5-(Chloromethyl)furfural (CMF): A Platform for Transforming Cellulose

In such cases, a company in the sustainable chemistry space may choose to first ..... Mascal, M.; Nikitin, E. B. Co-processing of Carbohydrates and Li...
0 downloads 0 Views 2MB Size
Perspective Cite This: ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

pubs.acs.org/journal/ascecg

5‑(Chloromethyl)furfural (CMF): A Platform for Transforming Cellulose into Commercial Products Mark Mascal*

ACS Sustainable Chem. Eng. Downloaded from pubs.acs.org by UNIV OF TEXAS AT DALLAS on 03/05/19. For personal use only.

Department of Chemistry, University of California Davis, 1 Shields Avenue, Davis, California 95616, United States ABSTRACT: 5-(Chloromethyl)furfural (CMF) is a carbohydrate-derived platform molecule that is gaining traction as a more practical alternative to 5-(hydroxymethyl)furfural (HMF). This perspective introduces the chemocatalytic approach to biorefining as the driving force behind the development of multifunctional chemical platforms. The main advantage of CMF over HMF is that it can be produced in high yield under mild conditions directly from raw biomass. Its stability and hydrophobicity markedly facilitate isolation. CMF is also a precursor to levulinic acid (LA), another versatile biobased intermediate. The logistics of CMF production are discussed, including reactor materials, HCl handling and management, byproducts, and the fate of collateral biomass components (hemicellulose, lipids, proteins, lignin). Examples of commercial markets that can be unlocked by synthetic manipulation of CMF are broken out into two derivative manifolds, furanic and levulinic, which are distributed over three product family trees: renewable monomers, fuels, and specialty chemicals. Selected examples of CMFand LA-based routes to these products are presented. Finally, a model for the integration of the CMF process into biorefinery practice is put forward. KEYWORDS: 5-(Chloromethyl)furfural, 5-(Hydroxymethyl)furfural, Biomass, Biorefinery, Sustainable chemistry



INTRODUCTION The recent explosion of interest in sustainable chemistry has brought about a flurry of activity in the search for practical routes to synthetically versatile platform molecules.1 This movement has been largely identified with the chemicalcatalytic approach to biorefining (Figure 1), which brings to

Figure 2. Fractions of biomass that provide the raw material for chemical-catalytic platform molecule synthesis. Green fields denote straightforward derivations; red denotes complexity.

derived platform molecules, i.e. 5-(chloromethyl)furfural, or CMF. Carbohydrates are the major component of two eukaryotic kingdoms; plants and fungi. They also figure centrally in invertebrate animals in the form of chitin. Indeed, it has been stated that cellulose is the most abundant organic molecule on the planet, not even counting hemicellulose, simple sugars, oligosaccharides, and starches. Complex carbohydrates can be broken down to their monosaccharide components, which in most cases means the C6 sugar glucose, although other hexoses and their derivatives may be variously represented, along with the C5 sugars xylose, arabinose, and some deoxy pentoses. Disregarding stereochemistry, two molecules thus constitute the great majority of terrestrial carbohydrates: 2,3,4,5,6pentahydroxyhexanal and 2,3,4,5-tetrahydroxypentanal. This

Figure 1. Biomass processing methods and their typical outputs.

bear the full arsenal of synthetic organic chemistry on the conversion process. Fermentative methods on the other hand generally focus on market-ready target compounds, while pyrolytic methods lead to complex mixtures. Even within the chemical-catalytic grouping, it is carbohydrate- and lipid-based processes that are associated with multifunctional platform development, while the heterogeneity of lignin and protein tends to complicate such endeavors (Figure 2). For lipids, the platform revolves around multiple pathways for the synthetic manipulation of glycerol. For information about that, the reader may consult reviews.2−4 This leaves carbohydrates, and it is here where the majority of effort has been concentrated. This perspective will give an overview of the state of the art with respect to an innovation in the field of carbohydrate© XXXX American Chemical Society

Received: December 13, 2018 Revised: February 6, 2019

A

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Perspective

ACS Sustainable Chemistry & Engineering

that may be any nonmiscible solvent that is not compromised by aqueous acid. Bench scale work generally prefers dichloromethane or 1,2-dichloroethane, although others, including hydrocarbons, have been employed.14 The optimal reaction temperature is feedstock dependent but varies between 80 and 120 °C. The reaction time is also feedstock dependent and may be as short as a few minutes in a flow reactor or up to 3 h in batch mode. The organic phase generally contains CMF as the only identifiable product, which can be isolated by removal of the solvent and either vacuum distillation or filtration through a pad of silica gel. When pure, it occurs as a colorless, low melting solid (mp 37−38 °C) but is frequently observed as a supercooled liquid with a slight yellow tint. The major component that makes up the carbon mass balance of the reaction is levulinic acid (LA) 3, which can be isolated from the aqueous phase by continuous extraction. An equimolar quantity of formic acid, useful as a hydrogen equivalent, is coproduced. LA is a platform molecule in its own right, and its chemistry, as a derivative of CMF, also figures centrally in this review. A question regarding this process that typically arises is this: why does this work as well as it does, when the production of HMF from cellulosic biomass under similar conditions leads to poor yields? This can be graphically illustrated as shown in Scheme 2. Cellulose hydrolysis, isomerization of glucose to

dramatically simplifies synthetic manipulation compared to other biomass fractions, such as lignin and proteins. The icon of the chemical-catalytic approach to carbohydrate processing is 5-(hydroxymethyl)furfural, or HMF, 1. Its combination of synthetic handles, i.e. aldehyde group, hydroxymethyl group, and furan ring, makes it an exceptionally versatile platform for accessing a variety of target molecules, which have variously included biofuels, renewable monomers, and other chemicals.5 However, from an industrial perspective, HMF is the revolution in sustainable chemistry that is still waiting to happen. A torrent of interest has surrounded 1: Over 2000 publications to date have appeared with HMF in the title, to which currently over 200 papers are added each year, and many more articles and patents (>12 000) include HMF somewhere in the document. A recent paper describing catalysts useful for HMF synthesis currently has nearly 1200 citations.6 Despite this, HMF is at present only practically derived from fructose, the cost of which limits the commercial potential of 1 as a platform. There is also growing opposition to the production chemicals and fuels from grain (the major source of fructose),7 and the commonly proposed exit from this predicament is to obtain glucose from cellulosic sources and then employ isomerases to access fructose. Whether this will provide an economically viable alternative path to HMF is unknown.

Scheme 2. Process for the Conversion of Cellulose to CMFa

The main obstacle to the efficient production of HMF from raw biomass is that the dehydration of sugars is an acidcatalyzed process, and HMF decomposes under these conditions to humin, a polyfuranic resin.8,9 The dehydration of fructose is fast and takes place under generally mild conditions, making high yields of 1 possible. For other sugars, however, both isomerization and dehydration are involved, which complicates the kinetics. While reports are abundant describing alternative preparations of HMF from cellulose or raw biomass under a range of conditions,5 to date, none have been adopted industrially. An answer to the above issues has been identified in the form of 5-(chloromethyl)furfural (CMF), 2. Although 2 had been observed as the product of the treatment of carbohydrates with HCl as early as 1901,10 procedures for its synthesis in good yield (from fructose) did not appear until the 1980s,11 and it was not until a 2008 report that the production of CMF in high yield from glucose and cellulose was published.12 This latter work was followed up by a process intensification study which culminated the results shown in Scheme 1.13 Thus, a carbohydrate feedstock, whether glucose, sucrose, cellulose, or raw biomass, is mixed with commercial grade hydrochloric acid typically sold at concentrations between 35 and 37% w/w. A second phase is introduced

a

Blue denotes an aqueous medium; green denotes an organic solvent.

fructose, and the dehydration of fructose to HMF all occur in the acidic aqueous phase. Since in this case the acid is HCl, the hydroxyl group in the HMF is quickly substituted for chlorine.

Scheme 1. Synthesis of CMF along with Typical Yield Ranges

B

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Perspective

ACS Sustainable Chemistry & Engineering

(e.g., Hastelloy) or stainless steel coated with tantalum, which has been known to provide decades of service. Another option is the use of azeotropic hydrochloric acid (20 w/w% or 6 M, bp 108 °C), which eliminates the requirement for a closed vessel, unless temperatures higher than the boiling point are needed. In the case of an open (batch) reactor, a higher density solvent with boiling point lower than that of the azeotrope is advantageous, since its reflux can provide a means of continuous extraction of the aqueous layer without excessive mixing of the phases, which is known to decrease CMF yield by back extraction of HMF/CMF into the aqueous acid. What about waste streams? Both the aqueous phase and organic solvent are recycled: The hydrochloric acid is brought back up to the desired titer by the ingassing of hydrogen chloride and reused, and the distilled solvent is simply looped directly back into the reactor. The carbohydrate to CMF reaction is a net producer of three molar equivalents of water, and subsequent chemistry on the chloromethyl group also liberates HCl. The recycling of hydrogen chloride is well established in industry, and multiple technologies for its recovery from solution are available, including membrane distillation,24 pervaporation,25 acid−base couple extraction,26,27 solvent extraction,28−30 diffusion dialysis,31,32 and electrodialysis.33 Technologies have also been developed for the oxidative upcycling of dilute hydrochloric acid to chlorine.34,35 Other than water and HCl, depending on the biomass source, there may also be minor extractives, which might or might not survive the reaction conditions. In cases where they may be of value (e.g., waxes from straw), they can be extracted prior to charging the reactor. Evaporation of the aqueous phase often reveals a mixture of products that appear to be undigested carbohydrates of some description, but these tend to be present in minor quantities and are unlikely to present a toxic hazard. Finally, inorganics will accumulate in the aqueous phase. Although the same hydrochloric acid has been cycled multiple times in batch mode without loss of CMF yield, at some point the aqueous phase will have to be reconditioned. The ash components (Si, Ca, K, Mg, Fe, Al, S, and P oxides)36 are used in soil amendments and fertilizers and have potential applications in the production of concrete, bricks, ceramics, glazes, and silica-based adsorbents.37

Whether that occurs in the aqueous phase or on HMF that has been extracted into the hydrogen chloride-saturated organic phase is still a matter of debate. Evidence that points to the latter is that the yield of CMF is markedly dependent on the ability of the solvent to partition HMF into the organic phase.15 The key point is that, unlike HMF, CMF is hydrophobic and is sequestered into the organic solvent where it is safe from the decomposition pathways that have long plagued HMF synthesis. When raw biomass is the feed, components other than cellulose are also taken through the process. If lipids are present, the triglyceride esters quickly hydrolyze and the fatty acids partition into the organic phase.16 This makes the CMF process a potentially useful approach for extracting oil from algae or oilseed crops for the purposes of utilizing the fatty acids (e.g., in biodiesel), since the cellular matrix is broken down and thus extraction efficiency is essentially quantitative. The carbohydrates are at the same time converted into CMF. Where proteins are a significant fraction, the CMF process converts them into a hydrolysate of amino acid hydrochlorides that remain in the aqueous phase. Amino acids are considered useful platforms in their own right.1 Hemicellulose is converted into furfural, although satisfactory yields are only observed if the CMF process is conducted at lower temperatures. This is probably due to the hydrophilicity of furfural, which limits its extraction into the organic phase and leads to rehydration and resinification on prolonged exposure to the aqueous acid environment. Since pentose and hexose sugars have different process parameters in terms of optimal temperature and acid concentration, the last section of this paper will propose the preliminary fractionation of biomass to remove hemicellulose before submitting to the CMF process. Finally, lignin exits the CMF reactor as a fine black solid, which has been studied in some detail. One interesting characteristic of this lignin is that it is low in ash.17 Since the CMF process is biphasic, most of the inorganics are washed into the aqueous layer. Ash can lead to abrasion, slagging, and fouling of gasifiers and boilers, particularly at higher temperatures;18−20 thus, low-ash carbon sources are superior solid fuels compared to untreated biomass. CMF-process lignin most closely resembles Willstaetter lignin in its highly condensed character and low solubility.21 An intriguing discovery was that it possessed a significant degree of porosity (63 m2 g−1), straight out of the reactor.22 Most remarkably, a large fraction of this involved mesopores (pore width 2−50 nm). Pyrolysis of CMF lignin at temperatures above 400 °C dramatically increased the porosity, mainly through the generation of micropores (pore width 60% overall yield. Ranitidine 37 was the world’s first billion dollar drug. It is a histamine H2-receptor antagonist used in the management of gastroesophageal reflux disease and the treatment of gastric and duodenal ulcers. Although it has been largely superseded by more effective drugs, particularly proton pump inhibitors, Ranitidine is still sold over the counter under the trade name Zantac. CMF figures centrally in one of the most efficient syntheses of 37 reported to date, as described in Scheme 17.95 The chemistry is straightforward, involving first the substitution of the chloride in 2 for commercial N-acetylcysteamine. Simple reductive amination of 34 with dimethylamine gives 35, which is deprotected by hydrolysis to 36. Substitution at the primary amine with 1-methylthio-1methylamino-2-nitroethylene, prepared from CS2, MeNO2, H

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Perspective

ACS Sustainable Chemistry & Engineering Scheme 16a

a Reagents and conditions: a. BuOH, cat. HCl, 99%. b. EtMgCl, Ni(acac)2, diallyl ether, −30 °C. c. aq. HCl, 83% over 2 steps. d. (9carboxynonyl)triphenylphosphonium iodide, LiHMDS. e. H2, Pd/C, THF, 92% over 2 steps. f. (CH2O)n, HBr, AcOH. g. H2, Pd/C, 80% over 2 steps.

Scheme 17a

a

Reagents and conditions: a. HSCH2CH2NHAc, NaH, THF, 91%. b. Me2NH, NaBH4, MeOH, 90%. c. 2 M KOH, reflux, 94%. d. 1-methylthio-1methylamino-2-nitroethylene, H2O, 55 °C, 88%.

Scheme 18a

a Reagents and conditions: a. TMS-acetylene, CuI, K2CO3, CH3CN, 55 °C, 88%. b. 1 M aq HCl, 97%. c. NaBH4, MeOH, 0 °C, 98%. d. chrysanthemic acid chloride, benzene, reflux, 95%. e. Bu4NF, THF, −20 °C, 84%.

Scheme 19a

Reagents and conditions: a. NaN3, solvent free, 110 °C. b. O2, rose bengal, hν, MeOH. c. H2, Pd/C, methanolic HCl.

a

I

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Perspective

ACS Sustainable Chemistry & Engineering Scheme 20. Integration of the CMF Process into the Biorefinery

terpenes) may be removed by treatment with an organic solvent. Since the CMF process has been optimized to operate most effectively on hexoses, a preliminary fractionation of hemicellulose, which is predominantly C5 sugars, would be advantageous. Hemicellulose is conveniently separated from lignocellulosic biomass by treatment with hot water or steam.100 It could be valorized in one of two ways, depending various economic considerations, including the prevailing markets. The first option is to dehydrate the pentoses into furfural,101 itself a commercially traded product. Since some of the above biofuels and monomers descend from LA as a platform, the known conversion of furfural to LA via the rehydration of furfuryl alcohol could form one manifold of this biorefinery.102 Alternatively, a hydrolysate of the hemicellulose fraction103 could be fermented to n-butanol. Various strains of microorganisms such as Clostridium sp. and Thermoanaerobacterium sp. have been shown to utilize pentose sugars as well as hexoses in the production of biobutanol.104,105 It may well be said that, once you have butanol, you essentially have access to the whole range of chemicals derivable from the light naphtha raffinate of petroleum.106 The cellulose-lignin matrix is then fed into the CMF process, where the cellulose is converted into CMF, which branches out into the monomer, biofuel, and chemical family trees. The insoluble lignin byproduct is separated out to potentially serve as a porous material for various applications or a low-ash feed for gasifiers, as was described above. Ultimately, the most profitable use of lignin would be to serve as a source of fuels or chemicals. Indeed, this is one of the most intensively researched subjects in the field of sustainable chemistry.107,108 The recalcitrance and heterogeneity of lignin generally thwart any expectation that a practical pathway to a single, highly useful derivative will emerge therefrom. Instead, it seems most likely that the technologies that best valorize lignin will be those that embrace its complexity, rather than fight it. For example, recent work describes the catalytic hydrogenation of the aromatic rings in lignin to give a polyether matrix that is hydrodeoxygenated to yield a mixture of aliphatic hydrocarbons in the C6−C20 range, with the highest concentration predictably at C9.109 Since motor fuels are mixtures in any case,

MeI, and MeNH2, is also facile, concluding a four step synthesis in overall 68% yield. Prothrin 41 is a known pyrethroid insecticide.96 As in the above cases, the 2,5-disubstituted furan core is clearly visible in its structure, inviting yet another CMF-based synthesis. This was accomplished as shown in Scheme 18. Thus, dibutyl acetal 30, borrowed from Scheme 16, undergoes Cu-mediated coupling with trimethylsilyl (TMS) acetylene to give 38. Deprotection of the aldehyde and reduction gives hydroxymethylfuran 40, which is esterified with chrysanthemic acid chloride. Removal of the TMS protecting group finally gives prothrin 41 in 6 steps from CMF and 69% overall yield. Tests of the insecticidal activity of 41 were carried out against the industry standard promethrin, where the former showed an LD50 for the A. aegypti mosquito within an order of magnitude of the latter. Finally, the photodynamic therapy drug δ-aminolevulinic acid (DALA) 46 was also prepared from CMF as described in Scheme 19.97 Here, the chemistry of the furan ring features centrally, where singlet O2 adds to give trioxabicycloheptene 43. This unstable intermediate decarbonylates by reaction with methanol to give butenolide 44, which opens up to azidooxopentenoic acid 45. This product was not isolated but the azido group and CC bond were simultaneously hydrogenated to give the product 46 as the stable HCl salt in a remarkable 3 steps from CMF and 68% overall yield. Commercial DALA is produced biosynthetically,98 and various sources place the cost of 46 at about USD 100 per gram; whereas the route in Scheme 19, assuming catalyst recycling, could provide 46 at USD 1 or less per gram. It should also be noted that DALA is also considered a promising agrochemical, both as a herbicide and insecticide.99



POTENTIAL ROLE OF THE CMF PROCESS IN THE INTEGRATED BIOREFINERY The core concept of the integrated biorefinery is essentially to exploit all of the fractions of raw biomass to their highest commercial potential. Scheme 20 shows how the CMF process might be incorporated into a biorefinery project. First, prior to fractionation, any potentially valuable extractives (e.g., waxes, J

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Perspective

ACS Sustainable Chemistry & Engineering

source to valuable commodity chemicals. Chem. Soc. Rev. 2008, 37, 527−549. (3) Pagliaro, M.; Ciriminna, R.; Kimura, H.; Rossi, M.; Della Pina, C. From glycerol to value-added products. Angew. Chem., Int. Ed. 2007, 46, 4434−4440. (4) Bagheri, S.; Julkapli, N. M.; Yehye, W. A. Catalytic conversion of biodiesel derived raw glycerol to value added products. Renewable Sustainable Energy Rev. 2015, 41, 113−127. (5) van Putten, R.-J.; van der Waal, J. C.; de Jong, E.; Rasrendra, C. B.; Heeres, H. J.; de Vries, J. G. Hydroxymethylfurfural, A Versatile Platform Chemical Made from Renewable Resources. Chem. Rev. 2013, 113, 1499−1597. (6) Zhao, H.; Holladay, J. E.; Brown, H.; Zhang, Z. C. Metal Chlorides in Ionic Liquid Solvents Convert Sugars to 5-Hydroxymethylfurfural. Science 2007, 316, 1597−1600. (7) Shurson, G. C. The Role of Biofuels Coproducts in Feeding the World Sustainably. Annu. Rev. Anim. Biosci. 2017, 5, 229−254. (8) Wang, S.; Lin, H.; Zhao, Y.; Chen, J.; Zhou, J. Structural characterization and pyrolysis behavior of humin by-products from the acid-catalyzed conversion of C6 and C5 carbohydrates. J. Anal. Appl. Pyrolysis 2016, 118, 259−266. (9) Herzfeld, J.; Rand, D.; Matsuki, Y.; Daviso, E.; Mak-Jurkauskas, M.; Mamajanov, I. Molecular Structure of Humin and Melanoidin via Solid State NMR. J. Phys. Chem. B 2011, 115, 5741−5745. (10) Fenton, H. J. H.; Gostling, M. Derivatives of Methylfurfurals. J. Chem. Soc., Trans. 1901, 79, 807−816. (11) Szmant, H. H.; Chundury, D. D. The preparation of 5chloromethylfurfuraldehyde from high fructose corn syrup and other carbohydrates. J. Chem. Technol. Biotechnol. 1981, 31, 205−212. (12) Mascal, M.; Nikitin, E. B. Direct, High-Yield Conversion of Cellulose into Biofuel. Angew. Chem., Int. Ed. 2008, 47, 7924−7926. (13) Mascal, M.; Nikitin, E. B. Dramatic Advancements in the Saccharide to 5-(Chloromethyl)furfural Conversion Reaction. ChemSusChem 2009, 2, 859−861. (14) Breeden, S. W.; Clark, J. H.; Farmer, T. J.; Macquarrie, D. J.; Meimoun, J. S.; Nonne, Y.; Reid, J. E. S. J. Microwave heating for rapid conversion of sugars and polysaccharides to 5-chloromethyl furfural. Green Chem. 2013, 15, 72−75. (15) Lane, D. R.; Mascal, M.; Stroeve, P. Experimental studies towards optimization of the production of 5-(chloromethyl)furfural (CMF) from glucose in a two-phase reactor. Renewable Energy 2016, 85, 994−1001. (16) Mascal, M.; Nikitin, E. B. Co-processing of Carbohydrates and Lipids in Oil Crops To Produce a Hybrid Biodiesel. Energy Fuels 2010, 24, 2170−2171. (17) Analysrapport; Eurofins Environment Sweden AB (Lidköping), 2013. (18) Benson, S. A.; Zygarlicke, C. J.; Sondreal, E. A. Efficient and clean power production: minimizing impacts of inorganic components in coal and other fuels. Preprints of SymposiaAmerican Chemical Society, Division of Fuel Chemistry 2000, 45, 88−92. (19) Cai, Y.; Tay, K.; Zheng, Z.; Yang, W.; Wang, H.; Zeng, G.; Li, Z.; Keng Boon, S.; Subbaiah, P. Modeling of ash formation and deposition processes in coal and biomass fired boilers: A comprehensive review. Appl. Energy 2018, 230, 1447−1544. (20) Hupa, M. Ash-Related Issues in Fluidized-Bed Combustion of Biomasses: Recent Research Highlights. Energy Fuels 2012, 26, 4−14. (21) Rassow, B.; Zickmann, P. Willstätter Lignin. J. Prakt. Chem. 1929, 123, 189−234. (22) Budarin, V. L.; Clark, J. H.; Henschen, J.; Farmer, T. J.; Macquarrie, D. J.; Mascal, M.; Nagaraja, G. K.; Petchey, T. H. M. Processed Lignin as a Byproduct of the Generation of 5(Chloromethyl)furfural from Biomass: A Promising New Mesoporous Material. ChemSusChem 2015, 8, 4172−4179. (23) Moeller, K.; Bein, T. Mesoporosity - a new dimension for zeolites. Chem. Soc. Rev. 2013, 42, 3689−3707. (24) Tomaszewska, M.; Gryta, M.; Morawski, A. W. Recovery of hydrochloric acid from metal pickling solutions by membrane distillation. Sep. Purif. Technol. 2001, 22−23, 591−600.

the complexity of lignin can be seen here as an advantage. It may be that technologies of this general nature will provide future value streams for lignin and serve to best complete this arm of the scheme.



CONCLUSION The chemical-catalytic approach to biomass processing has significant advantages over competing methods: Fermentation relies on sugar feeds and is no match for the production velocity of catalytic technologies, while energy-intensive pyrolytic methods lead to complex mixtures, which necessitates expensive upgrading (bio-oil) or the handling of large volumes of vapor (biogas). There are currently two multifunctional, chemocatalytic platform molecules that are industrially produced in high yield directly from feedstocks that are noncompetitive with food: CMF and LA, the latter of which can also be efficiently made from CMF. CMF is functionally equivalent to the iconic fructose derivative HMF, and its generation from carbohydrates parallels that of HMF. However, final substitution of the OH for Cl renders the CMF molecule lipophilic and its production in a biphasic reactor allows it to be extracted out of the aqueous acid layer, protecting it from the decomposition pathways that surround HMF. The derivative chemistries of CMF and LA are broad, encompassing first a series of monomers that include bifunctional aromatics and aliphatics, both drop-ins and new materials. Likewise, a number of biobased fuels have been described. Furan and levulinate-based oxygenates and fuel additives can be accessed directly from CMF, while highoctane gasoline blendstocks are produced by combining levulinate coupling reactions with hydrodeoxygenation and hydrodecarboxylation chemistry. Finally, CMF figures centrally in efficient routes to specialty products, for example in the healthcare and agrochemical field. CMF may be integrated into a model of the integrated biorefinery where cellulosic biomass is prefractionated into hemicellulose and a lignin-cellulose conglomerate. The former may be most advantageously converted into LA or hydrolyzed to sugars and fermented. The latter enters the CMF process, resulting in CMF and solid lignin. The CMF can be sent down any of the synthetic pathways described above (monomer, biofuel, specialty chemical), while the lignin has applications as a structured, porous support for separations/catalysis, a lowash gasifier feed, or ultimately a source of chemicals, the latter however being qualified as most useful in products that embrace complexity, such as liquid fuels.



AUTHOR INFORMATION

Corresponding Author

*Email address: [email protected]. ORCID

Mark Mascal: 0000-0001-7841-253X Notes

The author declares no competing financial interest.



REFERENCES

(1) Farmer, T. J.; Mascal, M. Platform Molecules. In Introduction to Chemicals from Biomass, Second ed.; Clark, J., Deswarte, F., Eds.; John Wiley & Sons: Chichester, 2015; pp 89−155. (2) Zhou, C.-H.; Beltramini, J. N.; Fan, Y.-X.; Lu, G. Q. Chemoselective catalytic conversion of glycerol as a biorenewable K

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Perspective

ACS Sustainable Chemistry & Engineering (25) Schuchardt, U.; Joekes, I.; Duarte, H. C. Hydrolysis of Sugar Cane Bagasse with Hydrochloric Acid: Separation of the Acid by Pervaporation. Evaluation of the Bergius Process. J. Chem. Technol. Biotechnol. 1988, 41, 51−60. (26) Baniel, A.; Eyal, A. A process for the recovery of HCl from a dilute solution thereof. WO 2008111045, 2008. (27) Sarangi, K.; Padhan, E.; Sarma, P. V. R. B.; Park, K. H.; Das, R. P. Removal/recovery of hydrochloric acid using Alamine 336, Aliquat 336, TBP and Cyanex 923. Hydrometallurgy 2006, 84, 125−129. (28) Gaddy, J. L.; Clausen, E. C. Recovery of concentrated hydrochloric acid from a product comprising sugars and hydrochloric acid from acid hydrolysis of biomass. US Patent 4645658, 1987. (29) Crittenden, E. D.; Hixson, A. N. Extraction of Hydrogen Chloride from Aqueous Solutions. Ind. Eng. Chem. 1954, 46, 265− 274. (30) Forster, A. V.; Martz, L. E.; Leng, D. E. Recovering concentrated hydrochloric acid from the crude product obtained from acid hydrolysis of cellulose. Eur. Pat. Appl. EP18621, 1980. (31) Zhang, X.; Li, C.; Wang, X.; Wang, Y.; Xu, T. Recovery of hydrochloric acid from simulated chemosynthesis aluminum foils wastewater: An integration of diffusion dialysis and conventional electrodialysis. J. Membr. Sci. 2012, 409−410, 257−263. (32) Xu, J.; Lu, S.; Fu, D. Recovery of hydrochloric acid from the waste acid solution by diffusion dialysis. J. Hazard. Mater. 2009, 165, 832−837. (33) Rohman, F. S.; Aziz, N. Optimization of batch electrodialysis for hydrochloric acid recovery using orthogonal collocation method. Desalination 2011, 275, 37−49. (34) Han, M.; Chang, P.; Hu, G.; Chen, Z.; Wang, D.; Wei, F. Conversion of hydrogen chloride to chlorine by catalytic oxidation in a two-zone circulating fluidized bed reactor. Chem. Eng. Process. 2011, 50, 593−598. (35) Mortensen, M.; Minet, R. G.; Tsotsis, T. T.; Benson, S. W. The development of a dual fluidized-bed reactor system for the conversion of hydrogen chloride to chlorine. Chem. Eng. Sci. 1999, 54, 2131− 2139. (36) Vassilev, S. V.; Baxter, D.; Andersen, L. K.; Vassileva, C. G. An overview of the composition and application of biomass ash. Part 1. Phase−mineral and chemical composition and classification. Fuel 2013, 105, 40−76. (37) Vassilev, S. V.; Baxter, D.; Andersen, L. K.; Vassileva, C. G. An overview of the composition and application of biomass ash. Part 2. Potential utilisation, technological and ecological advantages and challenges. Fuel 2013, 105, 19−39. (38) Mascal, M.; Nikitin, E. B. High-yield conversion of plant biomass into the key value-added feedstocks 5-(hydroxymethyl)furfural, levulinic acid, and levulinic esters via 5-(chloromethyl)furfural. Green Chem. 2010, 12, 370−373. (39) Yan, L.; Yao, Q.; Fu, Y. Conversion of levulinic acid and alkyl levulinates into biofuels and high-value chemicals. Green Chem. 2017, 19, 5527−5547. (40) Antonetti, C.; Licursi, D.; Fulignati, S.; Valentini, G.; Galletti, A. M. R. New Frontiers in the Catalytic Synthesis of Levulinic Acid: From Sugars to Raw and Waste Biomass as Starting Feedstock. Catalysts 2016, 6, 196. (41) Pileidis, F. D.; Titirici, M.-M. Levulinic Acid Biorefineries: New Challenges for Efficient Utilization of Biomass. ChemSusChem 2016, 9, 562−582. (42) Yan, K.; Jarvis, C.; Gu, J.; Yan, Y. Production and catalytic transformation of levulinic acid: A platform for speciality chemicals and fuels. Renewable Sustainable Energy Rev. 2015, 51, 986−997. (43) Rackemann, D. W.; Doherty, W. O. S. The conversion of lignocellulosics to levulinic acid. Biofuels, Bioprod. Biorefin. 2011, 5, 198−214. (44) Hayes, D. J.; Fitzpatrick, S.; Hayes, M. H. B.; Ross, J. R. H. The Biofine process−production of levulinic acid, furfural, and formic acid from lignocellulosic feedstocks. Biorefineries−Industrial Processes and Products; Kamm, B., Gruber, P. R., Kamm, M., Eds.; 2006; Vol. 1, pp 139−164.

(45) Chundury, D.; Szmant, H. H. Preparation of polymeric building blocks from 5-hydroxymethyl- and 5-chloromethylfurfuraldehyde. Ind. Eng. Chem. Prod. Res. Dev. 1981, 20, 158−163. (46) Moore, J. A.; Kelly, J. E. Polyesters derived from furan and tetrahydrofuran nuclei. Macromolecules 1978, 11, 568−573. (47) Satsangi, S. Polyethylene Terephthalate (PET) Market by Application (Beverages, Sheet & Films, Consumer Goods, Food Packaging, and Others) and End-use Industry (Packaging, Electrical & Electronics, Automotive, Construction, and Others): Global Opportunity Analysis and Industry Forecast, 2017−2023; Allied Market Research report, May 2017. (48) Papageorgiou, G. Z.; Tsanaktsis, V.; Bikiaris, D. N. Synthesis of poly(ethylene furandicarboxylate) polyester using monomers derived from renewable resources: thermal behavior comparison with PET and PEN. Phys. Chem. Chem. Phys. 2014, 16, 7946−7958. (49) Jiang, M.; Liu, Q.; Zhang, Q.; Ye, C.; Zhou, G. A series of furan-aromatic polyesters synthesized via direct esterification method based on renewable resources. J. Polym. Sci., Part A: Polym. Chem. 2012, 50, 1026−1036. (50) Burgess, S. K.; Karvan, O.; Johnson, J. R.; Kriegel, R. M.; Koros, W. J. Oxygen sorption and transport in amorphous poly(ethylene furanoate). Polymer 2014, 55, 4748−4756. (51) Zhang, J.; Li, J.; Tang, Y.; Lin, L.; Long, M. Advances in catalytic production of bio-based polyester monomer 2,5-furandicarboxylic acid derived from lignocellulosic biomass. Carbohydr. Polym. 2015, 130, 420−428. (52) Laugel, C.; Estrine, B.; Le Bras, J.; Hoffmann, N.; Marinkovic, S.; Muzart, J. NaBr/DMSO-Induced Synthesis of 2,5-Diformylfuran from Fructose or 5-(Hydroxymethyl)furfural. ChemCatChem 2014, 6, 1195−1198. (53) Dutta, S.; Wu, L.; Mascal, M. Production of 5-(chloromethyl)furan-2-carbonyl chloride and furan-2,5-dicarbonyl chloride from biomass-derived 5-(chloromethyl)furfural (CMF). Green Chem. 2015, 17, 3737−3739. (54) Mintz, M. J.; Walling, C. t-Butyl hypochlorite. Org. Synth. 1969, 49, 9−10. (55) Haworth, W. N.; Jones, W. G. M. The Conversion of Sucrose into Furan Compounds. Part I. 5-Hydroxymethylfurfuraldehyde and Some Derivatives. J. Chem. Soc. 1944, 667−670. (56) Hulea, V. Toward Platform Chemicals from Bio-Based Ethylene: Heterogeneous Catalysts and Processes. ACS Catal. 2018, 8, 3263−3279. (57) Maneffa, A.; Priecel, P.; Lopez-Sanchez, J. A. Biomass-Derived Renewable Aromatics: Selective Routes and Outlook for p-Xylene Commercialization. ChemSusChem 2016, 9, 2736−2748. (58) Karakhanov, E. A.; Maksimov, A. L.; Zolotukhina, A. V.; Vinokurov, V. A. Oxidation of p-Xylene. Russ. J. Appl. Chem. 2018, 91, 707−727. (59) Smith, P. B. Bio-based sources for terephthalic acid. Green Polymer Chemistry: Biobased Materials and Biocatalysis; ACS Symposium Series 2015; pp 453−469. (60) Smith, P. B.; Henton, D. R.; Dumitrascu, A.; Hucul, D. A.; Masuno, M.; Smith, R.; Bissell, J. Bio-based sources for p-xylene. Abstracts of Papers, 253rd ACS National Meeting & Exposition, 2017; CELL-406. (61) Spanjers, C. S.; Schneiderman, D. K.; Wang, J. Z.; Wang, J.; Hillmyer, M. A.; Zhang, K.; Dauenhauer, P. J. Branched Diol Monomers from the Sequential Hydrogenation of Renewable Carboxylic Acids. ChemCatChem 2016, 8, 3031−3035. (62) Arnaud, S. P.; Wu, L.; Wong Chang, M.-A.; Comerford, J. W.; Farmer, T. J.; Schmid, M.; Chang, F.; Li, Z.; Mascal, M. New biobased monomers: tuneable polyester properties using branched diols from biomass. Faraday Discuss. 2017, 202, 61−77. (63) van der Klis, F.; Knoop, R. J. I.; Bitter, J. H.; van den Broek, L. A. M. The effect of Me-substituents of 1,4-butanediol analogues on the thermal properties of biobased polyesters. J. Polym. Sci., Part A: Polym. Chem. 2018, 56, 1903−1906. L

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Perspective

ACS Sustainable Chemistry & Engineering (64) Iovel, I.; Goldberg, Y.; Shymanska, M. Hydroxymethylation of furan and its derivatives in the presence of cation-exchange resins. J. Mol. Catal. 1989, 57, 91−103. (65) Latifi, E.; Marchese, A. D.; Hulls, M. C. W.; Soldatov, D. V.; Schlaf, M. [Ru(triphos)(CH3CN)3](OTf)2 as a homogeneous catalyst for the hydrogenation of biomass derived 2,5-hexanedione and 2,5dimethyl-furan in aqueous acidic medium. Green Chem. 2017, 19, 4666−4679. (66) Schaefer, H. J. Recent contributions of Kolbe electrolysis to organic synthesis. Topics Curr. Chem. 1990, 152, 91−151. (67) Alexander, S.; Eastoe, J.; Lord, A. M.; Guittard, F.; Barron, A. R. Branched Hydrocarbon Low Surface Energy Materials for Superhydrophobic Nanoparticle Derived Surfaces. ACS Appl. Mater. Interfaces 2016, 8, 660−666. (68) Alexander, S.; Smith, G. N.; James, C.; Rogers, S. E.; Guittard, F.; Sagisaka, M.; Eastoe, J. Low-Surface Energy Surfactants with Branched Hydrocarbon Architectures. Langmuir 2014, 30, 3413− 3421. (69) Helberger, J. H.; Ulubay, S.; Civelekoglu, H. Ein einfaches Verfahren zur Gewinnung von α-Angelicalacton und über die hydrierende Spaltung sauerstoffhaltiger Ringe. Liebigs Ann. Chem. 1949, 561, 215−220. (70) Lukes, R.; Nemec, J.; Jary, J. Ü ber die Dimerisierung von αAngelicalacton. Collect. Czech. Chem. Commun. 1964, 29, 1663−1668. (71) Mascal, M.; Dutta, S.; Gandarias, I. Hydrodeoxygenation of the Angelica Lactone Dimer, a Cellulose-Based Feedstock: Simple, HighYield Synthesis of Branched C7-C10 Gasoline-like Hydrocarbons. Angew. Chem., Int. Ed. 2014, 53, 1854−1857. (72) Kabasci, S.; Bretz, I. Succinic acid: Synthesis of biobased polymers from renewable resources. Renewable Polymers 2011, 355− 379. (73) Andreessen, B.; Taylor, N.; Steinbuechel, A. Poly(3hydroxypropionate): a promising alternative to fossil fuel-based materials. Appl. Environ. Microbiol. 2014, 80, 6574−6582. (74) Karp, E. M.; Eaton, T. R.; Sanchez i Nogue, V.; Vorotnikov, V.; Biddy, M. J.; Tan, E. C. D.; Brandner, D. G.; Cywar, R. M.; Liu, R.; Manker, L. P.; Michener, W. E.; Gilhespy, M.; Skoufa, Z.; Watson, M. J.; Fruchey, O. S.; Vardon, D. R.; Gill, R. T.; Bratis, A. D.; Beckham, G. T. Renewable acrylonitrile production. Science 2017, 358, 1307− 1310. (75) Akhtar, J.; Idris, A.; Abd Aziz, R. Recent advances in production of succinic acid from lignocellulosic biomass. Appl. Microbiol. Biotechnol. 2014, 98, 987−1000. (76) Nghiem, N. P.; Kleff, S.; Schwegmann, S. Succinic acid: technology development and commercialization. Fermentation 2017, 3, 26. (77) Matsakas, L.; Hruzova, K.; Rova, U.; Christakopoulos, P. Biological production of 3-hydroxypropionic acid: an update on the current status. Fermentation 2018, 4, 13. (78) Kumar, V.; Ashok, S.; Park, S. Microbial production of 3hydroxypropionic acid from renewable sources: a green approach as an alternative to conventional chemistry. Bioprocessing of Renewable Resources to Commodity Bioproducts; Bisaria, V. S., Kondo, A., Eds.; 2014; pp 381−407. (79) Podolean, I.; Kuncser, V.; Gheorghe, N.; Macovei, D.; Parvulescu, V. I.; Coman, S. M. Ru-based magnetic nanoparticles (MNP) for succinic acid synthesis from levulinic acid. Green Chem. 2013, 15, 3077−3082. (80) Choudhary, H.; Nishimura, S.; Ebitani, K. Metal-free oxidative synthesis of succinic acid from biomass-derived furan compounds using a solid acid catalyst with hydrogen peroxide. Appl. Catal., A 2013, 458, 55−62. (81) Dalli, S. S.; Tilaye, T. J.; Rakshit, S. K. Conversion of WoodBased Hemicellulose Prehydrolysate into Succinic Acid Using a Heterogeneous Acid Catalyst in a Biphasic System. Ind. Eng. Chem. Res. 2017, 56, 10582−10590. (82) Falletta, E.; Della Pina, C.; Rossi, M.; He, Q.; Kiely, C. J.; Hutchings, G. J. Enhanced performance of the catalytic conversion of

allyl alcohol to 3-hydroxypropionic acid using bimetallic gold catalysts. Faraday Discuss. 2011, 152, 367−379. (83) Dutta, S.; Wu, L.; Mascal, M. Efficient, metal-free production of succinic acid by oxidation of biomass-derived levulinic acid with hydrogen peroxide. Green Chem. 2015, 17, 2335−2338. (84) Wu, L.; Dutta, S.; Mascal, M. Efficient, Chemical-Catalytic Approach to the Production of 3-Hydroxypropanoic Acid by Oxidation of Biomass-Derived Levulinic Acid With Hydrogen Peroxide. ChemSusChem 2015, 8, 1167−1169. (85) Daniel, R.; Xu, H.; Wang, C.; Richardson, D.; Shuai, S. Combustion performance of 2,5-dimethylfuran blends using dualinjection compared to direct-injection in a SI engine. Appl. Energy 2012, 98, 59−68. (86) Jackson, M. A.; Appell, M.; Blackburn, J. A. Hydrodeoxygenation of Fructose to 2,5-Dimethyltetrahydrofuran Using a Sulfur Poisoned Pt/C Catalyst. Ind. Eng. Chem. Res. 2015, 54, 7059− 7066. (87) Windom, B. C.; Lovestead, T. M.; Mascal, M.; Nikitin, E. B.; Bruno, T. J. Advanced Distillation Curve Analysis on Ethyl Levulinate as a Diesel Fuel Oxygenate and a Hybrid Biodiesel Fuel. Energy Fuels 2011, 25, 1878−1890. (88) Mikochik, P.; Cahana, A. Conversion of 5-(chloromethyl)-2furaldehyde into 5-methyl-2-furoic acid and derivatives thereof. PCT Int. Appl. 2012, WO2012024353; see also http://xftechnologies.com. (89) Chang, F.; Dutta, S.; Mascal, M. Hydrogen-Economic Synthesis of Gasoline-like Hydrocarbons by Catalytic Hydrodecarboxylation of the Biomass-derived Angelica Lactone Dimer. ChemCatChem 2017, 9, 2622−2626. (90) Wu, L.; Mascal, M.; Farmer, T. J.; Pérocheau Arnaud, S.; Wong Chang, M.-A. Electrochemical Coupling of Biomass-Derived Acids: New C8 Platforms for Renewable Polymers and Fuels. ChemSusChem 2017, 10, 166−170. (91) Li, Z.; Otsuki, A. L.; Mascal, M. Production of cellulosic gasoline via levulinic ester self-condensation. Green Chem. 2018, 20, 3804−3808. (92) Spiteller, G. Furan fatty acids: Occurrence, synthesis, and reactions. Are furan fatty acids responsible for the cardioprotective effects of a fish diet? Lipids 2005, 40, 755−771. (93) Xu, L.; Sinclair, A. J.; Faiza, M.; Li, D.; Han, X.; Yin, H.; Wang, Y. Furan fatty acids - Beneficial or harmful to health? Prog. Lipid Res. 2017, 68, 119−137. (94) Chang, F.; Hsu, W.-H.; Mascal, M. Synthesis of antiinflammatory furan fatty acids from biomass-derived 5(chloromethyl)furfural. Sus. Chem. Pharm. 2015, 1, 14−18. (95) Mascal, M.; Dutta, S. Synthesis of ranitidine (Zantac) from cellulose-derived 5-(chloromethyl)furfural. Green Chem. 2011, 13, 3101−3102. (96) Katsuda, Y.; Chikamoto, T.; Ogami, H.; Hirobe, H.; Kunishige, T. Novel insecticidal chrysanthemic esters. Agric. Biol. Chem. 1969, 33, 1361−1362. (97) Mascal, M.; Dutta, S. Synthesis of the natural herbicide δaminolevulinic acid from cellulose-derived 5-(chloromethyl)furfural. Green Chem. 2011, 13, 40−41. (98) Sasaki, K.; Watanabe, M.; Tanaka, T. Biosynthesis, biotechnological production and application of 5-aminolevulinic acid. Appl. Microbiol. Biotechnol. 2002, 58, 23−29. (99) Sasikala, C.; Ramana, C. V.; Rao, P. R. 5-Aminolevulinic acid: A potential herbicide/insecticide from microorganisms. Biotechnol. Prog. 1994, 10, 451−459. (100) Liu, S.; Lu, H.; Lei, Y.; Wood, C. D.; Amidon, T. E.; Liang, B.; Sun, R.; Scott, G. M.; Nichol, D. I.; Ward, A. Separation of hemicellulose by hot-water extraction from woody biomass. Integr. Biorefin. 2013, 651−706. (101) Luo, Y.g; Li, Z.; Li, X.; Liu, X.; Fan, J.; Clark, J. H.; Hu, C. The production of furfural directly from hemicellulose in lignocellulosic biomass: A review. Catal. Today 2019, 319, 14−24. (102) Gonzalez Maldonado, G. M.; Assary, R. S.; Dumesic, J.; Curtiss, L. A. Experimental and theoretical studies of the acidM

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX

Perspective

ACS Sustainable Chemistry & Engineering catalyzed conversion of furfuryl alcohol to levulinic acid in aqueous solution. Energy Environ. Sci. 2012, 5, 6981−6989. (103) Maki-Arvela, P.; Salmi, T.; Holmbom, B.; Willfor, S.; Murzin, D. Y. Synthesis of Sugars by Hydrolysis of Hemicelluloses- A Review. Chem. Rev. 2011, 111, 5638−5666. (104) Guan, W.; Xu, G.; Duan, J.; Shi, S. Acetone-Butanol-Ethanol Production from Fermentation of Hot-Water-Extracted Hemicellulose Hydrolysate of Pulping Woods. Ind. Eng. Chem. Res. 2018, 57, 775− 783. (105) Jiang, Y.; Liu, J.; Dong, W.; Zhang, W.; Fang, Y.; Ma, J.; Jiang, M.; Xin, F. The Draft Genome Sequence of Thermophilic Thermoanaerobacterium thermosaccharolyticum M5 Capable of Directly Producing Butanol from Hemicellulose. Curr. Microbiol. 2018, 75, 620−623. (106) Mascal, M. Chemicals from biobutanol: technologies and markets. Biof uels, Bioprod. Biofuels, Bioprod. Biorefin. 2012, 6, 483− 493. (107) Zakzeski, J.; Bruijnincx, P. C. A.; Jongerius, A. L.; Weckhuysen, B. M. The Catalytic Valorization of Lignin for the Production of Renewable Chemicals. Chem. Rev. 2010, 110, 3552− 3599. (108) Li, C.i; Zhao, X.; Wang, A.; Huber, G. W.; Zhang, T. Catalytic Transformation of Lignin for the Production of Chemicals and Fuels. Chem. Rev. 2015, 115, 11559−11624. (109) Cao, Z.; Dierks, M.; Clough, M. T.; Daltro de Castro, I. B.; Rinaldi, R. A Convergent Approach for a Deep Converting LigninFirst Biorefinery Rendering High-Energy-Density Drop-in Fuels. Joule 2018, 2, 1118−1133.

N

DOI: 10.1021/acssuschemeng.8b06553 ACS Sustainable Chem. Eng. XXXX, XXX, XXX−XXX