base-catalyzed depolymerization of biorefinery lignins · lignin can be obtained in two primary...

13
Base-Catalyzed Depolymerization of Biorenery Lignins Rui Katahira, Ashutosh Mittal, Kellene McKinney, Xiaowen Chen, Melvin P. Tucker, David K. Johnson, and Gregg T. Beckham* ,National Bioenergy Center, National Renewable Energy Laboratory, Golden, Colorado 80401, United States Biosciences Center, National Renewable Energy Laboratory, Golden, Colorado 80401, United States * S Supporting Information ABSTRACT: Lignocellulosic bioreneries will produce a substantial pool of lignin-enriched residues, which are currently slated to be burned for heat and power. Going forward, however, valorization strategies for residual solid lignin will be essential to the economic viability of modern bioreneries. To achieve these strategies, eective lignin depolymerization processes will be required that can convert specic lignin-enriched biorenery substrates into products of sucient value and market size. Base-catalyzed depolymeriza- tion (BCD) of lignin using sodium hydroxide and other basic media has been shown to be an eective depolymerization approach when using technical and isolated lignins relevant to the pulp and paper industry. To gain insights in the application of BCD to lignin-rich, biofuels-relevant residues, here we apply BCD with sodium hydroxide at two catalyst loadings and temperatures of 270, 300, and 330 °C for 40 min to residual biomass from typical and emerging biochemical conversion processes. We obtained mass balances for each fraction from BCD, and characterized the resulting aqueous and solid residues using gel permeation chromatography, NMR, and GCMS. When taken together, these results indicate that a signicant fraction (4578%) of the starting lignin-rich material can be depolymerized to low molecular weight, water-soluble species. The yield of the aqueous soluble fraction depends signicantly on biomass processing method used prior to BCD. Namely, dilute acid pretreatment results in lower water-soluble yields compared to biomass processing that involves no acid pretreatment. Also, we nd that the BCD product selectivity can be tuned with temperature to give higher yields of methoxyphenols at lower temperature, and a higher relative content of benzenediols with a greater extent of alkylation on the aromatic rings at higher temperature. Overall, this study shows that residual, lignin-rich biomass produced from conventional and emerging biochemical conversion processes can be depolymerized with sodium hydroxide to produce signicant yields of low molecular weight aromatics that potentially can be upgraded to fuels or chemicals. KEYWORDS: Dilute acid pretreatment, Biochemical conversion, Deacetylation, Kraft lignin, Lignin valorization, Lignin depolymerization INTRODUCTION Lignin is a heterogeneous, alkyl-aromatic polymer formed during plant cell wall synthesis from three phenyl-propanoid monomers connected by a variety of CO and CC linkages. In terrestrial plants, lignin comprises 15 to 30% of the cell wall, and plants utilize lignin for various functions including water and nutrient transport, defense against microbial pathogens, and strength and rigidity of plant tissues. 14 In Nature, lignin is partially broken down by the action of white-rot fungi and some ligninolytic bacteria, both of which employ powerful oxidative enzymes to cleave the diverse covalent linkages in lignin. 58 In a process context, lignin is currently produced worldwide in the pulp and paper industry in the form of Kraft or soda lignin. From these industrial streams, vanillin and lignosulfonates are the dominant products manufactured today. 1, 2, 9, 10 Going forward, the lignocellulosic biofuels industry, initially founded primarily for the production of ethanol from biomass-derived carbohydrates via fermentation, will begin to produce a vast amount of residual solid lignin. As a result, lignin valorization strategies targeting these biorenery lignin streams will become important for the economic viability and environmental sustainability of biofuels plants that utilize lignocellulose as a feedstock. 1,2,11 For many decades, studies have been conducted to develop eective depolymerization processes using either reductive or oxidative homogeneous and heterogeneous catalysts, lignino- lytic enzymes, whole-cell biocatalysts, and thermal depolyme- rization. Many of the developed strategies have focused on either technicallignin streams (e.g., ball-milled, dioxane- extracted lignin) or lignin from the pulp and paper industry (e.g., Kraft or soda lignin). The former is typically intended for the study of pureor isolated lignin whereas the latter is motivated by the need to valorize large, industrial scale streams that have been in production for many decades. However, the Received: November 6, 2015 Revised: January 8, 2016 Published: January 12, 2016 Research Article pubs.acs.org/journal/ascecg © 2016 American Chemical Society 1474 DOI: 10.1021/acssuschemeng.5b01451 ACS Sustainable Chem. Eng. 2016, 4, 14741486 This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Upload: others

Post on 14-May-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Base-Catalyzed Depolymerization of Biorefinery LigninsRui Katahira,† Ashutosh Mittal,‡ Kellene McKinney,† Xiaowen Chen,† Melvin P. Tucker,†

David K. Johnson,‡ and Gregg T. Beckham*,†

†National Bioenergy Center, National Renewable Energy Laboratory, Golden, Colorado 80401, United States‡Biosciences Center, National Renewable Energy Laboratory, Golden, Colorado 80401, United States

*S Supporting Information

ABSTRACT: Lignocellulosic biorefineries will produce asubstantial pool of lignin-enriched residues, which arecurrently slated to be burned for heat and power. Goingforward, however, valorization strategies for residual solidlignin will be essential to the economic viability of modernbiorefineries. To achieve these strategies, effective lignindepolymerization processes will be required that can convertspecific lignin-enriched biorefinery substrates into products ofsufficient value and market size. Base-catalyzed depolymeriza-tion (BCD) of lignin using sodium hydroxide and other basic media has been shown to be an effective depolymerizationapproach when using technical and isolated lignins relevant to the pulp and paper industry. To gain insights in the application ofBCD to lignin-rich, biofuels-relevant residues, here we apply BCD with sodium hydroxide at two catalyst loadings andtemperatures of 270, 300, and 330 °C for 40 min to residual biomass from typical and emerging biochemical conversionprocesses. We obtained mass balances for each fraction from BCD, and characterized the resulting aqueous and solid residuesusing gel permeation chromatography, NMR, and GC−MS. When taken together, these results indicate that a significant fraction(45−78%) of the starting lignin-rich material can be depolymerized to low molecular weight, water-soluble species. The yield ofthe aqueous soluble fraction depends significantly on biomass processing method used prior to BCD. Namely, dilute acidpretreatment results in lower water-soluble yields compared to biomass processing that involves no acid pretreatment. Also, wefind that the BCD product selectivity can be tuned with temperature to give higher yields of methoxyphenols at lowertemperature, and a higher relative content of benzenediols with a greater extent of alkylation on the aromatic rings at highertemperature. Overall, this study shows that residual, lignin-rich biomass produced from conventional and emerging biochemicalconversion processes can be depolymerized with sodium hydroxide to produce significant yields of low molecular weightaromatics that potentially can be upgraded to fuels or chemicals.

KEYWORDS: Dilute acid pretreatment, Biochemical conversion, Deacetylation, Kraft lignin, Lignin valorization,Lignin depolymerization

■ INTRODUCTION

Lignin is a heterogeneous, alkyl-aromatic polymer formedduring plant cell wall synthesis from three phenyl-propanoidmonomers connected by a variety of C−O and C−C linkages.In terrestrial plants, lignin comprises 15 to 30% of the cell wall,and plants utilize lignin for various functions including waterand nutrient transport, defense against microbial pathogens,and strength and rigidity of plant tissues.1−4 In Nature, lignin ispartially broken down by the action of white-rot fungi andsome ligninolytic bacteria, both of which employ powerfuloxidative enzymes to cleave the diverse covalent linkages inlignin.5−8 In a process context, lignin is currently producedworldwide in the pulp and paper industry in the form of Kraftor soda lignin. From these industrial streams, vanillin andlignosulfonates are the dominant products manufacturedtoday.1,2,9,10 Going forward, the lignocellulosic biofuelsindustry, initially founded primarily for the production ofethanol from biomass-derived carbohydrates via fermentation,will begin to produce a vast amount of residual solid lignin. As a

result, lignin valorization strategies targeting these biorefinerylignin streams will become important for the economic viabilityand environmental sustainability of biofuels plants that utilizelignocellulose as a feedstock.1,2,11

For many decades, studies have been conducted to developeffective depolymerization processes using either reductive oroxidative homogeneous and heterogeneous catalysts, lignino-lytic enzymes, whole-cell biocatalysts, and thermal depolyme-rization. Many of the developed strategies have focused oneither “technical” lignin streams (e.g., ball-milled, dioxane-extracted lignin) or lignin from the pulp and paper industry(e.g., Kraft or soda lignin). The former is typically intended forthe study of “pure” or isolated lignin whereas the latter ismotivated by the need to valorize large, industrial scale streamsthat have been in production for many decades. However, the

Received: November 6, 2015Revised: January 8, 2016Published: January 12, 2016

Research Article

pubs.acs.org/journal/ascecg

© 2016 American Chemical Society 1474 DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

This is an open access article published under an ACS AuthorChoice License, which permitscopying and redistribution of the article or any adaptations for non-commercial purposes.

chemical and physical properties of a “lignin” stream will oftensignificantly depend on the nature of the preprocessingconducted to isolate or enrich the lignin. These properties,for example, include the degree of polymerization, the identityand ratio of the chemical linkages (including new intermo-nomer linkages produced during processing), and the purity(related to the effects of residual carbohydrates, other plant cellwall components, or species introduced during the conversionprocess). The impact of processing will in turn affect the abilityof a given catalytic, thermal, or biological approach todepolymerize effectively lignin, such that a process that workswell on a given lignin stream will likely not be universallyeffective on all lignin streams. In this vein, the application ofdepolymerization strategies needs to be tailored to the process-relevant substrate(s) of interest.The need for tailoring lignin depolymerization strategies to a

process is especially acute for the growing biofuels industry,wherein lignin is currently being or slated to be burned for heatand power in most of the bioethanol plants operating orcoming online in the near future.1,3 Typical biochemicalconversion processes for bioethanol and advanced biofuelsemploy a mild thermochemical pretreatment step12,13 that isintended to disrupt the cell wall architecture (and in some casesremove lignin14−23) to make the polysaccharides moresusceptible to enzymatic depolymerization in a seconddeconstruction step through the use of (hemi)cellulolyticenzymes. The sugars produced from this two-step biomassdeconstruction strategy are then converted to ethanol viafermentation,24 or to a range of fuels or chemicals via biologicalor chemical catalysis approaches.25,26 As recently reviewed,1

lignin can be obtained in two primary points duringbiochemical conversion, namely during the pretreatment stepand/or as a lignin-enriched residual material followingpolysaccharide deconstruction. As the most prevalent pretreat-ment approaches do not extract appreciable amounts of lignin,lignin depolymerization research efforts adapted to residuallignin-enriched streams are thus a major priority in ligninvalorization.Current pretreatment approaches being used industrially

today mainly focus on the use of dilute-acid (e.g., sulfuric acid)pretreatment, hydrothermal pretreatment, or ammonia andalkaline-based pretreatment approaches.12,13 Dilute acid andhydrothermal environments have long been known to result inthe cleavage of prevalent C−O linkages in lignin (e.g., β-O-4linkages) followed by formation of more recalcitrant C−Clinkages.27 Recent work suggests that acid or hydrothermalenvironments are able to cleave C−O linkages from theterminal ends of the lignin polymer via an unzippingmechanism.28 The more “condensed” structure of dilute-acidor hydrothermal pretreated lignin is in turn a challengingsubstrate for subsequent depolymerization and upgrading.More recently, pretreatment processes have been developedwithout the use of acid that result in similar extents ofpolysaccharide deconstruction and sugar yields relative todilute-acid pretreatment.29,30 Specifically, Chen et al. evaluatedthe use of deacetylation, which consists of a mild alkalinetreatment with sodium hydroxide at 80 °C, and various formsof mechanical refining, the latter in lieu of dilute-acidpretreatment. The authors demonstrated that appreciablesugar yields are attainable via this approach. Given the absenceof high temperature and an acid catalyst, it is likely that thelignin from this process may be more amenable todepolymerization.

As mentioned above, many approaches have been employedto study lignin depolymerization. Reductive metal catalysis hasbeen reported as far back as the 1940s,31,32 and many recentstudies have been conducted using noble metal catalysts forlignin depolymerization.2,23,33−38 Oxidative transformations oflignin have also been reasonably well studied, historically in thecontext of aldehyde production, but more recently in a moregeneral context.39−42 Homogeneous alkaline catalysis has alsobeen a mainstay of lignin depolymerization research for quitesome time. Indeed, soda pulping is based on the use of sodiumhydroxide as a catalyst (oftentimes with anthraquinone as acocatalyst) to depolymerize lignin in whole biomass.20 Onlignin-enriched, solid substrates, the base-catalyzed depolyme-rization (BCD) process has been studied by multiple groups forover 2 decades. Thring conducted an early study using Alcelllignin with sodium hydroxide, and focused on liquid products,which were obtained in yields up to 30%.43 Later, Shabtai,Chornet, Johnson and co-workers published a series of studiesin which they describe a process employing BCD for lignindepolymerization followed by catalytic hydrodeoxygenation toproduce gasoline-range aromatic fuels.44−49 These studies wereconducted primarily on pulping lignins with a few experimentson dilute-acid pretreated substrates. A subset of the sameauthors developed a fractionation procedure to obtain severallignin-derived aromatic monomers from the BCD processthrough a series of extraction, distillation, chromatography, andcrystallization steps.48 In an additional set of studies, Miller andco-workers examined a variety of bases and showed thatstronger bases (e.g., KOH, NaOH) are able to produce higheryields of low molecular weight species than weaker bases, e.g.,LiOH.50,51 Roberts et al. published a more recent study whereinthey used boric acid as a capping agent to inhibitrepolymerization. It was not thoroughly discussed if boronwas chemically incorporated into the resulting lignin after theBCD reaction, which may cause challenges downstream incatalytic upgrading or in recycle of boron. Additional “capping”BCD experiments have been reported recently as well.Toledano et al. demonstrated the BCD concept on organosolvlignin from olive prunings and measured the oil yield (whichwas acidified to pH 1−2 after the BCD reaction andsubsequently extracted with ethyl acetate from the acidifiedliquid phase), which varied between 5 and 20%.52 Theysubsequently reported that the addition of phenol to the BCDreaction was able to dramatically improve the oil yield, providedthe phenol:lignin ratio was optimized.53

It is noteworthy that many of the aforementioned BCDstudies have been conducted to optimize the “oil” yield, whichis typically the organic solvent extractable material afteracidification. This approach is primarily motivated by the factthat the organic solvent extractable material will likely beamenable to catalytic hydrodeoxygenation, whereas the water-soluble fraction would not due to catalyst instability andpoisoning. Acidification will also precipitate any high molecularweight species. However, the downstream lignin valorizationstrategy will dictate the objective function of the depolymeriza-tion approach. Strategies are now emerging that are able toaccept heterogeneous mixtures of liquid-phase aromatic speciesas substrates for upgrading to value-added compounds based ona combined biological funneling and chemical catalysisapproach,54−57 such that the objective function will be totune the selectivity to specific product slates and maximize theconversion of lignin to low molecular weight species that arestable at biological pH (often near ∼7).

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1475

In light of the previous BCD work, and toward the ultimategoal of developing lignin depolymerization processes forbiorefinery lignins, here we systematically evaluate BCD onlignin-rich residues from three biorefinery processes namely,dilute-acid pretreatment followed by enzymatic hydrolysis(DAP-EH), deacetylation, disk refining, and enzymatichydrolysis (DDR-EH), deacetylation, twin-screw refining, andenzymatic hydrolysis (DTSR-EH), and two model lignins:Kraft lignin, and the lignin-enriched fraction from a CleanFractionation process.21 All lignin BCD experiments wereconducted in batch reactors for 40 min at either 2 or 4% NaOHconcentration with 10 wt % solids loadings at temperatures of270, 300, and 330 °C. We quantify the yield of liquid-phaseproducts, the molecular weight distributions of each resultingfraction (both the liquid and remaining solid fractions), andfingerprint prevalent species in the liquid-phase products.Overall, this study represents a comprehensive, quantitativestudy of BCD applied to process-relevant, biorefinery lignin-enriched substrates and demonstrates that BCD is able toproduce high yields of low molecular weight, aromatic speciesthat will likely be suitable for upgrading to value-addedcompounds through emerging lignin valorization pro-cesses.54−57

■ MATERIALS AND METHODSMaterials. Deacetylated/Disk refined/Enzymatic hydrolysis residue

from corn stover (DDR-EH):58 Corn stover was extracted with 0.1 M(0.4 wt %) NaOH at 80 °C for 2 h in a deacetylation step. Thedeacetylated corn stover was mechanically treated in a disk refiner(high consistency Sprout 36-in. commercial scale double disk refiner,model 401, and Durametal 36104 fine bar pattern segments used tomake up the rotating disk plates) followed by enzymatic hydrolysis.The deacetylated and disk-refined corn stover was enzymaticallyhydrolyzed at 15% solids loading at 50 °C for 5 days. NovozymesCellic CTec3 (60 mg protein/g cellulose) and HTec3 (40 mg protein/cellulose) enzyme preparations were added to the suspension. Wenote that these enzyme loadings are very high to efficiently hydrolyzepolysaccharides, but low enzyme loadings and high carbohydrate yieldshave been reported with this process.59 The enzymatically hydrolyzedslurry was repeatedly washed with DI water followed by centrifugation(10,000 rpm) to yield the enzymatic hydrolysis residue (DDR-EH).Deacetylated/Twin screw extruder refined/Enzymatic hydrolysis

residue from corn stover (DTSR-EH):58 Deacetylation of corn stoverwas performed by the same manner as that for the DDR-EH substrate.The deacetylated corn stover was ground in a twin screw extruder(Coperion ZSK-25 twin screw, corotating extruder) at 50 °C followedby enzymatic hydrolysis. Enzymatic hydrolysis was conducted at 20%solids loading under the same conditions as those for the DDR-EHsubstrate. The pretreated slurry was repeatedly washed with DI waterfollowed by centrifugation (10,000 rpm) to yield the enzymatichydrolysis residue (DTSR-EH).Dilute-acid pretreated/Enzymatic hydrolysis residue from corn

stover (DAP-EH): Pretreatment of whole corn stover was performedin a 1 ton/d continuous horizontal pretreatment reactor usingprocedures described previously.60 Corn stover samples wereimpregnated with 0.8% (w/w) sulfuric acid and pressed toapproximately 45%−50% total solids and then pretreated at 160 °Cfor 10 min. Pretreated corn stover was neutralized with NH4OH atroom temperature (22 °C) to pH 5.0. Neutralized, pretreated stoverwas enzymatically hydrolyzed at 20% solids loading at 50 °C for 120 hwith Novozymes Cellic CTec2 solution (20 mg protein/g cellulose).At completion of enzymatic hydrolysis, the slurry underwent solid−liquid separation using a Western States Q-80 (Fairfield, OH) basketcentrifuge with a 25 nm × 105 nm nominal pore size filtration cloth.The filter cake was resuspended in 10 L of DI water then recentrifugedin the same manner as the original slurry. This process was repeateduntil the glucose concentration in the supernatant was <1 g/L. The

washed, lignin-rich solids were collected and stored at 4 °C prior toanalysis and further processing.

Clean fractionation lignin-enriched (CF-LEF) residue from cornstover: Corn stover (10 g) in a single-phase mixture of methyl isobutylketone (MIBK)/acetone/H2O (11/44/44, g/g/g, 100 mL) withsulfuric acid (0.1 M) was heated at 140 °C for 56 min. The liquidfraction after filtration of the reaction mixture was phase-separated toobtain the lignin-enriched fraction (CF-LEF). Additional details of theCF-LEF protocol can be found in Katahira et al.21

Kraft lignin (KL) prepared from Norway spruce obtained fromSigma-Aldrich (Catalog #370959) was also used to evaluate theefficacy of BCD on a model lignin substrate that is currently availablecommercially.

BCD Experiments. BCD experiments were conducted in 50 mLstainless steel reactors. The experiments were conducted at threedifferent temperatures (270, 300, and 330 °C) at two NaOHconcentrations (2 and 4%) (Table 1). The autoclaves were heated

in a sand bath maintained at the target temperature where theyremained for the desired reaction period (40 min, once the desiredtemperature was reached). It took 8−12 min to preheat the reactor tothe desired reaction temperature. At the end of the reaction, thereactors were quenched to room temperature by immersion in coldwater. After removal from the water bath, the reactor was dried andweighed, opened to release the generated gas, and then reweighed.The difference of these two weights corresponds to the weight of thegas generated during the BCD experiment. Some part of the CO2 gasgenerated during BCD reaction may have reacted with NaOH to yieldNa2CO3, which would react with HCl during neutralization togenerate CO2. Gas yields in Figure 3 do not include the fraction of theCO2 absorbed by NaOH. The workup procedure was the same for allexperiments and is illustrated in Scheme 1. The reaction mixture wasfirst neutralized with HCl to pH 7 followed by centrifugation. Thisresulted in the separation of low molecular weight lignin products

Table 1. BCD Experimental Conditions

substratetemp(°C)

NaOH(wt %)

solidloading(wt %)

270 2 10Deacetylated Corn Stover/Disk Refined/EH residue [DDR-EH]

4 10

Deacetylated Corn Stover/Twin ScrewRefined/EH residue [DTSR-EH]

300 2 10

Dilute-acid Pretreated Corn Stover/EHresidue [DAP-EH]

4 10

Kraft lignin [KL] 330 2 104 10

Corn Stover CF-Lignin Enriched Fraction[CF-LEF]

300 4 10

Scheme 1. BCD Procedure Used in This Study forSeparation of the Aqueous and Solid Fractions

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1476

(supernatant−aqueous fraction) from the high molecular weight ligninproducts (solid residue).Gel Permeation Chromatography (GPC) Analysis. Starting

substrates and each freeze-dried BCD fraction (20 mg) were acetylatedin a mixture of pyridine (0.5 mL) and acetic anhydride (0.5 mL) at 40°C for 24h with stirring. The reaction was terminated by addition ofmethanol (0.2 mL). The acetylation solvents were then evaporatedfrom the samples at 40 °C under a stream of nitrogen gas. The sampleswere further dried in a vacuum oven at 40 °C overnight. A final dryingwas performed under vacuum (1 Torr) at room temperature for 1 h.The dried, acetylated lignin samples were dissolved in tetrahydrofuran(THF, Baker, HPLC grade). The dissolved samples were filtered (0.45μm nylon membrane syringe filters) before GPC analysis. Theacetylated samples were completely soluble in THF. GPC analysis wasperformed using an Agilent HPLC with 3 GPC columns (PolymerLaboratories, 300 × 7.5 mm) packed with polystyrene−divinylbenzenecopolymer gel (10 μm beads) having nominal pore diameters of 104,103, and 50 Å. The eluent was THF and the flow rate 1.0 mL/min. Aninjection volume of 25 μL was used. The HPLC was attached to adiode array detector measuring absorbance at 260 nm (bandwidth 80nm). Retention time was converted into molecular weight (Mw) byapplying a calibration curve established using polystyrene standards ofknown molecular weight (1 × 106 to 580 Da) plus toluene (92 Da).Gas Chromatography Mass Spectrometry (GC−MS) Anal-

ysis. Analysis was performed on an Agilent 6890N GC equipped witha 5973N MSD (Agilent Technologies, Palo Alto, CA). Samplecompounds were separated using a 30m × 0.25 mm × 0.25 mm HP-5MS column (122−5532, Agilent). HP MSD Chemstation software(Agilent) equipped with NIST11 database Rev. 2.0G (May 19, 2011build) was used to identify the unknown compounds found within thesamples. Aqueous fractions were freeze-dried and extracted withacetone. Each acetone-soluble sample was placed on an autosampler(Agilent) and injected at a volume of 10 μL into the GC−MS(Agilent). The GC−MS method consisted of a front inlet temperatureof 270 °C, MS transfer line temperature of 280 °C, and a scan rangefrom 35 to 550 m/z. A starting temperature of 35 °C was held for 3min and then ramped at 15 °C/min to a temperature of 225 °C with 1min hold time, then continued at a ramped rate of 15 °C/min to 300°C and held for 5 min. The carrier gas, He, was held at a constant flowof 1.0 mL/min. The method resulted in a run time of 27 min for eachsample.Solid-State NMR. High-resolution 13C CP/MAS NMR measure-

ments were carried out using a Bruker Avance 200 MHz spectrometeroperating at 50.13 MHz at room temperature. The spinning speed was7000 Hz, variable contact pulse, using a 50% ramp on the protonchannel of 2 ms, acquisition time 32.8 ms, and delay between pulses of1 s.Alkaline-Nitronbenzene Oxidation. DDR-EH and DAP-EH

were analyzed by alkaline-nitrobenzene oxidation.61,62 50 mg ofsubstrate was suspended in 40 mL of 2 M NaOH with 2.5 mL ofnitrobenzene. The mixture was heated at 170 °C for 2.5 h in a stainlesssteal reactor (Parr instrument company, Illinoia, USA). After cooldown of the reactor, the reaction mixture was extracted with CH3Clthree times. The aqueous layer was acidified with 2 N HCl, and thenextracted with diethyl ether three times. The internal standard(acetosyringone) was added to the reaction products, following whichthe entire mixture was silylated with BSTFA and analyzed by GC−MS.

■ RESULTS AND DISCUSSION

Characterization of Initial Substrates. In this work, wesystematically evaluated BCD on lignin-rich residues from threebiorefinery processes from corn stover: namely, dilute-acidpretreatment followed by enzymatic hydrolysis (DAP-EH),deacetylation, disk refining, and enzymatic hydrolysis (DDR-EH), deacetylation, twin-screw refining, and enzymatichydrolysis (DTSR-EH). Two model substrates were alsoexamined with the BCD process: Kraft lignin from spruceand organosolv lignin from a clean fractionation process applied

to corn stover (specifically the lignin-enriched fraction, dubbedCF-LEF).The compositions of these five substrates are shown in Table

2.63 The lignin content in DAP-EH is higher than those in

DDR-EH and DTSR-EH, while the xylan content in DAP-EHis lower than those in the other two EH residues. This isbecause hemicellulose is effectively hydrolyzed and removed inDAP. DDR-EH has more lignin and less glucan contents thanin DTSR-EH, suggesting that disk refining can increase thecellulose surface area for enhanced enzymatic hydrolysis.59 Ashcontent in DDR-EH is much lower that in DTSR-EH andDAP-EH. For DTSR-EH, this is due to the contamination ofmetals from the twin extruder. DAP-EH exhibits 13.3% ashcontent because the original ash in corn stover might beaccumulated and remain in DAP-EH; the extensive washing ofthe DDR-EH substrate likely removed significant amounts ofash.On the basis of the compositional analysis, these substrates

can be broadly divided into two main categories: (1) highlylignin-enriched substrates (KL and CF-LEF) with lignincontent >80% with minimal or no carbohydrates present and(2) process-relevant, biorefinery lignin-enriched substrates(DDR-EH, DTSR-EH and DAP-EH) with moderate to highlignin content and low to medium carbohydrate (glucan andxylan) content. The objective of selecting these substrates withvarying composition was to investigate the influence ofsubstrate composition and preprocessing on product yieldand quality (namely the yield of low molecular weight, aromaticspecies) that may be influenced by the presence of nonlignincomponents (i.e., glucan, xylan, ash) during BCD.

13C solid-state NMR spectra of the five BCD substrates areshown in Figure 1. DTSR-EH has less intensity in the aromaticcarbon peaks region at 109−165 ppm compared to DDR-EHand DAP-EH, which agrees with the lower lignin content ofDTSR-EH provided in the compositional analysis data. DAP-EH shows large lignin aromatic peaks (110−165 ppm) and alarge methoxyl peak at 56 ppm, and these aromatic peaks looksquite similar to the lignin aromatic peaks in the DDR-EH.Aromatic peaks from lignin are larger in DAP-EH because ofthe higher lignin content. C4/6 peaks from the crystalline andamorphous cellulose structure clearly appear, suggesting a largeportion of xylan was removed during the dilute-acid pretreat-ment process, in line with the results in Table 1. Both KL andCF-LEF have prominent lignin peaks with very small peaks forcarbohydrates.The molecular weight distributions of the lignin (which is the

primary component that absorbs in the UV range) in the fivelignin-rich BCD substrates are shown in Figure 2A and their

Table 2. Compositional Analysis Data for five BCDSubstrates (wt %)

DDR-EH DTSR-EH DAP-EH KL CF-LEF

Lignin 52.8 30.2 67.3 93.2 80.0Glucan 14.7 22.5 13.0 0.0 0.49Xylan 14.2 11.0 1.77 0.0 0.35Galactan 1.37 0.9 0.00 0.0 0.00Arabinan 2.86 2.2 0.38 0.0 0.68Acetate 0.41 0.3 0.80 4.0 0.25Ash 2.16 15.8 13.3 3.34 5.04Total sugar 33.1 36.5 15.2 0.0 1.5Total 88.5 82.8 96.5 101 86.9

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1477

weight-average Mw and polydispersity (PD) results are shownin Figure 2B. The lignin in each substrate exhibits a uniquemolecular weight distribution. DTSR-EH lignin has the highestMw of 16 100 Da with a small fraction of low Mw species (100−600 Da), whereas DDR-EH lignin exhibits the second highestMw of 9600 Da, suggesting that the refining method affects the

Mw of the residual lignin. The PD of DTSR-EH lignin is 12.4and that of DDR-EH is 7.2. DAP-EH lignin exhibits a muchlower Mw of 2900 Da compared to the lignin in the other twoenzymatic hydrolysis residues (Figure 2B), indicating that theether and/or ester bonds in corn stover lignin were cleaved to alarge extent during dilute-acid pretreatment. KL has relativelyhigh Mw of 5100 Da. CF-LEF has the lowest Mw (2000 Da)with the lowest polydispersity (Figure 2B) and the largestportion of low Mw species in the 200−500 Da range of the fivesubstrates (Figure 2A).

BCD Yields. Mass balances are key to the effectiveevaluation of any biomass conversion technology, thus amajor focus of the current study is aimed at obtaining effectivemass closure during the BCD reactions. The mass yields of theaqueous fraction, solid residue, and gas obtained after BCD areshown in Figure 3. The total mass yields for various substratesobtained in all six conditions ranged from 93 to 100% for DDR-EH (Figure 3A), 89−103% for DTSR-EH (Figure 3B), 94−102% for DAP-EH (Figure 3C), and 86−92% for KL (Figure3D), respectively. Overall, DDR-EH and DTSR-EH exhibithigher yields of aqueous fractions and lower solid residuescompared with the other three substrates. This suggests thatthe mechanically refined lignins are more native like and arethus more easily broken down; additionally, the lower lignincontents of the mechanically refined substrates may also lead tohigher yields of liquid species resulting from the residualcarbohydrates.The yield of solid residue decreased steadily with increasing

BCD temperature in both 2 and 4% NaOH. This trend wasobserved in all four substrates, except for BCD of KL at 300 °Cin 4% NaOH (Figure 3D). The gas yield increased steadily withan increase in the reaction temperature at both NaOHconcentrations for all four substrates, except for BCD ofDAP-EH at 300 °C in 4% NaOH (Figure 3C).The yields of the aqueous fraction from DDR-EH exhibit

only very slight trends with reaction temperature with the yieldincreasing at 300 and 330 °C with 4% NaOH; with 2% NaOH,the yields are virtually identical at 57−60%. With DTSR-EH

Figure 1. 13C CP-MAS solid-state NMR spectra of five substrates used in the BCD experiments. (A) DDR-EH, (B) DTSR-EH, (C) DAP-EH, (D)KL, and (E) CF-LEF.

Figure 2. Gel permeation chromatograms (A) and Mw and PD (B) ofthe five substrates.

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1478

Figure 3. Yield of different fractions obtained after BCD from four substrate (wt %). (A) DDR-EH, (B) DTSR-EH, (C) DAP-EH, and (D) KL.

Figure 4. Gel permeation chromatograms of the BCD aqueous fraction and solid residue from four substrates (300 °C in 4% NaOH). (A) DDR-EH,(B) DTSR-EH, (C) DAP-EH, and (D) Kraft lignin.

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1479

(Figure 3B), the gas yield increased (possibly due togasification of the carbohydrate components) and solid residueyield slightly decreased with increasing temperature. Theaqueous fraction decreased at higher temperature with both 2and 4% NaOH, in an opposite trend to that with DDR-EH and4% NaOH. Aqueous fraction yields from DTSR-EH with 4%NaOH are higher than those from DDR-EH, whereas the lignincontent in DTSR-EH (30%) is much lower than that in DDR-EH (53%) and the carbohydrate content of DTSR-EH isslightly higher (36.5%) than that of DDR-EH (33.1%) (Table2). These data suggest that some of the lignin and most of thecarbohydrates were solubilized in both DTSR-EH and DDR-EH lignins, and it is likely the carbohydrates were converted tocarboxylic acids.In DAP-EH, the yields of all three fractions exhibit a roughly

similar trend to that of DDR-EH: namely the yields of gas andaqueous fractions increase and the solid residues decrease withincreasing temperature. However, the aqueous fraction yieldfrom DAP-EH was much lower than that from DDR-EH andDTSR-EH and the solid residue yield was higher. This ispotentially because DAP-EH exhibits a lower initial carbohy-drate content or because the lignin in DAP-EH is morerefractory. The monomeric yield from DDR-EH after alkaline-nitrobenzene oxidation is 25.2 wt % based on the lignin contentof DDR-EH, which is higher than 20.6 wt % monomeric yieldfrom DAP-EH, suggesting that the lignin in DAP-EH exhibits amore condensed structure than DDR-EH. It is likely that thelower yield of the aqueous fraction from DAP-EH is alsobecause the more condensed lignin structure, generated duringdilute-acid pretreatment, exhibits greater resistance to alkalinetreatment. The increase in gas yields with increasing temper-ature suggests that some of the depolymerized compoundsfrom lignin and carbohydrates were broken down to gas-phasemolecules at higher temperature. The data also show, for theenzymatic hydrolysis residues (A, B, and C), the yields of theaqueous fraction in 4% NaOH are higher than those in 2%NaOH.For BCD of KL, the yields of solid fractions are much higher

than the yields of aqueous fraction compared to the other threesubstrates, which is also likely due to either the morecondensed structure of KL or its no carbohydrate content.The yields of the aqueous fraction and gas increased and thoseof the solid residue decreased with increasing reactiontemperature with 2% NaOH, whereas such obvious trends inthe solid residue and aqueous fraction were not observed with4% NaOH. For BCD of the CF-LEF (Figure S1 in SupportingInformation), the yields of aqueous fraction, solid residue, andgas were 48, 26, and 14%, respectively.GPC chromatograms are shown in Figure 4A−D for the

aqueous fraction and solid residue obtained from foursubstrates after BCD at 300 °C in 4% NaOH. As previouslynoted, the GPC chromatograms should primarily represent thearomatic, lignin-derived fraction. In all four substrates, theaqueous fractions and solid residue have much lower Mw thanthose of original substrates. It is interesting to note that despitea varying Mw distribution of the original substrates in DDR-EHand DTSR-EH (Figure 2), the Mw distribution of all aqueousfractions are quite similar showing a bimodal distribution, withpeaks at 300 and 500 Da. The average Mw of the aqueousfraction from DDR-EH is 660 Da and polydispersity is 1.7. TheMw distribution of the aqueous fraction from DTSR-EH issimilar to that of DDR-EH with the majority of productsexhibiting Mw between 200 and 2000 Da, whereas DAP-EH

exhibits a larger portion of products between 100 and 400 Daand lower overall Mw compared to the DDR-EH and DTSR-EH aqueous fractions, which could be attributed to the lowerMw of the original DAP-EH. The KL aqueous fraction containssome quite low Mw compounds in the range of 100−250.Clearly, the presence of low Mw compounds in the KL aqueousfraction shows the effectiveness of BCD in depolymerizing evenheavily processed lignin. It is surprising that some of thedepolymerized products in the 500−2000 Da range remainwater-soluble at pH 7, which is likely advantageous forupgrading these lignin depolymerization products to value-added compounds via biological approaches.54,57 In the solidfractions (Figure 4), all substrates also exhibit much lower Mwthan those of original substrates, and higher Mw distributionsthan those in the aqueous fraction. The Mw distribution fromeach BCD solid residue from DDR-EH, DTRS-EH, and DAP-EH are similar, in the range of 860−980 Da. The KL solidresidue is slightly higher Mw (1200 Da) than the other threesubstrates.To illustrate the effect of temperature and NaOH loading on

the resulting Mw distributions, GPC traces are shown in Figure5 for the aqueous fractions and solid residues obtained fromBCD of the DDR-EH residue at each temperature. At 270 °C(Figure 5A), the aqueous fraction obtained in 4% NaOHexhibits a slightly lower Mw distribution than that obtained in2% NaOH, suggesting that more lignin depolymerizationoccurs at higher NaOH concentration. Both chromatogramshave bimodal peaks, with low Mw species from 200 to 350 Daand higher MW species from 350 to 2000 Da. The former peakcorresponds to monomeric and dimeric compounds and thelatter to oligomers. The Mw distribution for the solid residueobtained in 4% NaOH also shows marginally lower Mwdistribution than that obtained in 2% NaOH, and both havepeaks at 700 Da. At 300 °C (Figure 5B), the aqueous fractionobtained in 2% NaOH contains more monomeric and dimericcompounds, whereas the aqueous fraction obtained in 4%NaOH exhibits a larger percentage of oligomeric products. Thisimplies that a larger amount of the higher Mw solid residue wasdegraded to oligomers in 4% NaOH. The chromatograms ofsolid residues obtained in both 2 and 4 wt % NaOH shift tolower Mw, suggesting greater thermal depolymerization oflignin at a higher reaction temperature. At 330 °C (Figure 5C),the trends continue in the same fashion. In the range of 100−3000 Da, the chromatograms of the aqueous fractions madewith 2% NaOH at 330 °C are similar to those made with 4%NaOH; however, there is a small peak near 10 000−20 000 Dawith 4% NaOH suggesting that some repolymerizationoccurred in 4% NaOH. The chromatograms of solid residuesobtained with both 2 and 4% NaOH are also similar to eachother. These GPC results clearly show that more depolyme-rization of the DDR-EH residue to monomeric and dimericcompounds occurs as the reaction temperature and NaOHconcentration increase. As shown in Figure 5D, higher severityBCD pretreatments tend to produce a lower Mw distributionand polydispersity with 2 wt% NaOH.Additionally, it is of interest to gain insights into product

selectivity by fingerprinting species in the BCD aqueousfractions as a function of NaOH loading and temperature. Tothis end, the BCD aqueous fractions from DDR-EH wereneutralized, freeze-dried, and then extracted with acetone.Acetone was used as the extraction solvent because most oflignin-derived monomeric compounds are soluble in acetone.The yields of the acetone extractives from DDR-EH were 2.7−

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1480

5.3% and 2.1−5.1% from DAP-EH. With 2% NaOH, bothacetone extractives increased with increasing temperature, butdecreased with 4% NaOH. GC-MS chromatograms of these

acetone extractives from DDR-EH are shown in Figure 6. Eachpeak was assigned and its relative abundance is summarized inTable 3. There are a small number of peaks from non-aromaticketones, 4-methyl-4-penten-2-one, 4-methyl-3-penten-2-one, 4-hydroxy-4-methyl-2-pentanone and 2,3-dimethyl-2-cyclopent-en-1-one (peak 1−3, 6). It is known that carbohydrates presentin biomass are degraded through peeling reactions in basicconditions, and a large portion of the hemicellulose is removedduring alkaline treatments.64 The non-aromatic ketones (peaks1−3, 6) are likely thus generated mainly from carbohydrates inDDR-EH. Their contents increased with increasing severitywith 2% NaOH, especially 4-methyl-3-penten-2-one (peak 2),because both the relative abundance of these non-aromaticketones and the yields of acetone extractives in 2% NaOHincreased with increasing temperature. The starting substrate,DDR-EH, had 14.7% glucan and 18.4% hemicellulose (Table2). It is likely that the carbohydrates in DDR-EH weredegraded to the four pentenone products and other low Mwdegradation products (i.e., most likely into carboxylic acids).Conversely, there are many peaks assigned to aromaticcompounds. Phenol (peak 5), guaiacol (peak 12), catechol(peak 17), syringol (peak 26), and acetosyringone (peak 33)are the main monomeric products at 270 °C based on peak area(Figures 6A-i and 6B-i). Their peak intensities decreased withincreasing reaction temperature except for catechol, whichexhibited a trend similar to that observed in BCD solution fromhemp wherein catechol intensity reaches a maximum at 300°C.65 The relative abundances of syringol (29.1%) andacetosyringone (14.7%) at 270 °C with 2% NaOH are muchlarger than those of guaiacol (6.79%) and acetovanillone(2.51%), suggesting that syringyl units are more easily releasedfrom lignin molecules than guaiacyl units under theseconditions. This is likely because syringyl units have one lessposition in the aromatic ring where formation of carbon−carbon bond condensation structures can occur than in guaiacylunits. A similar trend for peaks 12, 26, 32, and 33 was observedat 270 °C with 4% NaOH, but the syringol content was muchlower than that with 2% NaOH. The relative abundance ofcatechol (peak 17) content maximized at 300 °C with 2 and 4%NaOH (Table 3). Other benzenediol compounds 3-methyl-catechol (peak 22), 4-methylcatechol (peak 24), and 4-ethylcatechol (peak 28) were more prevalent at higher reactiontemperature. 4-Methylcatechol and 4-ethylcatechol are themain aromatic compounds in the BCD aqueous fraction inaddition to catechol, suggesting that 4-methylguaiacol (peak16) and 4-ethylguaiacol (peak 23) were converted to 4-methylcatechol and 4-ethylcatechol, respectively, throughdemethylation. These phenomena were observed using modelcompound studies with 4-methylguaiacol and 4-ethylguaiacol(Johnson D., unpublished). Many of the species present herewould be amenable to hydrodeoxygenation reactions orpotentially to be fed to an organism that could funnel theseheterogeneous aromatics into value-added compounds. Addi-tionally, 4-ethylguaiacol (peak 23) decreased while 4-ethyl-catechol and 4-ethylphenol (peak 14) increased. These resultssuggest that demethylation of the methoxyl group andfollowing dehydroxylation on the aromatic ring proceedsduring BCD. Additional demethylation of guaiacol also yieldscatechol (peak 17). The demethoxylation reaction was alsoobserved during BCD of hemp and softwood.65 Vanillic acid(peak 30) was also generated in small amounts at 270 °C,disappearing at the higher temperatures examined. Vanillin(peak 29), acetovanillone (peak 32), and acetosyringone (peak

Figure 5. Gel permeation chromatograms of aqueous fraction andsolid residue from DDR-EH after BCD with 2 and 4% NaOH at (A)270 °C, (B) 300 °C, and (C) 330 °C, and (D) Mw and PD of eachaqueous fraction. The dashed lines in panels A−C show 300 and 700Da.

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1481

33) were generated in small amounts at 270 and 300 °C, butdisappeared at 330 °C, suggesting that aldol condensationbetween aldehydes in vanillin and ketones in acetovanillone andacetosyringone may be occurring to generate dimers.To characterize further the reaction products formed during

BCD, the lumped peak areas of species categorized byfunctional group are summarized in Figure 7. We note thatanisole (peak 4), 1,2-dimethoxybenzene (veratrole, peak 10),and 1,2,3−trimethoxybenzene (peak 25) are not included inthe “total phenols”. The relative abundance of monomeric,depolymerized compounds derived from lignin (total lignin)decreased from 90.0% to 67.0% with increasing reactiontemperatures in both 2 and 4% NaOH, whereas the relativeabundance of non-aromatic ketones increased. This suggeststhat lignin depolymerization was predominant and proceedsrapidly under low severity BCD conditions and that conversionof both the lignin and sugar-derived species accelerates athigher temperature. This trend of non-aromatic ketonesindicates that residual carbohydrates in the DDR-EH substratewere converted to carboxylic acids via peeling reactions andthat these acids were subsequently converted to ketones viaketonization. The abundance of benzendiols such as catecholalso increases at higher severity, suggesting that demethylationof the methoxyl group in guaiacylic compounds occurs athigher severity BCD. The abundance of acetophenones,aldehydes, and acids decreases at higher temperature becausethey readily undergo condensation or decarboxylation reactions

with other degradation products due to their high reactivity.The relative abundance of p-hydroxyphenyl (H), guaiacyl (G),and syringyl (S) types are summarized in Figure 8. S-typeproducts are the largest at 270 °C whereas no S-type productwas detected at 330 °C. Both S- and G-derived productsdecreased dramatically with increasing reaction temperature,while H-type products decreased slightly. These data supportthe phenomenon of demethylation of methoxyl groups on thearomatic rings, leading to catechols that can further react.In a similar fashion, the relative abundance of monomeric

compounds from DTSR-EH, DAP-EH, and KL are shown inTables S1−S4 and Figures S2−S7, in Supproting Information.With DTSR-EH, the abundances of total lignin, acetophenones,and acids monotonically decrease and non-aromatic ketonesincrease with increasing reaction temperature (Table S1 andFigures S2, S3). The BCD products from DTSR-EH havehigher contents of non-aromatic ketones, compared to theproducts from DDR-EH, reflecting DTSR-EH’s high content ofresidual carbohydrates. The BCD products from DAP-EH(Table S2 and Figures S4, S5) have relative abundances oflignin-derived and carbohydrate-derived products that arebetween the levels seen for DDR-EH and DTSR-EH, whichis somewhat surprising given the low carbohydrate content ofDAP-EH, but possibly this reflects the more condensedstructure of the lignin in DAP-EH due to the use of acid inthe pretreatment, resulting in production of lower levels ofmonophenolic species.

Figure 6. GC−MS chromatograms of aqueous fractions from DDR-EH after BCD at (i) 270 °C, (ii) 300 °C, and (iii) 330 °C in (A) 2% NaOH and(B) 4% NaOH. Peak numbers are summarized in Table 3.

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1482

With KL, the relative abundance of lignin-derived productsdecreased to zero at 330 °C in 2% NaOH and at 300−330 °Cin 4% NaOH, whereas the abundance of non-aromatic ketones

increased dramatically at higher temperature (Table S3 andFigures S6, S7). Given that Kraft lignin does not containsubstantial carbohydrates (Table 2), this implies thatcompounds 1−2 in Table 3 are also generated from ligninand/or sugar degradation products, which were generatedduring the Kraft pulping process and contaminated in KL.

Table 3. Relative Abundance of Monomers Analyzed by GC−MS after Extraction of BCD Aqueous Fractions from DDR-EHwith Acetone

2% NaOH 4% NaOH

Peak # ID 270 °Ca 300 °Ca 330 °Ca 270 °Ca 300 °Ca 330 °Cb

1 4-methyl-4-penten-2-one 3.05 2.86 7.80 4.62 9.77 9.632 4-methyl-3-penten-2-one 4.62 5.84 19.5 7.32 17.1 15.63 4-hydroxy-4-methyl-2-pentanone 2.20 2.98 5.38 2.75 3.89 5.614 anisole 0.16 0.12 0.05 0.61 0.24 05 phenol 7.53 3.60 2.04 8.26 3.44 1.616 2,3-dimethyl-2-cyclopenten-1-one 0.12 0.18 0.31 0.19 0.21 0.247 2-methylphenol 0.08 0.16 0.22 0.14 0.21 0.338 acetophenone 0.02 0.10 0.05 0.01 0.01 0.429 2-ethylphenol 0.35 0.48 0.41 0.25 0.24 0.3510 veratrole 0.48 0.48 0.07 0.94 0.10 011 p-cresol 0.08 0.56 0.81 0.42 0.71 1.7012 guaiacol 6.79 2.00 0.26 6.36 0.73 0.2413 2,4-dimethylphenol 0.06 0.04 0.13 0 0.09 0.5914 4-ethylphenol 3.24 4.56 4.07 2.44 2.36 4.8815 2-methoxy-5-methylphenol 0.04 0.21 0.11 0.09 0.13 016 4-methylguaiacol 0.64 0.88 0.19 0.88 0.31 0.0317 catechol 8.02 32.9 19.0 15.7 26.2 12.918 4-isopropylphenol 0.20 0.39 0.44 0.16 0.20 0.6619 3-methoxyphenol 0.44 0.38 0 0.22 0.02 020 3,4-dimethoxyphenol 0.39 0.29 0 0.43 0 021 4-propylphenol 0.04 0.38 0.69 0.08 0.21 1.2322 3-methylcatechol 0 0 2.17 0 1.41 4.0823 4-ethylguaiacol 3.37 2.99 0.37 2.38 0.49 0.0624 4-methylcatechol 1.36 9.50 13.9 4.32 11.2 17.925 1,2,3-trimethoxybenzene 0.54 0.21 0 0.28 0 026 syringol 29.1 4.85 0 9.32 0.11 027 4-hydroxybenzaldehyde 0.27 0.23 0 0.21 0 028 4-ethylcatechol 3.44 16.9 20.9 8.74 16.0 19.229 vanillin 0.78 0.16 0 0.98 0 030 vanillic acid 4.38 1.48 0 2.12 0.01 031 4-hydroxyacetophenone 1.07 1.29 1.06 1.95 1.86 2.5832 acetovanillone 2.51 1.20 0 3.54 1.16 033 acetosyringone 14.3 1.52 0 13.9 1.43 034 2,2′-methylenebisphenol 0.32 0.19 0.03 0.40 0.09 0.11

total 100 100 100 100 99.9 100aThe values reported are the averages for the experiments conducted in triplicates. bIn duplicate.

Figure 7. Relative abundance of monomers in aqueous fractions fromDDR-EH under 6 BCD conditions analyzed by GC−MS. “Totallignin” includes all aromatic compounds. Anisole, 1,2-dimethoxyben-zene, and 1,2,3-trimethoxybenzene are not included in the “totalphenols”.

Figure 8. Relative abundance of monomers of p-hydroxyphenyl type(H), guaiacyl type (G), and syringyl type (S) in aqueous fractionsfrom DDR-EH under 6 BCD conditions analyzed by GC−MS.

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1483

Benzenediol, aldehydes, and acids were not detected in theBCD liquor from KL, and acetophenones were the mainproducts of lignin depolymerization. As expected the BCDaqueous fraction obtained from KL only contained guaiacyl-type lignin depolymerization products, as it was made from asoftwood (Norway Spruce), which contains only guaiacyl-typelignin.

13C-Solid-State NMR Analysis of BCD Solid Residues.The solid-state 13C NMR spectra of BCD solid residues fromfour of the substrates are shown in Figure 9. In DDR-EH (A),peaks from carbohydrates and oxygenated lignin side chains(60−109 ppm) decrease and disappear at 270 °C in 4% NaOH,whereas aliphatic carbon peaks (0−40 ppm) increase withhigher reaction severity. The methoxyl peak at 56 ppm alsodecreases and aromatic peaks in the range of 110−160 ppmincrease under severe conditions. These results indicate thatlignin and carbohydrates in DDR-EH residues were degradedfollowed by repolymerization to give high Mw products withhigh carbon content. Similar trends were observed in BCDsolid residues from the other three substrates.

■ CONCLUSIONS

Base-catalyzed depolymerization was applied to five lignin-richsubstrates, and the resulting mass balances of the aqueous,insoluble, and gas fractions were quantified. In all cases, themass balances are near total mass closure (86−103 wt %).DDR-EH and DTSR-EH exhibited higher liquid yields of 56−78 wt %, whereas DAP-EH, KL, and CF-LEF all exhibited loweryields of aqueous soluble fractions at all conditions. Thisdifference in the susceptibility to BCD between mechanicallyrefined substrates and acid-pretreated lignins as measured by

yield of an aqueous soluble fraction, is likely a consequence ofboth pretreatment method and the differing lignin andcarbohydrate contents of the substrates. Minimal structuralchanges in lignin and a higher carbohydrate content in asubstrate lead to higher liquid-phase yields after BCD. Theresulting molecular weight distributions in all cases of theaqueous fractions demonstrated a substantial amount of lignindepolymerization to monomeric and oligomeric species. TheBCD selectivity in all cases tended toward a higher content ofphenolic monomers at low severity, shifting to higher extents ofdealkylation as severity was increased. Going forward, base-catalyzed depolymerization of lignin-enriched biorefinerysubstrates may provide effective streams for either biologicalor catalytic upgrading strategies aimed toward lignin valor-ization.

■ ASSOCIATED CONTENT

*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acssusche-meng.5b01451.

Yields of BCD fraction from CF-LEF, Relativeabundances of individual and categorized monomericcompounds in aqueous fractions from DSTR-EH, DAP-EH, Kraft lignin, and CF-LEF. Relative abundances ofmonomers of p-hydroxyphenyl type, guaiacyl type, andsyringyl type in the aqueous fractions from thesesubstrates (PDF).

Figure 9. Solid-state 13C NMR spectra of BCD solid residues under six different conditions from the four substrates.

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1484

■ AUTHOR INFORMATIONCorresponding Author*G. Beckham. E-mail: [email protected] authors declare no competing financial interest.

■ ACKNOWLEDGMENTSWe thank the U.S. Department of Energy (DOE) BioenergyTechnologies Office (BETO) for funding this work. We thankErik Kuhn for supplying DAP-EH and William Michener forGC-MS analysis. The U.S. Government retains and thepublisher, by accepting the article for publication, acknowledgesthat the U.S. Government retains a nonexclusive, paid up,irrevocable, worldwide license to publish or reproduce thepublished form of this work, or allow others to do so, for U.S.Government purposes.

■ REFERENCES(1) Ragauskas, A. J.; Beckham, G. T.; Biddy, M. J.; Chandra, R.;Chen, F.; Davis, M. F.; Davison, B. H.; Dixon, R. A.; Gilna, P.; Keller,M. Lignin Valorization: Improving Lignin Processing in theBiorefinery. Science 2014, 344 (6185), 1246843.(2) Zakzeski, J.; Bruijnincx, P. C. A.; Jongerius, A. L.; Weckhuysen, B.M. The Catalytic Valorization of Lignin for the Production ofRenewable Chemicals. Chem. Rev. 2010, 110 (6), 3552−3599.(3) Chundawat, S. P.; Beckham, G. T.; Himmel, M. E.; Dale, B. E.Deconstruction of Lignocellulosic Biomass to Fuels and Chemicals.Annu. Rev. Chem. Biomol. Eng. 2011, 2, 121−145.(4) Boerjan, W.; Ralph, J.; Baucher, M. Lignin Biosynthesis. Annu.Rev. Plant Biol. 2003, 54 (1), 519−546.(5) Kirk, T.; Farrell, R. Enzymatic ″Combustion″: The MicrobialDegradation of Lignin. Annu. Rev. Microbiol. 1987, 41, 465−505.(6) Martinez, A. T.; Speranza, M.; Ruiz-Duenas, F. J.; Ferreira, P.;Camarero, S.; Guillen, F.; Martinez, M. J.; Gutierrez, A.; del Rio, J. C.Biodegradation of Lignocellulosics: Microbial, Chemical, andEnzymatic Aspects of the Fungal Attack of Lignin. Int. Microbiol.2005, 8 (3), 195−204.(7) Bugg, T. D.; Ahmad, M.; Hardiman, E. M.; Rahmanpour, R.Pathways for Degradation of Lignin in Bacteria and Fungi. Nat. Prod.Rep. 2011, 28 (12), 1883−96.(8) Bugg, T. D.; Ahmad, M.; Hardiman, E. M.; Singh, R. TheEmerging Role for Bacteria in Lignin Degradation and Bio-ProductFormation. Curr. Opin. Biotechnol. 2011, 22 (3), 394−400.(9) Kumar, M. N. S.; Mohanty, A. K.; Erickson, L.; Misra, M. Ligninand Its Applications with Polymers. J. Biobased Mater. Bioenergy 2009,3 (1), 1−24.(10) Rodrigues Pinto, P.; Borges da Silva, E.; Rodrigues, A. Lignin asSource of Fine Chemicals: Vanillin and Syringaldehyde. In BiomassConversion; Baskar, C.; Baskar, S.; Dhillon, R. S., Eds.; Springer: Berlin,2012; pp 381−420.(11) Davis, R.; Tao, L.; Tan, E.; Biddy, M. J.; Beckham, G. T.;Scarlata, C.; Jacobson, J.; Cafferty, K.; Ross, J.; Lukas, J.; Knorr, D.;Schoen, P. Process Design and Economics for the Conversion ofLignocellulosic Biomass to Hydrocarbons: Dilute-Acid Prehydrolysis andEnzymatic Hydrolysis Deconstruction of Biomass to Sugars and BiologicalConversion of Sugars to Hydrocarbons; Report NREL/TP-5100-60223;NREL: Golden, CO, 2013.(12) Mosier, N.; Wyman, C.; Dale, B.; Elander, R.; Lee, Y. Y.;Holtzapple, M.; Ladisch, M. Features of Promising Technologies forPretreatment of Lignocellulosic Biomass. Bioresour. Technol. 2005, 96(6), 673−686.(13) Karp, E. M.; Resch, M. G.; Donohoe, B. S.; Ciesielski, P. N.;O’Brien, M. H.; Nill, J. E.; Mittal, A.; Biddy, M. J.; Beckham, G. T.Alkaline Pretreatment of Switchgrass. ACS Sustainable Chem. Eng.2015, 3 (7), 1479−1491.(14) Li, C. L.; Knierim, B.; Manisseri, C.; Arora, R.; Scheller, H. V.;Auer, M.; Vogel, K. P.; Simmons, B. A.; Singh, S. Comparison of

Dilute Acid and Ionic Liquid Pretreatment of Switchgrass: BiomassRecalcitrance, Delignification and Enzymatic Saccharification. Bio-resour. Technol. 2010, 101 (13), 4900−4906.(15) Banerjee, G.; Car, S.; Scott-Craig, J. S.; Hodge, D. B.; Walton, J.D. Alkaline Peroxide Pretreatment of Corn Stover: Effects of Biomass,Peroxide, and Enzyme Loading and Composition on Yields of Glucoseand Xylose. Biotechnol. Biofuels 2011, 4, 16.(16) Bozell, J. J.; Black, S. K.; Myers, M.; Cahill, D.; Miller, W. P.;Park, S. Solvent Fractionation of Renewable Woody Feedstocks:Organosolv Generation of Biorefinery Process Streams for theProduction of Biobased Chemicals. Biomass Bioenergy 2011, 35 (10),4197−4208.(17) Bozell, J. J.; O’Lenick, C. J.; Warwick, S. Biomass Fractionationfor the Biorefinery: Heteronuclear Multiple Quantum Coherence-Nuclear Magnetic Resonance Investigation of Lignin Isolated fromSolvent Fractionation of Switchgrass. J. Agric. Food Chem. 2011, 59(17), 9232−9242.(18) Banerjee, G.; Car, S.; Liu, T. J.; Williams, D. L.; Meza, S. L.;Walton, J. D.; Hodge, D. B. Scale-up and Integration of AlkalineHydrogen Peroxide Pretreatment, Enzymatic Hydrolysis, andEthanolic Fermentation. Biotechnol. Bioeng. 2012, 109 (4), 922−931.(19) Chen, X. W.; Shekiro, J.; Franden, M. A.; Wang, W.; Zhang, M.;Kuhn, E.; Johnson, D. K.; Tucker, M. P. The Impacts of DeacetylationPrior to Dilute Acid Pretreatment on the Bioethanol Process.Biotechnol. Biofuels 2012, 5, 8.(20) Karp, E. M.; Donohoe, B. S.; O’Brien, M. H.; Ciesielski, P. N.;Mittal, A.; Biddy, M. J.; Beckham, G. T. Alkaline Pretreatment of CornStover: Bench-Scale Fractionation and Stream Characterization. ACSSustainable Chem. Eng. 2014, 2, 1481.(21) Katahira, R.; Mittal, A.; McKinney, K.; Ciesielski, P. N.;Donohoe, B. S.; Black, S.; Johnson, D. K.; Biddy, M. J.; Beckham, G. T.Evaluation of Clean Fractionation Pretreatment for the Production ofRenewable Fuels and Chemicals from Corn Stover. ACS SustainableChem. Eng. 2014, 2, 1364.(22) Resch, M. G.; Donohoe, B. S.; Ciesielski, P. N.; Nill, J. E.;Magnusson, L.; Himmel, M. E.; Mittal, A.; Katahira, R.; Biddy, M. J.;Beckham, G. T. Clean Fractionation Pretreatment Reduces EnzymeLoadings for Biomass Saccharification and Reveals the Mechanism ofFree and Cellulosomal Enzyme Synergy. ACS Sustainable Chem. Eng.2014, 2, 1377.(23) Parsell, T.; Yohe, S.; Degenstein, J.; Jarrell, T.; Klein, I.; Gencer,E.; Hewetson, B.; Hurt, M.; Kim, J. I.; Choudhari, H.; Saha, B.; Meilan,R.; Mosier, N.; Ribeiro, F.; Delgass, W. N.; Chapple, C.; Kenttamaa, H.I.; Agrawal, R.; Abu-Omar, M. M. A Synergistic Biorefinery Based onCatalytic Conversion of Lignin Prior to Cellulose Starting fromLignocellulosic Biomass. Green Chem. 2015, 17, 1492.(24) Himmel, M. E.; Ding, S. Y.; Johnson, D. K.; Adney, W. S.;Nimlos, M. R.; Brady, J. W.; Foust, T. D. Biomass Recalcitrance:Engineering Plants and Enzymes for Biofuels Production. Science 2007,315 (5813), 804−807.(25) Keasling, J. D.; Chou, H. Metabolic Engineering Delivers Next-Generation Biofuels. Nat. Biotechnol. 2008, 26 (3), 298−299.(26) Alonso, D. M.; Bond, J. Q.; Dumesic, J. A. Catalytic Conversionof Biomass to Biofuels. Green Chem. 2010, 12 (9), 1493−1513.(27) Moxley, G.; Gaspar, A. R.; Higgins, D.; Xu, H. StructuralChanges of Corn Stover Lignin During Acid Pretreatment. J. Ind.Microbiol. Biotechnol. 2012, 39 (9), 1289−1299.(28) Sturgeon, M. R.; Kim, S.; Lawrence, K.; Paton, R. S.; Chmely, S.C.; Nimlos, M.; Foust, T. D.; Beckham, G. T. A MechanisticInvestigation of Acid-Catalyzed Cleavage of Aryl-Ether Linkages:Implications for Lignin Depolymerization in Acidic Environments.ACS Sustainable Chem. Eng. 2014, 2 (3), 472−485.(29) Chen, X. W.; Tao, L.; Shekiro, J.; Mohaghaghi, A.; Decker, S.;Wang, W.; Smith, H.; Park, S.; Himmel, M. E.; Tucker, M. ImprovedEthanol Yield and Reduced Minimum Ethanol Selling Price (MESP)by Modifying Low Severity Dilute Acid Pretreatment withDeacetylation and Mechanical Refining: 1) Experimental. Biotechnol.Biofuels 2012, 5, 60.

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1485

(30) Tao, L.; Chen, X. W.; Aden, A.; Kuhn, E.; Himmel, M. E.;Tucker, M.; Franden, M. A. A.; Zhang, M.; Johnson, D. K.; Dowe, N.;Elander, R. T. Improved Ethanol Yield and Reduced MinimumEthanol Selling Price (MESP) by Modifying Low Severity Dilute AcidPretreatment with Deacetylation and Mechanical Refining: 2) Techno-Economic Analysis. Biotechnol. Biofuels 2012, 5, 69.(31) Brewer, C. P.; Cooke, L. M.; Hibbert, H. Studies on Lignin andRelated Compounds. Lxxxiv. The High Pressure Hydrogenation ofMaple Wood: Hydrol Lignin1. J. Am. Chem. Soc. 1948, 70 (1), 57−59.(32) Pepper, J. M.; Hibbert, H. Studies on Lignin and RelatedCompounds. Lxxxvii. High Pressure Hydrogenation of Maple Wood1.J. Am. Chem. Soc. 1948, 70 (1), 67−71.(33) Rinaldi, R.; Schuth, F. Design of Solid Catalysts for theConversion of Biomass. Energy Environ. Sci. 2009, 2 (6), 610−626.(34) Sturgeon, M. R.; O’Brien, M. H.; Ciesielski, P. N.; Katahira, R.;Kruger, J. S.; Chmely, S. C.; Hamlin, J.; Lawrence, K.; Hunsinger, G.B.; Foust, T. D. Lignin Depolymerisation by Nickel SupportedLayered-Double Hydroxide Catalysts. Green Chem. 2014, 16 (2), 824−835.(35) Hicks, J. C. Advances in C−O Bond Transformations in Lignin-Derived Compounds for Biofuels Production. J. Phys. Chem. Lett.2011, 2 (18), 2280−2287.(36) Wang, X.; Rinaldi, R. A Route for Lignin and Bio-OilConversion: Dehydroxylation of Phenols into Arenes by CatalyticTandem Reactions. Angew. Chem., Int. Ed. 2013, 52 (44), 11499−11503.(37) Wang, X.; Rinaldi, R. Solvent Effects on the Hydrogenolysis ofDiphenyl Ether with Raney Nickel and Their Implications for theConversion of Lignin. ChemSusChem 2012, 5 (8), 1455−1466.(38) Xu, C.; Arancon, R. A. D.; Labidi, J.; Luque, R. LigninDepolymerisation Strategies: Towards Valuable Chemicals and Fuels.Chem. Soc. Rev. 2014, 43 (22), 7485−7500.(39) Zakzeski, J.; Jongerius, A. L.; Weckhuysen, B. M. TransitionMetal Catalyzed Oxidation of Alcell Lignin, Soda Lignin, and LigninModel Compounds in Ionic Liquids. Green Chem. 2010, 12 (7),1225−1236.(40) Zakzeski, J.; Bruijnincx, P. C.; Weckhuysen, B. M. In SituSpectroscopic Investigation of the Cobalt-Catalyzed Oxidation ofLignin Model Compounds in Ionic Liquids. Green Chem. 2011, 13 (3),671−680.(41) Pandey, M. P.; Kim, C. S. Lignin Depolymerization andConversion: A Review of Thermochemical Methods. Chem. Eng.Technol. 2011, 34 (1), 29−41.(42) Das, L.; Kolar, P.; Sharma-Shivappa, R. Heterogeneous CatalyticOxidation of Lignin into Value-Added Chemicals. Biofuels 2012, 3 (2),155−166.(43) Thring, R. W. Alkaline-Degradation of Alcell(R) Lignin. BiomassBioenergy 1994, 7 (1−6), 125−130.(44) Chornet, E.; Shabtai, J. S.; Zmierczak, W. W. Process forConversion of Lignin to Reformulated Hydrocarbon Gasoline. U.S.Patent 5,959,167 A, September 28, 1999.(45) Johnson, D. K.; Chornet, E.; Zmierczak, W.; Shabtai, J.Conversion of Lignin into a Hydrocarbon Product for Blending withGasoline. Abstr. Pap. Am. Chem. Soc. 2002, 223, U583−U584.(46) Shabtai, J. S.; Zmierczak, W. W.; Chornet, E. Process forConversion of Lignin to Reformulated, Partially Oxygenated Gasoline.U.S. Patent 6,172,272 B1, January 9, 2001.(47) Shabtai, J. S.; Zmierczak, W. W.; Chornet, E.; Johnson, D.Process for Converting Lignins into a High Octane Additive. U.S.Patent 20,030,100,807, May 29, 2003.(48) Vigneault, A.; Johnson, D. K.; Chornet, E. Base-CatalyzedDepolymerization of Lignin: Separation of Monomers. Can. J. Chem.Eng. 2007, 85 (6), 906−916.(49) Vigneault, A.; Johnson, D.; Chornet, E. Advance in the ThermalDepolymerization of Lignin Via Base-Catalysis. In Science in Thermaland Chemical Biomass Conversion; CPL Press: Thatcham, U. K., 2006;pp 1401−1419.(50) Miller, J.; Evans, L.; Littlewolf, A.; Trudell, D. BatchMicroreactor Studies of Lignin and Lignin Model Compound

Depolymerization by Bases in Alcohol Solvents. Fuel 1999, 78 (11),1363−1366.(51) Miller, J. E.; Evans, L. R.; Mudd, J. E.; Brown, K. A. BatchMicroreactor Studies of Lignin Depolymerization by Bases. 2. AqueousSolvents; Report SAND2002-1318; Sandia National Laboratories:Albuquerque, NM, 2002.(52) Toledano, A.; Serrano, L.; Labidi, J. Organosolv LigninDepolymerization with Different Base Catalysts. J. Chem. Technol.Biotechnol. 2012, 87 (11), 1593−1599.(53) Toledano, A.; Serrano, L.; Labidi, J. Improving Base CatalyzedLignin Depolymerization by Avoiding Lignin Repolymerization. Fuel2014, 116, 617−624.(54) Linger, J. G.; Vardon, D. R.; Guarnieri, M. T.; Karp, E. M.;Hunsinger, G. B.; Franden, M. A.; Johnson, C. W.; Chupka, G.;Strathmann, T. J.; Pienkos, P. T. Lignin Valorization throughIntegrated Biological Funneling and Chemical Catalysis. Proc. Natl.Acad. Sci. U. S. A. 2014, 111 (33), 12013−12018.(55) Vardon, D. R.; Franden, M. A.; Johnson, C. W.; Karp, E. M.;Guarnieri, M. T.; Linger, J. G.; Salm, M. J.; Strathmann, T. J.;Beckham, G. T. Adipic Acid Production from Lignin. Energy Environ.Sci. 2015, 8, 617−628.(56) Johnson, C. W.; Beckham, G. T. Aromatic Catabolic PathwaySelection for Optimal Production of Pyruvate and Lactate from Lignin.Metab. Eng. 2015, 28, 240−247.(57) Salvachua, D.; Karp, E. M.; Nimlos, C. T.; Vardon, D. R.;Beckham, G. T. Towards Lignin Consolidated Bioprocessing:Simultaneous Lignin Depolymerization and Product Generation byBacteria. Green Chem. 2015, 17, 4951.(58) Chen, X. W.; Shekiro, J.; Elander, R.; Tucker, M. ImprovedXylan Hydrolysis of Corn Stover by Deacetylation with High SolidsDilute Acid Pretreatment. Ind. Eng. Chem. Res. 2012, 51 (1), 70−76.(59) Chen, X.; Shekiro, J.; Pschorn, T.; Sabourin, M.; Tao, L.;Elander, R.; Park, S.; Jennings, E.; Nelson, R.; Trass, O.; Flanegan, K.;Wang, W.; Himmel, M. E.; Johnson, D.; Tucker, M. P. A HighlyEfficient Dilute Alkali Deacetylation and Mechanical (Disc) RefiningProcess for the Conversion of Renewable Biomass to Lower CostSugars. Biotechnol. Biofuels 2014, 7, 98.(60) Schell, D.; Chapeaux, A.; Dowe, N.; Kuhn, E.; Shekiro, J.Demonstrate Integrated Pilot Plant Performance Meeting the 2012Conversion Targets; NREL: Golden, CO, 2012.(61) Katahira, R.; Nakatsubo, F. Determination of NitrobenzeneOxidation Products by Gc and H-1-Nmr Spectroscopy Using 5-Iodovanillin as a New Internal Standard. J. Wood Sci. 2001, 47 (5),378−382.(62) Esteves Costa, C. A.; Rodrigues Pinto, P. C.; Rodrigues, A. E.Evaluation of Chemical Processing Impact on E. Globulus WoodLignin and Comparison with Bark Lignin. Ind. Crops Prod. 2014, 61,479−491.(63) Sluiter, A. D.; Hames, B. R.; Ruiz, R. O.; Scarlata, C. J.; Sluiter, J.B.; Templeton, D. W.; Crocker, D. P. Determination of StructuralCarbohydrates and Lignin in Biomass; NREL: Golden, CO, 2006.(64) Green, J. W.; Pearl, I. A.; Hardacker, K. W.; Andrews, B. D.;Haigh, F. C. Peeling Reaction in Alkaline Pulping. Tappi 1977, 60(10), 120−125.(65) Lavoie, J.-M.; Bare, W.; Bilodeau, M. Depolymerization ofSteam-Treated Lignin for the Production of Green Chemicals.Bioresour. Technol. 2011, 102 (7), 4917−4920.

ACS Sustainable Chemistry & Engineering Research Article

DOI: 10.1021/acssuschemeng.5b01451ACS Sustainable Chem. Eng. 2016, 4, 1474−1486

1486