
yields as compared to 1, and this could result from greater studies including NMR of recovered lignin, and X-ray
lignin removal by 9 and lesser xylan removal by 11. diffraction along with glycome profiling of pretreated biomass
An experiment using a 1:1:1 mixture (by mass) of 10:11:12 are warranted.
gave 62% glucose and 61% xylose yields. While the glucose
yields obtained by pretreatment with these quaternary
5. SPECTRAL DATA
ammonium ILs are slightly lower than those reported using
protic lignin-derived ILs,27 it is likely due to the fact that the Compound 1 (1-ethyl-3-methyl-imidazolium acetate) clear
sugar yields reported from the protic ILs were calculated from liquid. 1H NMR: 1.49 (3H, t, J = 7.3 Hz), 1.90 (3H, s), 3.89
the total possible glucose available from glucan in the (3H, s), 4.22 (2H, q, J = 7.3 Hz), 7.42 (1H, s), 7.49 (1H, s),
pretreated biomass. In the current work, the glucose yields 8.72 (1H, s). 13C NMR: 15.09, 23.77, 36.19, 45.38, 122.48,
were calculated on the basis of the total amount of glucan 124.06, 136.18, 181.48. HR-MS: [C6H11N2]+, found 111.0919,
available in the untreated switchgrass. This assumption is calcd 111.0917, Δ = 2 ppm.
corroborated by the fact that switchgrass pretreated with Compound 2 (1-(4-methoxyphenyl)-N,N-dimethylmethan-
EMIM OAc also gave decreased glucose yields as compared to amine) clear liquid. 1H NMR: 2.22 (6H, s), 3.36 (2H, s), 3.78
previous results,27,28 which were calculated on the basis of the (3H, s), 6.85 (2H, d), 7.21 (2H, d). 13C NMR: 45.19, 55.29,
total glucan available after pretreatment. The water-pretreated 63.72, 113.68, 130.39, 130.76, 158.82; HR-MS: M =
switchgrass control gave 32% glucose and 20% xylose yields, C10H15NO, [M + H]+, found 166.1225, calcd 166.1225, Δ =
which represents a 1.60-fold-higher yield of glucose as 0.6 ppm.
compared to that of xylose. The IL-pretreated samples gave Compound 3 (1-(3,4-dimethoxyphenyl)-N,N-dimethyl-
an average of 1.04-fold higher glucose than xylose, also methanamine) clear liquid. 1H NMR: 2.21 (6H, s), 3.35
indicating that some glucan was lost during pretreatment. (2H, s), 3.84 (3H, s), 3.86 (3H, s), 6.78 (2H, s), 6.87 (1H, s).
13
Glucose and xylose yields were quantified at 48 h of enzyme C NMR: 45.25, 55.92, 64.18, 110.75, 112.12, 121.31, 131.28,
incubation and were approximately 4−6% lower for all ILs as 148.20, 148.97. HR-MS: C11H17NO2, [M + H]+, found
compared to respective yields after 70 h. There was no 196.1328, calcd 196.1332, Δ = 2 ppm.
statistical difference among the ILs tested. These data indicate Compound 4 (N,N-dimethyl-1-(3,4,5-trimethoxyphenyl)-
that the enzymatic hydrolysis rate reached a maximum in the methanamine) clear liquid. 1H NMR: 2.24 (6H, s), 3.35
first 48 h. To compare glucose yields to those observed after (2H, s), 3.82 (3H, s), 3.84 (6H, s), 6.54 (2H, s); 13C NMR:
pretreatment with aprotic lignin-derived ILs,27 glucose yields 45.43, 56.19, 60.92, 64.77, 105.84, 134.50, 137.03, 153.19; HR-
for ILs 1, 9, and 11 were also calculated from glucan recovery MS: C12H19NO3, [M + H]+, found 226.1435, calcd 226.1438,
(Table 2). As expected, glucose yields were found to be 97% Δ = 1 ppm.
for 1, 94% for 9, and 95% for 11. Standard deviations for the Compound 5 (N,N,N-trimethyl-1-phenylmethanaminium
aforementioned monosaccharide yields were between 2 and methyl carbonate) clear liquid. 1H NMR: 3.10 (9H, s), 3.35
7%. (2H, s), 4.48 (2H, s), 7.6 (5H, m). 13C NMR: 53.97 (t, J = 16
F DOI: 10.1021/acs.iecr.9b00640
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Hz), 70.23 (t, J = 10 Hz), 128.02, 129.81, 131.46, 133.43,


160.94. HR-MS: [C10H16N]+, found 150.1276, calcd 150.1277,
■ AUTHOR INFORMATION
Corresponding Author
Δ = 0.7 ppm. *E-mail: sochaa@queens.edu.
Compound 6 (1-(4-methoxyphenyl)-N,N,N-trimethyl-
ORCID
methanaminium methyl carbonate) clear liquid. 1H NMR:
3.06 (9H, s), 3.35 (2H, s), 3.87 (3H, s), 4.42 (2H, s), 7.10 Roland S. Kalb: 0000-0002-8462-4317
(2H, d, J = 8.2 Hz), 7.49 (2H, d, J = 8.2 Hz). 13C NMR: 52.69 Aaron M. Socha: 0000-0002-9976-3184
(t, J = 16 Hz), 56.11, 69.80 (t, J = 10 Hz), 115.15, 120.50, Notes
134.98, 160.95, 161.36. HR-MS: [C11H18NO]+, found The authors declare no competing financial interest.
180.1378, calcd 180.1383, Δ = 3 ppm.
Compound 7 (1-(3,4-dimethoxyphenyl)-N,N,N-trimethyl-
methanaminium methyl carbonate) clear liquid. 1H NMR:
■ ACKNOWLEDGMENTS
The LCMS used for mass spectrometry was acquired through
3.08 (9H, s), 3.87 (3H, s), 3.88 (3H, s), 4.41 (2H, s), 7.1 (3H, an NSF-MRI grant (CHE-1624377) awarded to D.N.B.
m). 13C NMR: 52.85 (t, J = 16 Hz), 56.35, 56.43, 70.15 (t, J = Student support for this project was made possible by an
10 Hz), 112.36, 116.18, 120.72, 127.07, 148.84, 150.70, NSF-ATE grant (1601636) awarded to A.M.S. The authors
160.96. HR-MS: [C12H20NO2]+, found 210.1490, calcd thank collaborators at Lawrence Berkeley National Laboratory
210.1489, Δ = 0.5 ppm. (Joint Bioenergy Institute) for the gift of switchgrass.
Compound 8 (N,N,N-trimethyl-1-(3,4,5-trimethoxyphenyl)-
methanaminium methyl carbonate) clear liquid. 1H NMR:
3.12 (9H, s), 3.82 (3H, s), 3.89 (6H, s), 4.42 (2H, s), 6.86
■ DEDICATION
This article is dedicated to Professor Yuzuru Shimizu.
(2H, s). 13C NMR: 53.14 (t, J = 16 Hz), 56.86, 61.58, 70.35,
111.02, 124.40, 139.25, 153.34, 160.97. HR-MS:
[C13H22NO3]+, found 240.1594, calcd 240.1594, Δ = 0 ppm.
■ REFERENCES
(1) Ralph, J.; Brunow, G.; Boerjan, W. Lignins. Encyclopedia of Life
Compound 9 (N,N,N-trimethyl-1-phenylmethanaminium Sciences; Wiley: New York, 2007.
acetate) clear liquid. 1H NMR: 1.91 (3H, s), 3.10 (9H, s), (2) Ralph, J.; Grabber, J. H.; Hatfield, R. D. Lignin-ferulate cross-
4.49 (2H, s), 7.6 (5H, m). 13C NMR: 23.87, 52.96 (t, J = 16 links in grasses: active incorporation of ferulate polysaccharide esters
Hz), 70.22 (t, J = 10 Hz), 128.01, 129.81, 131.46, 133.42, into ryegrass lignins. Carbohydr. Res. 1995, 275, 167.
181.85. HR-MS: [C10H16N]+, found 150.1274, 150.1277, Δ = (3) Koshijima, T.; Watanabe, T. Associations between Lignin and
2 ppm. Carbohydrates in Wood and Other Plant Tissues; Springer: Berlin, 2003.
Compound 10 (1-(4-methoxyphenyl)-N,N,N-trimethyl- (4) Strassberger, Z.; Tanase, S.; Rothenberg, G. The pros and cons
of lignin valorisation in an integrated biorefinery. RSC Adv. 2014, 4,
methanaminium acetate) clear liquid. 1H NMR: 1.91 (3H, 25310.
s), 3.06 (9H, s), 3.87 (3H, s), 4.43 (2H, s), 7.10 (2H, d, J = 8.2 (5) Tolbert, A.; Akinosho, H.; Khunsupat, R.; Naskar, A. K.;
Hz), 7.49 (2H, d, J = 8.2 Hz). 13C NMR: 23.90, 52.69 (t, J = Ragauskas, A. J. Characterization and analysis of the molecular weight
16 Hz), 69.80 (t, J = 10 Hz), 115.16, 120.49, 134.98, 161.37, of lignin. Biofuels, Bioprod. Biorefin. 2014, 8, 836.
181.93. HR-MS: [C11H18NO]+, found 180.1377, calcd (6) Jung, W.; Savithri, D.; Sharma-Shivappa, R.; Kolar, P. Changes in
180.1383, Δ = 3 ppm. lignin chemistry of switchgrass due to delignification by sodium
Compound 11 (1-(3,4-dimethoxyphenyl)-N,N,N-trimethyl- hydroxide pretreatment. Energies 2018, 11, 376.
methanaminium acetate) clear liquid. 1H NMR: 1.91 (3H, s), (7) Gonçalves, A. R.; Schuchardt, U.; Bianchi, M. L.; Curvelo, A. A.
3.08 (9H, s), 3.88 (3H, s), 3.90 (3H, s), 4.42 (2H, s), 7.1 (3H, S. Piassava fibers (Attalea f unifera): NMR spectroscopy of their lignin.
m). 13C NMR: 23.89, 52.84 (t, J = 16 Hz), 56.38, 56.47, 70.15 J. Braz. Chem. Soc. 2000, 11, 491.
(8) Chen, F.; Dixon, R. A. Lignin modification improves fermentable
(t, J = 10 Hz), 112.41, 116.27, 120.77, 127.10, 148.87, 150.74, sugar yields for biofuel production. Nat. Biotechnol. 2007, 25, 759.
182.02. HR-MS: [C12H11NO2]+, found 210.1489, calcd (9) Shi, J.; Pattathil, S.; Ramakrishnan, P.; Anderson, N.; Im Kim, J.;
210.1489, Δ = 0 ppm. Venkatachalam, S.; Hahn, M. G.; Chapple, C.; Simmons, B. A.; Singh,
Compound 12 (N,N,N-trimethyl-1-(3,4,5-trimethoxy- S. Impact of engineered lignin composition on biomass recalcitrance
phenyl)methanaminium acetate) clear liquid. 1H NMR: 1.95 and ionic liquid pretreatment efficiency. Green Chem. 2016, 18, 4884.
(4H, s), 3.12 (9H, s), 3.82 (3H, s), 3.89 (6H, s), 4.43 (2H, s), (10) Stewart, D. Lignin as a base material for materials applications:
6.87 (2H, s). 13C NMR: 23.19, 53.14 (t, J = 16 Hz), 56.86, chemistry, application and economics. Ind. Crops Prod. 2008, 27, 202.
61.59, 70.35, 111.03, 124.40, 139.26, 153.35, 180.83. HR-MS: (11) Li, C.; Zhao, X.; Wang, A.; Huber, G. W.; Zhang, T. Catalytic
[C13H22NO3]+, found 240.1596, calcd 240.1594, Δ = 1 ppm. transformation of lignin for the production of chemicals and fuels.


Chem. Rev. 2015, 115, 11559.
(12) Zakzeski, J.; Bruijnincx, P. C. A.; Jongerius, A. L.; Weckhuysen,
ASSOCIATED CONTENT B. M. The catalytic valorization of lignin for the production of
*
S Supporting Information renewable chemicals. Chem. Rev. 2010, 110, 3552.
The Supporting Information is available free of charge on the (13) Pandey, M. P.; Kim, C. S. Lignin depolymerization and
ACS Publications website at DOI: 10.1021/acs.iecr.9b00640. conversion: a review of thermochemical methods. Chem. Eng. Technol.
2011, 34, 29.
EMIM acetate, 1H and 13C NMR spectra of compound (14) Beckham, G. T.; Johnson, C. W.; Karp, E. M.; Salvachúa, D.;
1; benzylammonium, 1H and 13C NMR spectra of Vardon, D. R. Opportunities and challenges in biological lignin
compounds 5 and 9; 4-methoxy benzylammonium, 1H valorization. Curr. Opin. Biotechnol. 2016, 42, 40.
and 13C NMR spectra of compounds 2, 6, and 10; 3,4- (15) Sun, Z.; Fridrich, B.; de Santi, A.; Elangovan, S.; Barta, K.
Bright side of lignin depolymerization: toward new platform
dimethoxy benzylammonium, 1H and 13C NMR spectra
chemicals. Chem. Rev. 2018, 118, 614.
of compounds 3, 7, and 11; 3,4,5-trimethoxy benzy- (16) Chatel, G.; Rogers, R. D. Oxidation of lignin using ionic liquids
lammonium, 1H and 13C NMR spectra of compounds 4, - an innovatives to produce renewable chemicals. ACS Sustainable
8, and 12 (PDF) Chem. Eng. 2014, 2, 322.

G DOI: 10.1021/acs.iecr.9b00640
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

(17) Fache, M.; Boutevin, B.; Caillol, S. Vanillin production from one-pot process for the production of cellulosic ethanol. Energy
lignin and its use as a renewable chemical. ACS Sustainable Chem. Eng. Environ. Sci. 2016, 9, 1042.
2016, 4, 35. (38) Achinivu, E. C.; Howard, R. M.; Li, G.; Graczb, H.; Henderson,
(18) Voitl, T.; von Rohr, P. R. Demonstration of a process for the W. A. Lignin extraction from biomass with protic ionic liquids. Green
conversion of kraft lignin into vanillin and methyl vanillate by acidic Chem. 2014, 16, 1114.
oxidation in aqueous methanol. Ind. Eng. Chem. Res. 2010, 49, 520. (39) Nowicki, J.; Muszyński, M.; Mikkola, J.-P. Ionic liquids derived
(19) Mota, M. I. F.; Pinto, P. C. R.; Loureiro, J. M.; Rodrigues, A. E. from organosuperbases: en route to superionic liquids. RSC Adv.
Recovery of vanillin and syringaldehyde from lignin oxidation: a 2016, 6, 9194.
review of separation and purification processes. Sep. Purif. Rev. 2016, (40) Swatloski, R. P.; Spear, S. K.; Holbrey, J. D.; Rogers, R. D.
45, 227. Dissolution of cellose with ionic liquids. J. Am. Chem. Soc. 2002, 124,
(20) Liu, S.; Shi, Z.; Li, L.; Yu, S.; Xie, C.; Song, Z. Process of lignin 4974.
oxidation in ionic liquid coupled with separation. RSC Adv. 2013, 3, (41) Li, W.; Sun, N.; Rogers, R. D. Rapid dissolution of
5789. lignocellulosic biomass in ionic liquids using temperatures above
(21) Jones, T.; Finnan, J. L.; Arvizzigno, J. Process for separation of the glass transition of lignin. Green Chem. 2011, 13, 2038.
vanillin by means of azeotropic distillation with dibenzyl ether. U.S. (42) Brandt, A.; Grasvik, J.; Hallett, J. P.; Welton, T. Deconstruction
Patent 5,510,006, 1996. of lignocellulosic biomass with ionic liquids. Green Chem. 2013, 15,
(22) Wang, X.; Wang, N.; Nguyen, T. T.; Qian, E. W. Catalytic 550.
depolymerization of lignin in ionic liquid ssing a continuous flow (43) Qin, Y.; Lu, X.; Sun, N.; Rogers, R. D. Dissolution or extraction
fixed-bed reaction system. Ind. Eng. Chem. Res. 2018, 57, 16995. of crustacean shells using ionic liquids to obtain high molecular
(23) Tian, J.; Fu, S.; Lucia, L. A. Ionic liquid-based molecular oxygen weight purified chitin and direct production of chitin films and fibers.
oxidation of Eucalyptus kraft lignin to obtain a suite of monomeric Green Chem. 2010, 12, 968.
aromatic by-products. J. Wood Chem. Technol. 2015, 35, 280. (44) Phillips, D. M.; Drummy, L. F.; Conrady, D. G.; Fox, D. M.;
(24) Moodley, B.; Mulholland, D. A.; Brookes, H. C. The chemical Naik, R. R.; Stone, M. O.; Trulove, P. C.; De Long, H. C.; Mantz, R.
oxidation of lignin found in Sappi Saiccor dissolving pulp mill effluent. A. Dissolution and regeneration of Bombyx mori silk fibroin using
Water SA 2012, 38, DOI: 10.4314/wsa.v38i1.1. ionic liquids. J. Am. Chem. Soc. 2004, 126, 14350.
(25) Hocking, M. Vanillin: Synthetic flavoring from spent sulfite (45) DeLong, H. C.; Trulove, P. C.; Haverhals, L. M.; Reichert, W.
liquor. J. Chem. Educ. 1997, 74, 1055. M. Natural Fiber Welding. U.S. Patent 8,202,379, 2012.
(26) Luterbacher, J. S.; Rand, J. M.; Alonso, D. M.; Han, J.; (46) Brandt, A.; Hallett, J. P.; Leak, D. J.; Murphy, R. J.; Welton, T.
Youngquist, J. T.; Maravelias, C. T.; Pfleger, B. F.; Dumesic, J. A. The effect of the ionic liquid anion in the pretreatment of pine wood
Nonenzymatic sugar production from biomass using biomass-derived chips. Green Chem. 2010, 12, 672.
γ−valerolactone. Science 2014, 343, 277. (47) Cheng, G.; Varanasi, P.; Arora, R.; Stavila, V.; Simmons, B. A.;
(27) Socha, A. M.; Parthasarathi, R.; Shi, J.; Pattathil, S.; Whyte, D.; Kent, M. S.; Singh, S. Impact of ionic liquid pretreatment conditions
Bergeron, M.; George, A.; Tran, K.; Stavila, V.; Venkatachalam, S.; on cellulose crystalline structure using 1-ethyl-3-methylimidazolium
Hahn, M. G.; Simmons, B. A.; Singh, S. Efficient biomass acetate. J. Phys. Chem. B 2012, 116, 10049.
pretreatment using ionic liquids derived from lignin and hemi- (48) Chowdhury, Z. Z.; Zain, S. M.; Abd Hamid, S. B.; Khalid, K.
cellulose. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, E3587. Catalytic role of ionic liquids for dissolution and degradation of
(28) Sun, N.; Parthasarathi, R.; Socha, A. M.; Shi, J.; Zhang, S.; biomacromolecules. BioResources 2013, 9, 1787.
Stavila, V.; Sale, K. L.; Simmons, B. A.; Singh, S. Understanding (49) Kalb, R. S. CBILSs − 10 Years of Carbonate Based Ionic
pretreatment efficacy of four cholinium and imidazolium ionic liquids Liquids Synthesis; 6th International Congress on Ionic Liquids COIL-6,
by chemistry and computation. Green Chem. 2014, 16, 2546. June 16−20, 2015, Jeju, Korea.
(29) Hulsbosch, J.; De Vos, D. E.; Binnemans, K.; Ameloot, R. (50) Kalb, R. S.; Stepurko, E. N.; Emel’yanenko, V. N.; Verevkin, S.
Biobased ionic liquids: Solvents for a green processing industry. ACS P. Carbonate based ionic liquid synthesis (CBILS): thermodynamic
Sustainable Chem. Eng. 2016, 4, 2917. analysis. Phys. Chem. Chem. Phys. 2016, 18, 31904.
(30) Kamlet, M.; Abboud, J.; Taft, R. The solvatochromic (51) Gottlieb, H. E.; Kotlyar, V.; Nudelman, A. NMR Chemical
comparison method. 6. The.pi.* scale of solvent polarities. J. Am. Shifts of Common Laboratory Solvents as Trace Impurities. J. Org.
Chem. Soc. 1977, 99, 6027. Chem. 1997, 62, 7512.
(31) Ab Rani, M. A.; Brant, A.; Crowhurst, L.; Dolan, A.; Lui, M.; (52) Sluiter, A.; Hames, B.; Ruiz, R.; Scarlata, C.; Sluiter, J.;
Hassan, N. H.; Hallett, J. P.; Hunt, P. A.; Niedermeyer, H.; Perez- Templeton, D.; Crocker, D. Determination of Structural Carbohydrates
Arlandis, J. M.; Schrems, M.; Welton, T.; Wilding, R. Understanding and Lignin in Biomass: Laboratory Analytical Procedure; National
the polarity of ionic liquids. Phys. Chem. Chem. Phys. 2011, 13, 16831. Renewable Energy Laboratory: Golden, CO, 2008.
(32) Klein-Marcuschamer, D.; Simmons, B. A.; Blanch, H. W. (53) Shi, J.; Gladden, J. M.; Sathitsuksanoh, N.; Kamban, P.;
Techno-economic analysis of a lignocellulosic ethanol biorefinery with Sandoval, L.; Mitra, D.; Zhang, S.; George, A.; Singer, S. W.;
ionic liquids pretreatment. Biofuels, Bioprod. Biorefin. 2011, 5, 562. Simmons, B. A.; Singh, S. One-pot ionic liquid pretreatment and
(33) Welton, T. Ionic liquids: a brief history. Biophys. Rev. 2018, 10, saccharification of switchgrass. Green Chem. 2013, 15, 2579.
691. (54) Stueber, D.; Patterson, D.; Mayne, C. L.; Orendt, A. M.; Grant,
(34) Kalb, R. S.; Shiflett, M. Towards industrialization of ionic D. M.; Parry, R. W. Carbonates, Thiocarbonates, and the
liquids. Commercialization of Ionic Liquids; Springer; ahead of print. Corresponding Monoalkyl Derivatives. 1. Their Preparation and
(35) Socha, A. M.; Singh, S.; Simmons, B. A.; Bergeron, M. Synthesis Isotropic 13C NMR Chemical Shifts. Inorg. Chem. 2001, 40, 1902.
of novel ionic liquids from lignin-derived compounds. U.S. Patent (55) Mani, F.; Peruzzini, M.; Stoppioni, P. CO2 adsorption by
10,155,735. 2018. aqueous NH3 solutions: speciation of ammonia carbamate,
(36) George, A.; Brandt, A.; Tran, K.; Zahari, S. M. S. N.; Klein- bicarbonate and carbonate by a 13C NMR study. Green Chem.
Marcuschamer, D.; Sun, N.; Sathitsuksanoh, N.; Shi, J.; Stavila, V.; 2006, 8, 995.
Parthasarathi, R.; Singh, S.; Holmes, B. M.; Welton, T.; Simmons, B. (56) Pironti, C.; Cucciniello, R.; Camin, F.; Tonon, A.; Motta, O.;
A.; Hallett, J. P. Design of low-cost ionic liquids for lignocellulosic Proto, A. Determination of the 13C/12C Carbon Isotope Ratio in
biomass pretreatment. Green Chem. 2015, 17, 1728. Carbonates and Bicarbonates by 13C NMR Spectroscopy. Anal. Chem.
(37) Xu, F.; Sun, J.; Murthy Konda, N.; Shi, J.; Dutta, T.; Scown, C.; 2017, 89, 11413.
Simmons, B.; Singh, S. Transforming biomass conversion with ionic (57) Prakoso, N. I.; Pangestu, P. H.; Wahyuningsih, T. D. O-
liquids: process intensification and the development of a high-gravity, methylation of natural phenolic compounds based on green chemistry

H DOI: 10.1021/acs.iecr.9b00640
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

using dimethyl carbonate. IOP Conf. Ser.: Mater. Sci. Eng. 2016, 107,
012065.
(58) Greaves, T. L.; Drummond, C. J. Protic ionic liquids: Properties
and applications. Chem. Rev. 2008, 108, 206.
(59) Fukaya, Y.; Hayashi, K.; Wada, M.; Ohno, H. Cellulose
dissolution with polar ionic liquids under mild conditions: required
factors for anions. Green Chem. 2008, 10, 44.
(60) Murray, S. M.; O’Brien, R. A.; Mattson, K. M.; Ceccarelli, C.;
Sykora, R. E.; West, K. N.; Davis, J. H. The Fluid-Mosaic Model,
Homeoviscous Adaptation, and Ionic Liquids: Dramatic Lowering of
the Melting Point by Side-Chain Unsaturation. Angew. Chem., Int. Ed.
2010, 49, 2755.

I DOI: 10.1021/acs.iecr.9b00640
Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

View publication stats

You might also like