throbber
Exhibit 2006
`E.I. du Pont de Nemours & Co. and
`Acher-Daniels-Midland Co. v. Furanix Technologies BV
`IPR2015-01838
`
`

`
`cr-rnrrczu. ENGINEERING nzszucr-r AND DESIGN 8 7 (2 o o 9) 1318-1327
`
`1319
`
`food regulations. Nevertheless all the potential technologies
`(whether approved for food or non-food production) need to be
`able to overcome the pH and temperature instability and lim-
`ited solubility ir1 organic solvents. It is because of the nature
`of glucose therefore that one obvious starting point is to use
`enzymatic catalysis (water based and under mild conditions).
`In this paper we will review the alternative technologies and
`routes from glucose to FDA, and discuss some of the limita-
`tions and drallenges.
`
`Biomass as a raw material for
`2.
`biorefmeries
`
`Nature is producing vast amounts of biomass driven by sun-
`light via photosynthesis:
`
`YICO2 + nH2O —> (CI-120),. + 7102
`
`However, utilization of biomass for producing chemicals and
`fuels is sfill in its infancy with only 3.5% being used for food
`or non-food purposes. Plant biomass consists mainly of car-
`bohydrates, lignin, protein and fats. Out of an estimated 170
`billion metric tons of biomass produced every year roughly
`75% are in the form of carbohydrates which makes biomass
`carbohydrates the most abundant renewable resource (Roper,
`2002). Tbgether with their amenability towards enzymatic pro-
`cesses this makes carbohydrates the center of attention when
`looking for new and greener feedstocks to replace petroleum
`for producing commodity chemicals as well as fuels. In plant
`biomass most of the carbohydrates are stored as sugar poly-
`mers such as starch, cellulose or hernicellulose.
`Starch is the second largest biomass produced on earth and
`commonly found in vegetables, such as corn, wheat, rice, pota-
`toes and beans. The total world production in 2004 was 60
`million tons of whidl more than 70% came from corn. Starch
`
`consists of chains of glucose molecules, which are linked
`together by u-1,4 and u—1,6 glycosidic bonds. The two major
`parts of starch are amylose (20—30%), essentially linear a—1,4
`glucan chains and arnylopectin (70—80%), a branched molecule
`containing 4—5% u-1,6 linkages.
`Starch is industrially hydrolyzed to glucose by the three
`enzymes: a-amylase, glucoarnylase, and pullulanase (Schafer
`et al., 2007). Bacterial a-amylases (EC 3.2.1.1) catalyze the
`hydrolysis of internal a-1,4 glycosidic bonds. This reduces
`the viscosity, which is necessary for further processing. Glu-
`coamylase (EC 3.2.1.3) is an exo-amylase that is added to the
`partly hydrolyzed stardr after liquefaction. Glucose units are
`removed in a stepwise manner from the non—reducing end
`of the molecule. The third enzyme is pullulanase (EC 3.2.1.41).
`Industrially used pullulanases are heat stable enzymes, which
`act simultaneously with glucoarnaylase during saccharifica-
`tion. Pullulanases catalyze the hydrolysis of the a-1,6 linkages
`in amylopectin, and especially in partially hydrolysed amy-
`lopectin. Typical process conditions for production of glucose
`from starch are given in Table 1.
`Cellulose is a glucose polymer consisting of linear chains
`of glucopyranose units linked together via B-1,4 glucosidic
`
`bonds. Unlike starch, cellulose is a crystalline material where
`inter- and intra-molecular hydrogen bonding gives rise to the
`very stable cellulose fiber. Hernicellulose is a polysaccharide
`consisting of short highly branched chains of different car-
`bohydrate units, including five- as well as six-carbon units
`(e.g. xyloses, galactose, glucose, mannose and arabinose).
`Hernicelluloses are much easier to hydrolyze than cellu-
`lose. The structured portion of biomass, such as straw, com
`stover, grasses and wood, is made of lignocellulose com-
`posed mainly of cellulose (30—60%), hernicellulose (20—40%)
`and lignin (10—30%). Both cellulose and hernicellulose con-
`sist of carbohydrate components whereas lignin is a highly
`branched aromatic polymer.
`Currently, there is intensive research on the use of lig-
`nocellulosic raw material as a biomass source for producing
`chemicals and fuels (as exemplified by many of the other
`articles in this special edition). However this research still
`faces considerable challenges due to lignocellulose being
`remarkably resistant towards hydrolysis and enzymatic attack
`(Peters, 2007). Energy demanding thermal pre—treatment oflig-
`nocellulose is necessary in order to break up the extremely
`stable cellu1ose—hemicellulose—lignin composites prior to
`adding cellulose-hydrolyzing enzymes and the current sit-
`uation does not allow the efficient use of lignocellulosic
`materials. Nevertheless, there is little doubt given the great
`abundance of lignocellulose that in the future this will become
`an attractive option. It is therefore important to continue to
`develop processes that can economically convert lignocellu-
`lose into chemicals. Moreover, glucose is one of the most
`abundant monosaccharides in biomass, accessible by enzy-
`matic or chemical hydrolysis from starch, sugar or cellulose.
`Furthermore, a range of chemical products can be obtained
`from glucose which gives it a key position as a basic raw mate-
`rial/building block.
`
`3.
`
`Glucose — a biorefinery building block
`
`Fermentation of polymer building blocks is already under
`commercial introduction. For example, Cargill produces lac-
`tic acid by fermentation and products based on polylactic
`acid are being introduced to the market. Several companies
`focus on succinic acid as a polymer building block, but also as
`a potential raw material for chemicals (e.g. butanediol). 1,3-
`propandiol is marketed by DuPont Tate & Lyle BioProducts
`for Soronam polytrirnethylene terephthalate (P'I'I‘) polyester.
`Likewise Cargill is working on developing 3—hydroxypropionic
`acid (3—HP). 3-HP is a potential raw material for existing
`chemicals such as propanediol and acrylic acid. Polyhydrox—
`yalkanoate (PHA) is marketed by Telles, a JN between ADM
`and Metabolix. Roquette, the French starch producer, has com-
`mercialized isosorbide, a derivative of sorbitol. Isosorbide is
`used as a co-monomer for high temperature polyethylene
`terephthalate. However, even if commercialization of polymer
`building blocks made by fermentation is commercially under-
`way, the technology has certain drawbacks such as loss of
`carbon as C02, low yields and difficult recovery of the products
`
`Table 1 - Process conditions for production of glucose from starch.
`
`Process
`
`Temperature (°C)
`
`Dry substance content (%)
`
`Jet cooking/dextrinization
`Saccharification
`
`105/95
`60
`
`30-35
`30-35
`
`pH
`
`5.2-5.6
`43-4.5
`
`Process time (11)
`
`0.1/1-2
`25-50
`
`

`
`1320
`
`chemical engineering research and design 8 7 ( 2 0 0 9 ) 1318–1327
`
`from the fermentation broth. The technology presented here
`(combined chemical and enzymatic catalysis from glucose)
`has the potential to overcome these problems and represents
`a promising next generation technology.
`One chemical transformation (besides fermentations) of
`carbohydrate monomers for the degradation of functional-
`ity is the dehydration reaction. This facilitates the removal of
`some of the functional groups in carbohydrates and allows the
`formation of defined building blocks. Triple dehydration of glu-
`cose yields HMF—a building block molecule that subsequently
`can be transformed into a multitude of bio-based chemicals.
`By a subsequent hydration reaction or an oxidation, HMF can
`be converted into levulininc acid or FDA, respectively. Both
`of these molecules are on the list of the 12 bio-based plat-
`form chemicals identified as being of highest potential to be
`converted into new families of useful molecules (Werpy and
`Petersen, 2004). In the following we will focus on the dehydra-
`tion of glucose to HMF as an example of the need to efficiently
`combine enzymatic aqueous processes with inorganic het-
`erogeneous catalytic processes that have so far mainly been
`developed for running reactions within the petrochemical
`industry.
`HMF is in itself a rather unstable molecule. It can be found
`in natural products such as honey and a variety of heat pro-
`cessed food products formed in the thermal decomposition
`of carbohydrates. Interestingly, HMF can be chemically con-
`verted into a range of other valuable chemicals. The oxidation
`of HMF is of particular interest. Here, the ultimate objective
`is to obtain FDA as suggested by Schiwek et al. (1991). The
`diacid can be used as a replacement for terephthalic acid in the
`production of polyethylene terephthalate and polybutylene
`terephthalate (Gandini and Belgacem, 1997; Kunz, 1993) which
`was recently reviewed by Moreau et al. (2004). The partially oxi-
`dized compounds can also be used as polymer building blocks
`
`although these are more difficult to produce selectively. FDA is
`a chemically very stable compound. Its only current uses are
`in small amounts in fire foams and in medicine where it can
`be used to remove kidney stones.
`Several extensive reviews describing the chemistry of HMF
`and its derivatives have been reported (see Fig. 1). The most
`recent review focuses on chemical transformation of biomass
`to a variety of chemicals with particular emphasis on the
`dehydration of monosaccharides giving either furfural (from
`pentoses) or HMF from hexoses, respectively (Corma et al.,
`2007). Moreau et al. (2004) described the recent catalytic
`advances in substituted furans from biomass and focused
`especially on the ensuing polymers and their properties. A
`review by Lewkowski (2001) on the chemistry of HMF and its
`derivatives also appeared recently. Two other relevant reviews
`are from Cottier and Descotes (1991) and Kuster (1990).
`The mechanism for the dehydration of fructose to HMF has
`been interpreted to proceed via two different routes; either
`via acyclic compounds or cyclic compounds (Haworth and
`Jones, 1944; Kuster, 1990; Van Dam et al., 1986; Antal et al.,
`1990). Besides HMF, the acid-catalyzed dehydration can lead to
`several other by-products such as insoluble polymers, called
`humins or humic acids. In an industrial process it is very
`important to find the right process conditions that avoid the
`formation of humins as these, besides lowering the selectivity
`of the reaction, potentially can clog up your reactor or deacti-
`vate the heterogeneous catalysts.
`In spite of all the research carried out within this area an
`efficient way of producing HMF or its corresponding dicar-
`boxylic acid, FDA, still remains to be found. Traditionally,
`chemists have been struggling with finding an inexpensive
`way of producing pure HMF. Given the immense field of its
`application, it is interesting that relatively few of the listed
`reviews have described the challenges that might be faced in
`
`Fig. 1 – HMF as a precursor for a range of commercial chemicals.
`
`

`
`cr-rnnczu. ENGINERING RESEARCH AND DESIGN 8 7 (2 oo 9) 1318-1327
`
`1321
`
`Table 2 — Typical reaction conditions for immobilized glucose isomerase.
`
`Process
`Isornerization
`
`Temperature (°C)
`50-60
`
`Dry substance content (9%)
`40-50
`
`pH
`7-8
`
`Process time (11)
`0.3-3
`
`a biorefinery manufacturing HMF or its derivatives. The most
`likely bioreflnery scenario will not be restricted to one product
`but make a series of high and low value products (including
`fuel). This allows the biorefinery to shift focus from one prod-
`uct to another if the market changes. In the case of I-IMF or
`FDA production this means that producing purely HMF or FDA
`is not the ultimate target and side-strearns producing other
`valuable products besides HMF or FDA can actually be of ben-
`efit One potential by-product of value is levulinic acid. This is
`formed via a rehydration of HMF to give levulininc acid along
`with formic acid. Both of these molecules are valuable prod-
`ucts that are potentially worth isolating as side streams. In
`this respect the goal of completely selective dehydration may
`in the future be misplaced.
`The synthesis of HMF is based on the acid-catalyzed
`triple dehydration of C6-sugar monomers, mainly glucose and
`fructose. However, various polysaccharides have also been
`reported as I-IMF sources (Rapp, 1987). The most convenient
`method for the preparation of HMF is by dehydration of fruc-
`tose. When starting from ketohexoses (such as fructose) the
`dehydration reaction proceeds more efficiently and selec-
`tively. This can be explained by aldohexoses (such as glucose)
`only being able to enolyze to a low degree which is consid-
`ered the limiting step in the production of HMI-' from glucose.
`However, glucose is the favored source of HMF due to the
`lower cost of glucose compared to fructose. Fructose may be
`obtained by enzyme or acid-catalyzed hydrolysis of sucrose
`and inulin orby the isomerization of glucose to fructose. Inulin
`is a linear B-2,1 linked fructose polymer which is temiinated
`by a single glucose unit. It is found as a food reserve in a
`number of plants including Jemsalem artichoke and chicory.
`Industrially fructose is produced from glucose by the enzyme
`glucose isomerase (EC 5.3.1.5). The equilibrium conversion
`under industrial conditions is 50% making chromatographic
`separation necessary in order to obtain the industrial product
`of 55% fructose, which has sweetness similar to sucrose. Glu-
`cose isomerase is used industrially as an immobilized enzyme
`with typical reaction conditions as shown in Table 2.
`Commercial immobilized glucose isomerase preparations
`used in a packed column have half-lives between 100 and
`200 days. Most columns therefore last for more than 1 year
`and productivities are typically around 15 tons of syrup dry
`substancdkg immobilized enzyme.
`
`4.
`
`Case studies
`
`4.1.
`
`Case 1: conversion of glucose/fructose to HMI-'
`
`To date most of the work regarding the add-catalyzed con-
`version of fructose, and to a less extent glucose, into HMI-'
`has been carried out in aqueous reaction media. Obviously
`water being very abundant and non-hazardous is the pre-
`ferred solvent of choice when exploring green and sustainable
`chemistry. Furthermore water is a good solvent for dissolving
`the monosacdraride substrates (fructose and glucose) as well
`as the product, HMF. However the dehydration of fructose to
`yield HMF in aqueous media is hampered by a competitive
`rehydration process resulting in the by—products levulinic acid
`
`and formic acid. In addition soluble and insoluble polyrner—
`ization products (humins), that are thought to arise from the
`self- and cross-polymerization of HMF, fructose and other by-
`products seem to be more pronounced in an aqueous reaction
`medium than an organic one (Van Dam et al., 1986). Neverthe-
`less, several interesting papers have been published on the
`dehydration of fructose into HMF. The conversion of glucose
`into HMF is more difficult and as a result there are only a few
`publications on this process.
`
`4.1.1. Aqueous media
`Several mineral acids such as HCI, H2504 and H3P04 have
`been employed in the homogeneous catalyzed dehydration of
`fructose to yield HMF (Newth, 1951; Medniclr, 1962; Roman-
`Leshkov et al., 2006). So far, however, the yield and selectivity
`of reactions carried out in aqueous reaction media are not
`comparable to those observed in aprotic high—boiling organic
`solvents such as DMSO where the solvent also serves as
`
`the catalyst (Musau and Munavu, 1987). Despite high yields
`and selectivity, the cost of removing high—boiling solvents
`makes these solvents unsuitable for indusnial and large-scale
`processes. Heterogeneous catalysts have, due to separation
`and recycling considerations, drawn more attention than
`homogenous catalysts. The use of various acidic heteroge-
`neous catalysts such as niobic acid (Nb205-nH20) and niobium
`phosphate (Nb0PO4) have been reported to have an intermedi-
`ate selectivity of about 30% for the production of HMF at about
`80% conversion offructose (Camiti et al., 2006). Zirconium and
`titanium phosphates/pyrophosphates have been shown to
`have a very high selectivity ofup to 100% at 100 °C in a period of
`18 min for the formation of HMI-' in water. However as the reac-
`
`tion time increases, the selectivity drops fast which is thought
`to be due to the formation of polymeric by-products. Addi-
`tionally, titanium oxides (T102), zirconium oxides (Zr02) and
`H-forrn zeolites catalyze the dehydration reaction (Moreau et
`al., 1996). Especially interesting is the direct conversion of glu-
`cose to HMF which can be enhanced up to 5-fold compared
`to the hydrothermal dehydration, by employing an a—TiO2
`at 200°C (Watanabe et al., 2005a,b). The main disadvantage
`with these catalysts seems to be the high temperature needed
`in order for the reaction to proceed without limited selec-
`tivity and conversion rates. Highly acidic cation-exchange
`resins such as those derivatized with sulfonic acid groups are
`also effective catalysts, providing the acidity of mineral acids
`together with the advantages of the heterogeneous catalysts
`(Rigal et al., 1981). These, often polystyrene based resins, can
`only tolerate temperatures up to around 130 “C, which reduces
`the range oftheir application. However this temperature range
`seems to be sufficient to overcome the activation energ bar-
`rier, when simultaneously applying the effect of microwave
`heating (Qi et al., 2008).
`
`4.1.2. Modified aqueous media and two-phase systems
`Phase modifiers have within the last couple of years proved
`very effective in promoting the conversion of fructose to HMI-'.
`Polar organic solvents that are miscible with water are added
`in order to increase the rate of the reaction to HMF and reduce
`
`the rate of the rehydration process forming by—products (Van
`
`

`
`1322
`
`chemical engineering research and design 8 7 ( 2 0 0 9 ) 1318–1327
`
`Dam et al., 1986). Commonly employed aqueous phase mod-
`ifiers are acetone, DMSO and polyethylene glycol (PEG) (Qi et
`al., 2008; Chheda et al., 2007; Van Dam et al., 1986). A further
`modification of the aqueous phase system is the introduction
`of a second immiscible phase to create a two-phase reaction
`system. An organic phase extracts the HMF from the aqueous
`phase as it is produced and consequently reduces the forma-
`tion of rehydration and polymeric by-products. Even with an
`initial concentration of fructose as high as 50 wt%, remark-
`able results with selectivity of 77% and a conversion of 90% at
`◦
`180
`C with HCl as the catalyst have been reported. In compar-
`ison similar conditions in water resulted only in a selectivity
`of 28% and a conversion of 51% (Román-Leshkov et al., 2006).
`
`4.1.3. Non-aqueous organic solvents
`Until now, the best results for the dehydration of fructose to
`HMF have been made in high-boiling organic solvents. The low
`concentration of water prevents the rehydration of HMF to lev-
`ulinic acid and formic acid. Iodine catalyzes the dehydration
`◦
`of the fructose part of sucrose in anhydrous DMF at 100
`C.
`Glucose is unaffected under the same conditions (Bonner et
`al., 1960). High selectivity has also been obtained when using
`PEG-600 as a solvent together with catalytic HCl. With the
`acid present a 1:1 solution of fructose and PEG-600 can be
`◦
`obtained at 85
`C (Kuster and Laurens, 1977). The first really
`high yields were reported by Nakamura and Morikawa (1980)
`using a strongly acidic ion-exchange resin as the catalyst in
`◦
`DMSO at 80
`C. These conditions gave a yield of 90% after
`8 h. The rate of the reaction was strongly affected by the type
`of resin used (Nakamura and Morikawa, 1980). Quantitative
`yields, without the use of a catalyst, were reported soon after
`◦
`in DMSO at 100
`C for 16 h (Brown et al., 1982). Good results
`were also obtained during an investigation of the optimum
`fructose concentration in DMSO. With 8.5 molar equivalents
`of DMSO with respect to fructose, a yield of 92% was obtained
`◦
`at 150
`C without any catalyst after 2 h (Musau and Munavu,
`1987).
`None of the above examples are suitable for production
`on a large-scale. High-boiling aprotic solvents such as DMSO,
`DMF and NMP are all miscible with water as well as many
`other common organic solvents. This makes separation of the
`desired products very difficult. Furthermore, both DMF and
`NMP are considered to be teratogenic.
`
`Supercritical/subcritical solvents
`4.1.4.
`Since the best results for the dehydration of hexoses to HMF
`have been in high-boiling organic solvents, the use of low-
`boiling solvents in their sub- or supercritical state would be
`an interesting alternative. Subcritical water has emerged in
`recent years as a feasible alternative to organic solvents at
`larger scale. Its unique intrinsic acidic and basic properties,
`makes it particularly interesting as a reaction medium for the
`dehydration of carbohydrates. When glucose is dehydrated in
`pure subcritical water, HMF is formed with greater selectivity
`than when using sulfuric acid or sodium hydroxide as cata-
`lysts under the same pressures and temperatures (Simkovic
`et al., 1987). Watanabe et al. (2005a) explored the use of differ-
`ent TiO2 and ZrO2 catalysts in highly compressed water. The
`anatase-TiO2 catalyst showed both basic and acidic properties
`and catalyzed the conversion of glucose to HMF. Yields were
`only about 20%, but the selectivity was more than 90%. The
`basic properties of the catalyst were thought to catalyze the
`isomerization of glucose to fructose, whereas the acidic prop-
`erties were thought to catalyze the dehydration (Watanabe et
`
`al., 2005b). Yields of up to 50% were obtained when using fruc-
`tose as the starting sugar and different zirconium phosphates
`as catalysts in subcritical water. No rehydration products
`were observed, yet the highest selectivity was not more than
`61%. By-products were humins and furaldehyde (Asghari and
`Yoshida, 2006). Interesting results have recently been reported
`on the catalytic effect of H3PO4, H2SO4 and HCl in the direct
`conversion of glucose to HMF in water at 523 K. It was con-
`cluded that the weakest acid, H3PO4, was the best catalyst for
`the conversion of glucose into HMF and the strongest acid,
`HCl, was the best catalyst for the conversion of HMF to lev-
`ulinic acid. The best yield for HMF was 40% (Takeuchi et al.,
`2008). More extensive studies on the kinetics of the dehydra-
`tion of d-glucose and d-fructose in sub- and supercritical water
`have been made as well as the behavior of HMF under similar
`conditions (Kabyemela et al., 1999; Asghari and Yoshida, 2007;
`Chuntanapum et al., 2008).
`Nevertheless, the overall results from sub- and supercrit-
`ical water have so far been unsatisfactory in terms of yields.
`Bicker et al. (2003) explored other low-boiling solvents such
`as acetone, methanol and acetic acid. An acetone/water mix-
`◦
`ture at 180
`C and 20 MPa gave 99% conversion of fructose
`and a selectivity of 77% to HMF. This excellent result was
`explained by the structural similarities between acetone and
`DMSO, which would promote the furanoid form of fructose
`and hence favor the formation of HMF. The authors also pro-
`pose a continuous process for the reaction (Bicker et al., 2003,
`2005).
`
`Ionic liquids
`4.1.5.
`Another attractive alternative to high-boiling organic solvents
`is the use of ionic liquids. Their unique physical properties
`such as negligible vapor pressure and non-flammability make
`them particularly suitable as solvents for large-scale produc-
`tion. There is a possibility to design and functionalize the
`ions of the ionic liquid, giving them ability to work both as
`solvent and reagent for certain reactions. There are several
`examples of ionic liquids that have the ability to solubilize nat-
`ural polymers such as cellulose, starch and chitin. This opens
`an excellent opportunity to convert crude biomass into fine
`chemicals (Liu et al., 2005; El Seoud et al., 2007).
`The first dehydrations of fructose and glucose with the help
`of ionic liquids date back 25 years. Fructose was dehydrated
`in the presence of pyridinium chloride to HMF in high purity
`with 70% yield. The corresponding result for glucose was only
`5% (Fayet and Gelas, 1983). In 1-butyl-3-methylimidazolium
`tetrafluoroborate and 1-butyl-3-methylimidazolium hexaflu-
`orophosphate, yields up to 80% from fructose were obtained
`using DMSO as a co-solvent and Amberlyst-15 resin as the
`catalyst. The DMSO helped to solubilize the starting fructose
`and the reaction was faster than in DMSO alone. Performing
`the reaction in 1-butyl-3-methylimidazolium tetrafluorobo-
`rate alone gave a yield of 50% within 3 h (Lansalot-Matras
`and Moreau, 2003). The best results so far from fructose were
`made by using the acidic 1-H-3-methylimidazolium chloride
`as reaction medium. This acted both as solvent and cata-
`◦
`lyst giving a yield of 92% after 15–45 min at 90
`C. There was
`no sign of HMF decomposition and glucose remained com-
`pletely unreacted (Moreau et al., 2006). Recently remarkably
`good results were found using the ionic liquid 1-ethyl-3-
`methylimidazolium chloride together with CrCl2, giving a total
`yield of 70% HMF directly from glucose and virtually no lev-
`ulinic acid. The authors propose that the actual catalytic
`−
`specie is the CrCl3
`ion formed together with the solvent
`
`

`
`chemical engineering research and design 8 7 ( 2 0 0 9 ) 1318–1327
`
`1323
`
`based on fructose (Halliday et al., 2003). Carlini et al. (2005)
`reported that HMF, as a starting reagent or produced one pot
`from fructose, was oxidized to the corresponding dialdehyde
`in water with methylisobutylketone (MIBK), as well as pure
`organic solvents, with vanadyl phosphate (VPO) based cata-
`lysts (Zr, Nb, Cr, Fe modified) as such or using a TiO2 support
`◦
`at 75–200
`C and 1 MPa. However, the reported yields were
`low (H20:MIBK = 0:30–5:30, HMF conversion 3–10%, selectivity
`to DFF 100–60%, respectively). Considering the oxidation as a
`stand-alone reaction and changing the solvents to less polar
`ones (benzene, toluene) better conversion rates and selectiv-
`ity were obtained, and using MIBK as a solvent lead to 98%
`conversion with 50% selectivity. However, in DMF the results
`◦
`are even better (at 150
`C) giving 84% conversion and 97%
`selectivity.
`
`4.2.2. Oxidation of HMF to FDA
`The above-described DFF may either be used as a valuable by-
`product or as an intermediate for obtaining FDA. On the other
`hand, catalytic reactions leading to the formation of FDA are
`also reported.
`Partenheimer and Grushin (2000) obtained DFF from HMF
`using metal bromide catalysts (Co/Mn/Zr/Br). The reactions
`were carried out in acetic acid at atmospheric pressure and
`also at 70 bar; the yields were 57% and 63% with the conver-
`sion of HMF 98% and 92%, respectively. Cobalt as a catalyst
`was also used by Ribeiro and Schuchardt (2003). Using cobalt
`acetylacetonate as a bi-functional acidic and redox catalyst
`◦
`encapsulated in silica in an autoclave at 160
`C, they obtained
`FDA, from fructose via HMF formation, with 99% selectivity to
`FDA at 72% conversion of fructose. By in situ oxidation of HMF
`to FDA starting from fructose, Kröger et al. (2000) described a
`way of producing FDA via acid-catalyzed formation and sub-
`sequent oxidation of HMF in a MIBK/water mixture using solid
`acids for fructose transformation and PtBi-catalyst encap-
`sulated in silicone and swollen in MIBK. The reaction was
`carried out in a reactor divided with a PTFE-membrane in
`order to prevent the oxidation of fructose. However, though
`in principle the integration process has been described, the
`yields remain quite low. The resulting yield of FDA was 25%
`based on fructose. In the oxidation of HMF to FDA the use
`of noble metals was first studied by Vinke et al. (1991). Here,
`mainly Pd, Pt, Ru supported on different carriers were used as
`the aerobic oxidation catalysts. Although all the noble met-
`als revealed catalytic activities, only Pt supported on Al2O3
`remained stable and active and gave quantitative yields of
`FDA. The reactions were carried out in water at pH 9 using
`◦
`a reaction temperature of 60
`C and a partial oxygen pressure
`of 0.2.
`
`4.2.3. Oxidation of HMF to FDA derivatives
`A new approach to the oxidation of HMF has been reported
`recently by Taarning et al. (2008) using methanol as both
`solvent and reagent. They performed a reaction with a gold
`◦
`nanoparticle catalyst in an autoclave at 130
`C and 4 bars of
`dioxygen, and obtaining FDA with 98% yield (according to GC
`analysis) and 60% isolated yield after sublimation.
`
`5.
`
`Process technology
`
`Table 3 indicates some of the key features of possible routes for
`the conversion of fructose to HMF. A number of observations
`can be made:
`
`Fig. 2 – Oxidation of HMF to DFF and FDA.
`
`and that it catalyzes the isomerization of ␤-glucopyranose
`to fructofuranose, which is subsequently dehydrated to HMF
`(Zhao et al., 2007). Bao et al. (2008) concluded that ionic liquids
`with a Lewis acid moiety were more efficient than those with a
`Brønsted acid counterpart when dehydrating fructose. These
`ionic liquids were also successfully immobilized on silica, giv-
`ing a yield of up to 70% from fructose to HMF and completely
`retained their catalytic activity after five reaction cycles (Bao
`et al., 2008).
`
`Case 2: HMF oxidation to 2,5-diformylfuran and
`
`4.2.
`FDA
`
`FDA has been identified by the U.S. Department of Energy
`(DOE) biomass program as one of the 12 chemicals that in the
`future can be used as a feedstock from biomass in biorefiner-
`ies (Werpy and Petersen, 2004). Due to the presence of the
`two carboxylic acid groups, FDA is considered to be a biore-
`newable building block to form polymers from biomass and
`therefore become an alternative to terephthalic, isophthalic
`and adipic acids, which are all produced from fossil fuels. Sug-
`ars in the form of mono- and disaccharides are easily available
`from biomass. The hexose type monosaccharides such as glu-
`cose and fructose can be catalytically dehydrated into HMF
`(Corma et al., 2007; Gallezot, 2007; Moreau et al., 2004). HMF
`can then be oxidized into FDA using a variety of routes and
`reaction types with stochiometric amount of oxidants. Most of
`them are described in a review by Lewkowski (2001), including
`electrochemical oxidation, use of barium and potassium per-
`manganates, nitric acid and chromium trioxide. In this section
`we will focus on the recently reported catalytic routes for the
`oxidation of HMF into FDA.
`
`4.2.1. Oxidation of HMF to DFF
`Though production of FDA from HMF has been of great interest
`recently, there are few papers on catalytic aerobic oxidation of
`HMF. In the catalytic route to form FDA the partially oxidized
`intermediate 2,5-diformylfuran (DFF) is often observed (Fig. 2).
`The dialdehyde is a useful product to form other deriva-
`tives, and a number of studies have reported on the selective
`formation of DFF. Thus, Halliday et al. (2003) reported oxida-
`tion of HMF to DFF using an in situ reaction protocol where
`HMF was directly generated from fructose and not isolated.
`Hence, using ion-exchange resins and, then, VOP-type cata-
`lysts the authors obtained DFF with a maximum yield of 45%
`
`

`
`1324
`
`CHEMICAL ENGINEERING RESEARCH AND DESIGN 8 7 (2 0 o 9) 1318-1327
`
`Table 3 — Key features of possible routes for the conversion of fructose to I-IMF.
`
`Mode of
`operation“
`
`Catalyst”
`
`Temp.
`
`Fructose
`concentration
`
`Solvent media‘
`
`Highest
`yield
`
`Reference
`
`13
`B
`B
`B
`B
`
`B
`
`B
`
`B
`B
`C
`B
`C
`
`Hetero.
`Homo.
`Homo.
`Hetero.
`Hetero.
`
`Homo.
`
`80°C
`170°C
`90°C
`165 "C
`80°C
`
`180°C
`
`Hetero.
`
`90°C
`
`Hetero.
`
`Hetero.
`Hetero.
`
`110°C
`100°C
`85 °G
`100°C
`165 °c
`
`3-4% (wlw)
`10% (wlw)
`3—50% (wlw)
`10% (wlw)
`6% (wlw)
`3% (wlw)
`30% (wlw)
`50% (wlw)
`10% (wlw)
`30% (wlw)
`6—10% (wlw)
`6—10% (wlw)
`10-20%. (wlw)
`6% (wlw)
`05-3.5% (wlw)
`
`Water, MIBK
`Water, DMSO, MIBK, 2-butanol, DCM
`I-IMIM*CI‘
`Water, MIBK
`Wafef
`
`Water, DMSO, PVP, MIBK, 2-butanol
`
`Water, DMSO, PVP, MIBK, 2-butanol
`
`Water
`water, MIBK
`Water
`Water
`Water, MIBK
`
`41%
`87%
`92%
`69%
`42%
`59%
`76%
`71%
`59%
`54%
`31%
`74%
`26%
`85%
`-
`
`Carlini et a1. (2005)
`Chheda et al. (2007)
`Moreau et al. (2006)
`Moreau et al. (1996)
`Cartini et al. (2004)
`
`Roman-Leshkov et al. (2006)
`
`Roman-Leshkov et a1. (2006)
`
`Carlini et al. (1999)
`
`Benvenuti et al. (2000)
`Rivalier et al. (1995)
`
`‘ Process is continuous (C) or batch (B).
`" Catalyst is homogenous (homo.) or heterogenous (hetero).
`C Solent media are: methylisobutylketone (MIBK), dimethyl sulfoxide (DMSO), poly(1-vinyl-2-pyrrolidinone) (PVP), dicholorrnethane (DCM), and
`1-H-3-methyl imidazolium chloride (HMIM*C1').
`
`o Catalysttype
`A variety of catalysts like mineral and organic acids,
`salts, and solid acid catalysts such as ion-exchange resins
`and zeolites have been used in the dehydration reaction.
`The homogeneous acid-catalyzed processes are frequently
`associated with low selectivity (30—50%) for HMI-‘ at a
`relatively high conversion (50—70%) (Carlini et al., 1999).
`Moreover, problems related to separation and recycling of
`the mineral acid as well as of plant corrosion are expected.
`Thus, recent research has been based on heterogeneous
`acid catalysts which have considerable potential for indus-
`trial application (Carlini et al., 1999).
`o Mode of operati

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket