Comparison of Fast Pyrolysis Behavior of ... - ACS Publications

May 31, 2017 - Comparison of Fast Pyrolysis Behavior of Cornstover Lignins Isolated by Different ... Björkman method was broadly applied to isolated ...
0 downloads 0 Views 440KB Size
Subscriber access provided by Binghamton University | Libraries

Letter

Comparison of fast pyrolysis behavior of cornstover lignins isolated by different methods Jing Zhang, Kwang Ho Kim, Yong S Choi, Ali Hussain Motagamwala, James A. Dumesic, Robert C. Brown, and Brent H. Shanks ACS Sustainable Chem. Eng., Just Accepted Manuscript • Publication Date (Web): 31 May 2017 Downloaded from http://pubs.acs.org on June 4, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Sustainable Chemistry & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

1

Comparison of fast pyrolysis behavior of cornstover lignins isolated by different methods Jing Zhang,a Kwang Ho Kim,c Yong S. Choi,a Ali H. Motagamwala,de James A. Dumesic,bde Robert C. Brown,c and Brent H. Shanksab,∗ a

Department of Chemical and Biological Engineering, Iowa State University, 617 Bissell Road, Ames, IA 50011, USA b Center for Biorenewable Chemicals, Iowa State University, 617 Bissell Road, Ames, IA, 50011, United States. c Bioeconomy Institute, Iowa State University, 617 Bissell Road, Ames, IA 50011, USA d Department of Chemical and Biological Engineering, University of Wisconsin, 1415 Engineering Drive, Madison, WI, 53706, USA e Great Lakes Bioenergy Research Center, 1552 University Ave., Madison, WI, 53726, USA

Abstract The effect of lignin isolation methods on lignin fast pyrolysis was investigated utilizing an online product analysis methodology to give high mass balance closure for the products. Lignin samples with high purity were isolated from cornstover using three methods: Björkman milling, organosolv extraction and gamma-valerolactone (GVL) extraction. Although milled wood lignin has been extensively studied, characterization and the fast pyrolysis of milled cornstover lignin has not been previously reported. The current study shows the milled cornstover lignin contains a higher molecular weight and more aryl ether linkages relative to organosolv and GVL extracted cornstover lignin. During fast pyrolysis, considerably more phenolics, especially those with side chain double bonds derived from aryl ether cleavage, were formed from the milled cornstover lignin concomitant with less char formation.



Corresponding author. Address: Department of Chemical and Biological Engineering, Iowa State University, Ames, IA 50011, USA. Tel.: +1 515 294 1895. E-mail address: [email protected] (B.H. Shanks).

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2

Keywords Fast pyrolysis, Lignin characterization, Lignin isolation, Milled cornstover lignin, Phenolic compound

Introduction Lignin is an amorphous biopolymer consisting of phenylpropane monomer units with hydroxyl and methoxyl groups substituted on the aromatic ring.1,2 Unlike the uniform structure of cellulose and hemicellulose, lignin has a complicated structure since the number of methoxyl groups on the aromatic ring varies from zero to two and can be coupled in an almost random fashion. The three types of phenylpropane monomers, 4hydroxylphenylpropane (H), guaiacylpropane (G) and syringylpropane (S), correspond to the inclusion of zero, one or two methoxy groups on the phenylpropane ring, respectively. The content of H, G, S units in lignin is different among hardwood, softwood and herbaceous biomass.3,4 Due to intimate connection to cellulose and hemicellulose, lignin cannot be isolated from lignocellulose without altering its original structure. The Björkman method was broadly applied to isolated lignin from woody biomass, resulting in milled wood lignin that has been widely acknowledged to be the most closely representative of native lignin. Unfortunately, even this method causes some decomposition of the lignin polymer. Therefore, the structure of lignin depends on both the type of biomass and the manner in which it is isolated from lignocellulose. As the second most abundant component of lignocellulosic biomass, lignin has potential for producing fuels and chemicals. One possible deconstruction approach for converting lignin is fast pyrolysis. Several previous studies have examined the pyrolysis of different types of lignin. Wang et al. reported that the milled hardwood lignin produced

ACS Paragon Plus Environment

Page 2 of 16

Page 3 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

3

more methanol than milled softwood lignin because the former contained more methoxy groups.5 Wang et al. also reported that for a hardwood, lauan, the generated pyrolytic lignin had a lower molecular weight and initial thermal decomposition temperature than its milled wood lignin.6 However, both studies were performed based on thermogravimetric analysis (TGA), which could not provide the high heating rates that characterize fast pyrolysis. Importantly, they also did not quantify the products. Using a pyrolyzer with high heating rates, Obst found softwood lignin primarily produced G-type compounds, while hardwood lignin produced both S- and G-type compounds.7 Additionally, Evans et. al. reported vinyl phenol and coumaryl alcohol were the dominant volatile products from the fast pyrolysis of herbaceous lignin.8 However, these studies primarily focused on qualitative comparisons or, in a few instances, quantitative comparison of just a few products. Recently, Patwardhan et al. quantified a comprehensive product distribution for the fast pyrolysis of organosolv lignin form cornstover, but no other types of lignin were investigated.9 To our knowledge, there has been no systematic comparison of product distributions from the pyrolysis of cornstover lignins obtained by different isolation methods. We hypothesize that different isolation methods will alter the lignin structure to different degrees, thereby impacting the pyrolysis product distribution. Cornstover was chosen as the lignin source since it is the largest quantity of biomass residue in the United States.10 In the current study, we investigated three types of isolated cornstover lignin obtained from Björkman milling, gamma-valerolactone (GVL) solvent extraction, and organosolv processing. The GVL solvent extraction was recently explored as an approach to isolate native-like lignin from cornstover,11 while organosolv processing has been commercially used to produce lignin. The isolated lignins from these three processes are referred to in this work as milled

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

4

cornstover lignin (MCL), GVL extracted lignin (GVLL), and organosolv lignin (OSL) respectively.

Materials and Methods Lignin extraction Organosolv lignin from cornstover was procured from Archer Daniels Midland (ADM) Company. Milled cornstover lignin was prepared according to Björkman lignin method.12 This extraction was performed by solubilizing 100 g of cornstover powder in 1 L of methanol for 24 hours to remove extractives. The extractive-free cornstover was subjected to ball milling for 96 hours, followed by two separate extractions using 1 L of dioxane: water (96: 4, v: v) for 24 hours. The extracts were vacuum dried to obtain crude lignin. Further purification was performed by dissolving the crude lignin in 100 ml of 90% acetic acid followed by precipitation in ice water. The precipitated lignin was separated by centrifugation and freeze-drying. The lignin was then dissolved in 50 ml of 1,2dichloroethane: ethanol (2: 1, v: v) solution followed by centrifugation. The supernatant was added drop wise in 1 L of ethyl ether and centrifuged. The insoluble portion was freeze-dried and underwent ethyl ether washing three times to obtain the MCL. The GVLL was obtained by treating the cornstover in a batch reactor with a mixture of GVL, water and sulfuric acid at 120 °C for 30 min. After cooling, the slurry was filtered and washed with hot GVL. Then, DI water was added to the filtrate to precipitate crude lignin, followed by excessive DI water washing to obtain the GVLL. The experimental details has been reported elsewhere.11 Characterization of lignins

ACS Paragon Plus Environment

Page 4 of 16

Page 5 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

5

The amount of Klason lignin and acid-insoluble lignin were determined using the NREL LAP TP-510-42618 protocol.13 In the current study, the purity of MCL and OSL was taken as the sum of Klason lignin and acid soluble lignin. The purity of GVLL was qualitatively determined by NMR analysis as described by Luterbacher et al.11 A MettlerToledo Analytical (USA) thermogravimetric analyzer was used to determine the initial decomposition temperature, moisture and ash content for the lignin samples. The molecular weight of the lignin samples was determined by gel permeation chromatography (GPC) using a high-performance liquid chromatography system (Ultimate 3000, Dionex, Sunnyvale, USA) with a Shodex refractive index and diode array detector. Details about the TGA and GPC analyses are given in the supplementary materials. Fourier transform infrared (FTIR) spectrophotometry (Bruker Tensor 37, USA) was performed using dry KBr discs containing 1% of lignin powders. Each spectrum was recorded in the range from 4000 to 400 cm-1 with a resolution of 4 cm-1 over 20 scans. Background subtraction was applied to each spectrum. Fast pyrolysis of lignin The fast pyrolysis experiments were performed using a single-shot micropyrolyzer (Model 2020 iS, Frontier Laboratories, Japan). The micropyrolyzer was connected to a gas chromatograph (GC) system (7890A, Agilent Technologies, USA) for product identification and quantification. The GC system was equipped with three detectors including mass spectrometry (MS), flame ionization detector (FID) and thermal conductivity detector (TCD), which were used for product identification, condensable product quantification, and non-condensable product quantification, respectively. Details about the GC ramping program, product identification, product calibration and char

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

6

measurement can be found in previous publications.14,15 In the current study, all three lignin samples were pyrolyzed at four different temperatures (350 °C, 425 °C, 500 °C, 575 °C). Pyrolysis results are presented based on the average of triplicate trials.

Results and Discussion As shown in Table 1, the MCL and OSL samples were highly pure lignin. As indicated by NMR analysis, the GVLL was also a clean aromatic product except for some GVL residues.11 The residual GVL content was quantified by pyrolyzing it at a relatively low temperature. This revealed the GVL residue to be 17.5 wt% of the GVLL sample. Table 1 also shows that the three lignin samples had low ash and moisture content. Lignin purity was also indicated by the relative absence of pyrolysis products of non-lignin origin, as subsequently described. Table 1 shows the MCL sample had a much higher molecular weight than the OSL. Similar trends among isolated lignins from woody biomass have been reported.16 These results suggested that hydrolysis occurred to a lesser extent during Björkman milling than with organosolv processing. Table 1 shows that GVLL had a molecular weight more similar to that of the OSL than the MCL. Since NMR characterization indicated that GVLL had a lignin structure more closely resembling native lignin with little degradation other than limited β-ether cleavage,11 the molecular weight reduction for GVLL was likely due to cleavage of weak aryl ether bonds. The three delignification methods resulted in different degrees of depolymerization, as suggested by GPC analysis, which might be expected to lead to differences in the hydroxyl group and ether bond finctionalities. Fig. S1 shows the FTIR spectra of the three lignins with the magnified spectra for specific peaks shown in Fig. 1. As most clearly seen

ACS Paragon Plus Environment

Page 6 of 16

Page 7 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

7

in Fig. 1, a stronger signal at 1170 cm-1, attributed to aromatic ether linkage,17 was observed for MCL relative to the OSL and GVLL. Additionally, one peak at 1365 cm-1 was observed for the OSL, attributed to the phenolic hydroxyl group.18,19 A small shoulder at the same location was observed for the GVLL, as well. However, this peak was not apparent for MCL. This difference might be explained by hydrolytic breakage of the aromatic ether linkages leading to the formation of phenolic hydroxyl groups. As such, fragmentation of aromatic ether linkages, possibly the most sensitive β ether bonds,2 appeared to occur to a larger extent in the production of OSL and GVLL compared to MCL. Similar to woody biomass, the Bjorkman milling of cornstover appeared to preserve the native ether linkages in lignin to a larger extent thereby suppressing the generation of hydroxyl groups relative to the GVL extraction and organosolv process. TGA analysis results, which approximates the heating rate for slow pyrolysis, are shown in Fig. S2. The initial decomposition temperatures for MCL and OSL were nearly identical, at around 200 °C, suggesting the difference in molecular weight between MCL and OSL did not affect their thermal stability. The initial decomposition temperature for the GVLL was obscured by interference from residual GVL solvent in the sample. The char yield from TGA analysis was obtained using the sample weight before introducing the oxygen into the system (around 880 °C). For GVLL, the char yield was calculated on a GVL solvent-free basis. As shown in Table S1, similar char yields were observed for OSL and GVLL, both of which were higher than for MCL. Fast pyrolysis of the three lignin samples was performed using the Frontier micropyrolyzer in the temperatures range of 350 °C to 575 °C. As shown in Table S3 to S6, mass balances higher than 90% were achieved, which made possible quantitative

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

8

comparison of the product yields for the lignin samples. For GVLL, the residual solvent did not appear to decompose since 17.5 wt% GVL was observed among the products at all pyrolysis temperatures tested. To correct for the presence of this solvent, the pyrolysis product distribution of the GVLL was reported on a GVL solvent-free basis. For all the three lignins, the product yields not derived from lignin (grouped in Table S2) summed to less than 1%, which was further evidence of the high purity of the isolated lignins. Fig. 2 and Tables S3 to S6 compare the yields of char and total phenolic compounds produced from the three lignin samples for each pyrolysis temperature. Yields of phenolics, char and gas from the different lignins as a function of temperature are plotted in Fig. S3, which showed a decrease of char yield and increase of gas yield for all lignins from 350 °C to 575 °C while the yield of phenolics was maximized at 500 °C. Significantly higher yields of total phenolics were observed for MCL compared to GVLL and OSL at all pyrolysis temperatures. Concomitantly, a lower char yield was observed for the MCL relative to the GVLL and OSL, which was the same tend observed from TGA. The relative change in yields of phenolics as a function of temperature were similar for the MCL, OSL and GVLL, with the highest yield of total phenolics achieved at 500 °C for all cases. By dividing the phenolic products into phenolics with side chain double bonds (PWDB) and phenolics without side chain double bonds (PNDB) (grouped in Table S2), the main difference in the total phenolics yield between the MCL and OSL (or GVLL) was attributed to the higher yield of PWDB from the MCL, as seen in Fig. 2. Based on the aforementioned lignin characterization, the major difference between MCL and OSL/GVLL was the molecular weights. Ether bonds, possibly the most fragile β-ether bonds, were cleaved to a greater

ACS Paragon Plus Environment

Page 8 of 16

Page 9 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

9

extent during organosolv and GVL extraction compared to Björkman milling. Ether bonds surviving the Björkman milling process would readily cleave during fast pyrolysis. Our previous work with lignin model compounds provided insights into the β-ether bond cleavage during fast pyrolysis.20 Pyrolysis of 2-phenoxyphenylethanol (PPE), the simplest β-O-4 compound, primarily formed acetophenone, benzeneacetaldehyde and phenol, the first two of which are PWDB accounting for 60 wt% of total pyrolysis products.20 Similarly, pyrolysis of 2-phenoxyphenyl-1,3-propanediol (PPPD) primarily formed acetophenone, benzeneacetaldehyde, benzaldehyde and phenol, among which PWDB accounted for up to 50 wt% of the yield.20 Based on the product distribution, it was proposed that pyrolysis involved concerted reactions in which β-O-4 linkages were initially cleaved by a retro-ene reaction through a cyclic transition state followed by keto-enol tautomerization, demethylation, and hydroxyl group migration. In contrast, hydrolytic cleavage of β-O-4 bonds during the lignin isolation process could introduce hydroxyl groups instead of side chain double bonds. These hydroxyl groups decrease the volatility of the pyrolytic products and increase the tendency of intermolecular dehydration thereby facilitating char formation. For example, during pyrolysis of a p-hydroxy substituted PPPD, 1-(4-hydroxyphenyl)-2-phenoxypropane-1,3-diol (HH), 34% of the carbon appeared as char. In contrast, no noticeable char formation occurred for PPPD pyrolysis.20 The only difference between the HH and PPPD was that the former has one more phenolic hydroxyl group. Moreover, significantly less char was formed from native lignocellulose in which hemicellulose had been extracted compared to a physical mixture of cellulose and OSL with comparable amounts of cellulose and lignin.15 Since both kinds of cellulose-lignin mixtures were essentially free of minerals and pure cellulose generates very little char, the

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

10

difference in char yields was attributed to differences in the lignin in the two samples. Zhang et. al. hypothesized that this difference was due to the higher hydroxyl content of OSL compared to native lignin.15 Therefore, cleavage of aromatic ether linkages during fast pyrolysis formed significant amounts of PWDB while hydrolytic cleavage of these linkages during the process of lignin isolation introduced hydroxyl groups that facilitated char formation. Of the lignins tested, MCL undergoes the least amount of aromatic ether linkage hydrolytic cleavage during isolation, which allowed such ether linkages to subsequently be available for cleavage during pyrolysis. This effect resulted in a product containing more PWDB and less char than the other extracted lignins.

Conclusions Pyrolysis of cornstover lignins from three extraction methods was performed. For the MCL which had the highest molecular weight, the pyrolytic yield for total phenolics was 40% to 47% higher concomitant with less char formation compared to OSL and GVLL. Aromatic ether bond cleavage during fast pyrolysis formed significant amounts of phenolics with side chain double bonds, which primarily accounted for the higher phenolics and lower char yield from the MCL. Therefore, if the goal was to maximize carbon recovery from the pyrolytic deconstruction of lignin, the preferred lignin isolation method would be one that preserved aromatic ether bonds such as with the MCL.

Acknowledgements We are thankful to Patrick Johnston for his assistance in the gel permeation chromatography analysis.

Supporting Information Available

ACS Paragon Plus Environment

Page 10 of 16

Page 11 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

11

The supplementary information includes experimental details in TGA and GPC analysis, lignin char yield and weight loss curve from TGA analysis, FT-IR spectra of lignin samples in the range of 1400-1000 cm-1, grouping for pyrolysis products, and detailed pyrolysis product distribution from lignin at different temperatures.

References (1) Dolgonosov, B. M.; Gubernatorova, T. N. Modeling the biodegradation of multicomponent organic matter in an aquatic environment: 2. Analysis of the structural organization of lignin. Water Resour. 2010, 37, 320–331. (2) Pandey, M. P.; Kim, C. S. Lignin depolymerization and conversion: A review of thermochemical methods. Chem. Eng. Technol. 2011, 34, 29–41. (3) Faix, O.; Meier, D.; Grobe, I. Studies on isolated lignins and lignins in woody materials by pyrolysis-gas chromatography-mass spectrometry and off-line pyrolysis-gas chromatography with flame ionization detection. J. Anal. Appl. Pyrolysis 1987, 11, 403– 416. (4) Monteil-Rivera, F.; Phuong, M.; Ye, M.; Halasz, A.; Hawari, J. Isolation and characterization of herbaceous lignins for applications in biomaterials. Ind. Crops Prod. 2013, 41, 356–364. (5) Wang, S.; Wang, K.; Liu, Q.; Gu, Y.; Luo, Z.; Cen, K.; Fransson, T. Comparison of the pyrolysis behavior of lignins from different tree species. Biotechnol. Adv. 2009, 27, 562– 567. (6) Wang, S.; Lin, H.; Ru, B.; Sun, W.; Wang, Y.; Luo, Z. Comparison of the pyrolysis behavior of pyrolytic lignin and milled wood lignin by using TG-FTIR analysis. J. Anal. Appl. Pyrolysis 2014, 108, 78–85. (7) Obst, J. R. Analytical pyrolysis of hardwood and softwood lignins and its use in lignintype determination of hardwood vessel elements. J. Wood Chem. Technol. 1983, 3, 377– 397. (8) Evans, R. J.; Milne, T. A.; Soltys, M. N. Direct mass-spectrometric studies of the pyrolysis of carbonaceous fuels. III. Primary pyrolysis of lignin. J. Anal. Appl. Pyrolysis 1986, 9, 207–236. (9) Patwardhan, P. R.; Brown, R. C.; Shanks, B. H. Understanding the fast pyrolysis of lignin. ChemSusChem 2011, 4, 1629–1636. (10) Glassner, D. A.; Hettenhaus, J. R.; Schechinger, T. M. Corn stover collection project. BioEnergy ’98 Expand. BioEnergy Partnerships 1998, 2, 1100–1110. (11) Luterbacher, J. S.; Azarpira, A.; Motagamwala, A. H.; Lu, F.; Ralph, J.; Dumesic, J. A.; Gosselink, H.; Davis, R. J.; Dumesic, J. A.; Ralph, J.; et al. Lignin monomer production integrated into the γ-valerolactone sugar platform. Energy Environ. Sci. 2015, 8, 2657– 2663. (12) Björkman, A. Studies on finely divided wood. Part 1. Extraction of lignin with neutral solvents. Sven. papperstidning 1956, 59, 477–485.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

12

(13) Sluiter, A.; Hames, B.; Ruiz, R.; Scarlata, C.; Sluiter, J.; Templeton, D.; Crocker, D. Determination of structural carbohydrates and lignin in biomass: laboratory analytical procedure (LAP) (Revised July 2011). 2008. (14) Wang, K.; Zhang, J.; H. Shanks, B.; Brown, R. C. Catalytic conversion of carbohydrate-derived oxygenates over HZSM-5 in a tandem micro-reactor system. Green Chem. 2015, 17, 557–564. (15) Zhang, J.; Choi, Y. S.; Yoo, C. G.; Kim, T. H.; Brown, R. C.; Shanks, B. H. Cellulose– hemicellulose and cellulose–lignin interactions during fast pyrolysis. ACS Sustain. Chem. Eng. 2015, 3, 293–301. (16) Hage, R. El; Brosse, N.; Chrusciel, L.; Sanchez, C.; Sannigrahi, P.; Ragauskas, A. Characterization of milled wood lignin and ethanol organosolv lignin from miscanthus. Polym. Degrad. Stab. 2009, 94, 1632–1638. (17) She, D.; Nie, X. N.; Xu, F.; Geng, Z. C.; Jia, H. T.; Jones, G. L.; Baird, M. S. Physicochemical characterization of different alcohol-soluble lignins from rice straw. Cellul. Chem. Technol. 2012, 46, 3–4. (18) Hussin, M. H.; Rahim, A. A.; Mohamad Ibrahim, M. N.; Brosse, N. Physicochemical characterization of alkaline and ethanol organosolv lignins from oil palm (Elaeis guineensis) fronds as phenol substitutes for green material applications. Ind. Crops Prod. 2013, 49, 23– 32. (19) Tejado, A.; Peña, C.; Labidi, J.; Echeverria, J. M.; Mondragon, I. Physico-chemical characterization of lignins from different sources for use in phenol–formaldehyde resin synthesis. Bioresour. Technol. 2007, 98, 1655–1663. (20) Choi, Y. S.; Singh, R.; Zhang, J.; Balasubramanian, G.; Sturgeon, M. R.; Katahira, R.; Chupka, G.; Beckham, G. T.; Shanks, B. H.; Lee, J.; et al. Pyrolysis reaction networks for lignin model compounds: unraveling thermal deconstruction of β-O-4 and α-O-4 compounds. Green Chem. 2016, 18, 1762–1773.

ACS Paragon Plus Environment

Page 12 of 16

Page 13 of 16

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

13

Table 1 Purity, ash content, moisture content, and average molecular weight for the lignins: milled cornstover lignin (MCL), GVL extracted lignin (GVLL), and organosolv lignin (OSL). Lignin type Lignin purity Ash content Moisture Mna Mwb (wt%) (wt%) (wt%) (Da) (Da) MCL 97.2 ~0 0.7 2351 6740 OSL 96.5 0.1 0.6 864 2185 GVLL 0.2 0.5 827 1910 a Number average molecular weight; b weight average molecular weight.

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

14

1365

1170

MCL ORL GVLL

Transmittance, %

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 16

1400

1350

1300

1250 1200 1150 Wavenumber, cm-1

1100

1050

1000

Fig. 1. FT-IR spectra in the 1400-1000 cm-1 range for the lignin samples: milled cornstover lignin (MCL), GVL extracted lignin (GVLL), and organosolv lignin (OSL)

ACS Paragon Plus Environment

Page 15 of 16

15

(a)

(b)

40 15 30 10 20 5

GVLL

30 15

10 15 5

0

0 MCL

PWDB PNDB Char yield

20

10

0

45

0 MCL

OSL

(c)

OSL

(d) 40

30 25

20 30

15 20 10

Phenolics yield, wt%

PWDB PNDB Char yield

25

Char yield, wt%

25

20

GVLL

PWDB PNDB Char yield

20

15 15 10 10

Char yield, wt%

Phenolics yield, wt%

50 Phenolics yield, wt%

PWDB PNDB Char yield

20

25

Char yield, wt%

60

Char yield, wt%

25

Phenolics yield, wt%

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Sustainable Chemistry & Engineering

10 5

5

0

0 MCL

GVLL

OSL

5

0

0 MCL

GVLL

OSL

Fig. 2. Yield comparison for phenolics with side chain double bonds (PWDB), phenolics without side chain double bonds (PNDB), and char among the three samples, milled cornstover lignin (MCL), GVL extracted lignin (GVLL), and organosolv lignin (OSL), at pyrolysis temperature of (a) 350 °C, (b) 425 °C, (c) 500 °C, (d) 575 °C

ACS Paragon Plus Environment

ACS Sustainable Chemistry & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

For Table of Contents Use Only

Comparison of fast pyrolysis behavior of cornstover lignins isolated by different methods Jing Zhang, Kwang Ho Kim, Yong S. Choi, Ali H. Motagamwala, James A. Dumesic, Robert C. Brown, Brent H. Shanks

Synopsis Isolating lignin from cornstover by Björkman milling helps to preserve native aromatic ether bonds, leading to more phenolics and less char formation during the subsequent fast pyrolysis.

ACS Paragon Plus Environment

Page 16 of 16