13

Chemical routes for the conversion of cellulosic platform molecules into high-energy-density biofuels

J.A. Melero, J. Iglesias, G. Morales,  and M. Paniagua     Universidad Rey Juan Carlos, Mostoles, Madrid, Spain

Abstract

In this chapter, we will highlight different chemical routes to convert highly functionalized sugars (glucose and xylose) coming from the hydrolysis of nonedible lignocellulosic biomass into high-energy-density biofuels with reduced oxygen content and improved H/C ratio. Firstly, this chapter will be focused on the chemical catalytic upgrading of lignocellulosic-based platform molecules (5-HMF, levulinic acid, and furfural) to oxygenated second-generation biofuels and the effect of the blending with conventional fuels. Then, the latest developments in the effective transformation of oxygenated platform molecules into liquid hydrocarbon fuels will also be discussed. This valorization strategy is based on the combination of oxygen removal processes and CC coupling reactions to increase the molecular weight with the aim of yielding a mixture of liquid hydrocarbons with a carbon number in the range of conventional fuels. Both type of advanced biofuels are shown as a good alternative to conventional biofuels (biodiesel and bioethanol).

Keywords

5-HMF; Biofuels; Chemical valorization; Furfural; GVL; Levulinic acid; Lignocellulosic biomass

13.1. Introduction

Greening of air and land transport is one of the key environmental objectives for this century. This concern is driving the chemical industry toward the search for new sustainable and efficient alternatives that can substitute fossil sources. Lignocellulosic biomass is abundant, and it has the potential to significantly displace petroleum in the production of fuels and valuable chemicals. Lignocellulose must be separated into its constituents (lignin (15–30%), cellulose (35–50%) and hemicelluloses (25–30%)) and depolymerized to their corresponding building blocks. The building blocks of lignin are aromatic alcohols, but controlled lignin depolymerization is rather difficult on a technical scale and this problem has not yet been solved. Controlled cellulose depolymerization results in glucose, while the hemicellulose is depolymerized to a mixture of different sugars, mostly pentoses. These sugars are the key molecules for the production of the different platform molecules, such as hydroxymethylfurfural (5-HMF), furfural, and levulinic acid (Chatterjee et al., 2015; Bohre et al., 2015).
On the other hand, limitations of conventional biofuels (biodiesel and bioethanol) and new trends in legislation have stimulated the research for new technologies that allow high-energy-density, infrastructure-compatible fuels (advanced biofuels) which could be easily implemented in the existing hydrocarbon-based transportation infrastructure (eg, engines, fueling stations, distribution networks, and petrochemical processes) and, more importantly, not using edible biomass for their production.
An interesting approach for the use of platform molecules in the field of renewable fuels (advanced biofuels) is their transformation into oxygenated compounds which can be used as blending components in the reformulation of conventional fuels (gasoline and diesel) and in some cases improving some properties such as cold flow behavior and octane number as well as emissions reductions (Climet et al., 2014). Moreover, catalytic transformation of platform molecules into liquid hydrocarbon fuels is also shown to be an interesting approach. However, this strategy requires the removal of oxygen from the highly oxygenated platform molecules (by means of a great variety of reactions including dehydration, hydrogenolysis, hydrogenation, decarbonylation/descarboxylation, etc.) and CC coupling reactions (aldol condensation, ketonization, oligomerization, etc.) to increase the molecular weight with the aim of yielding a mixture of liquid hydrocarbons with a carbon number in the range of conventional fuels (De et al., 2015).
These processes involve a multistep transformation from the carbohydrate fraction to the value-added products which makes most of them far from commercialization. Hence, intensive efforts are still required to enable scale up of synthetic protocols developed on a lab-scale into industrial processes. Some of the current drawbacks might be overcome by the one-pot transformation of lignocellulose carbohydrates in value-added chemicals without isolation of the intermediate platform molecules (Delidovich et al., 2014). Moreover, nanoporous materials, such as acidic, basic or metallic catalysts (zeolites, mesoporous silicas, microporous/mesoporous carbons, resins, metal oxides, etc.), will play a crucial role in this biomass transformation (Wang and Xiao, 2015).
A variety of catalytic routes have been described in recent years for the chemical transformation of carbohydrates into hydrocarbon liquid fuels and oxygenated biofuels which will be discussed in this chapter. Special focus will be brought to the recent progress of integrated processes based on the use of multifunctional catalytic systems without isolation of the platform intermediates. Fig. 13.1 summarizes the main chemical routes for the conversion of cellulosic platform molecules into high-energy-density biofuels.

13.2. Oxygenated fuels via 5-HMF: furanic compounds

A large number of works described in the literature deal with the production of 5-HMF from the dehydration of hexoses (glucose and fructose) under a wide range of reaction conditions and over different acid catalytic systems. However, in general, yields to 5-HMF from monossacarides other than fructose are considerably low. Unfortunately, no full-scale commercial plants for 5-HMF manufacture have been set up yet (Mukherjee et al., 2015). However, 5-HMF is not an attractive fuel component by itself due to its lack of chemical stability, but it can be used as a starting point to produce a variety of furan derivatives with interesting properties to be used as energy-dense fuels. In this section, we will discuss different strategies to convert 5-HMF into oxygenated compounds that are suitable for fuel formulation.

13.2.1. 2,5-Dimethylfuran

One of the most promising furanic compounds to be used as a liquid transportation fuel is 2,5-dimethylfuran (DMF), which has received increasing attention in recent years (Quian et al., 2015). The most general method to obtain DMF is by reduction of both the formyl and hydroxyl groups of 5-HMF using supported metal catalysts in organic solvents such as butanol. The hydrogenolysis of 5-HMF has been the most successfully achieved when using multifunctional catalytic systems than can both hydrogenate and deoxygenate without excessively opening the furan ring (Zu et al., 2014; Huang et al., 2014; Nishimura et al., 2014; Nakawaga et al., 2014; Wang et al., 2014; Yang et al., 2015). However, despite these improvements, this strategy is limited to the isolation with good yields of the 5-HMF product.
image
Figure 13.1 Chemical transformations of biomass into valuable fuels.
Several works have been addressed to the production of DMF directly from biomass-derived hexoses, especially fructose. In this sense, the breakthrough of deriving DMF in good yields (76–79%) from fructose was first reported by Roman-Leshkov et al. (2007) in a two-step process. The process consisted of the production of 5-HMF starting from fructose (in an acid-catalyzed biphasic reactor to promote the simultaneous dehydration of fructose to 5-HMF and the solvent-extraction of the produced 5-HMF) followed by hydrogenation over a carbon-supported copper/ruthenium (Cu-Ru/C) catalyst using butanol as a solvent. From this pioneering work, similar works have been described in the literature using different solvents and catalysts. Recently, Upare et al. (2015) proposed an integrated process for the production of DMF from fructose with an outstanding yield of 92% in a two-step heterogeneous process in which fructose is firstly dehydrated to 5-HMF using Amberlyst-15 in butanol, and the 5-HMF in the resulting mixture is then converted to DMF by vapor-phase hydrogenolysis over an Ru-Sn/ZnO catalyst.
Another interesting approach to transform fructose into DMF in a one-pot process has been reported by Thananatthanachon and Rauchfuss (2010). The process is based on the use of formic acid, which can also be conveniently produced by biomass degradation, both as homogeneous acid catalyst for the dehydration of fructose into 5-HMF and hydrogen source to turn 5-HMF into the intermediate 2,5-dihydroxymethylfurfural (DHMF); finally acting again as a catalyst for the DMFH deoxygenation to give DMF. An overall yield of 51% was achieved, which is unsatisfactory for industrial applications. Other approaches based on hydrogen donors have been described in the literature but the DMF yields have been reduced.
Considering the use of glucose instead of fructose, a two-step approach for its conversion into DMF was also attempted in ionic liquids (ILs) in combination with acid catalysts (Chidambaram and Bell, 2010). This process involves the dehydration of glucose to 5-HMF in [EMIM]Cl and acetonitrile using 12-molybdophosphoric acid (12-MPA) as catalyst, followed by the subsequent conversion of 5-HMF into DMF over Pd/C in a one-pot method. However, such a system provides low 5-HMF conversion and poor DMF selectivity (13%), needing very high hydrogen pressures (62 bars) owing to its low solubility in ILs.

13.2.2. Ethers of 5-HMF: ethoxymethyfurfural

Ethers of 5-HMF and in particular 5-(ethoxymethyl)furfural (EMF), the main representative of the 5-alkoxymethylfurfural ethers family, are considered as excellent additives for diesel. The most common route for the production of EMF has been the etherification of 5-HMF with ethanol using solid acid catalysts (Lanzafame et al., 2011). While relatively high EMF yields can be obtained using this approach, the direct use of 5-HMF as precursor for the preparation of EMF is not industrially interesting.
A more attractive reaction pathway is the one-pot combination of the dehydration of a cheap and renewable source, such as fructose, to 5-HMF followed by its etherification into EMF using ethanol as solvent and a heterogeneous catalyst. Both transformations are driven by acid catalysis, being feasible to optimize the selectivity toward EMF through the proper selection of the solid acid catalyst and the reaction conditions. In a pioneering work, Brown et al. (1982) evaluated the preparation of ethers of 5-HMF, together with 5-HMF itself and alkyl levulinates, from fructose using ion-exchange resins in nonaqueous solvents. However, both the selectivity to EMF and the reaction rates were low. More recently, different catalytic systems have been used in a similar way (Balakrishnan et al., 2012; Liu et al., 2012; Kraus and Guney, 2012; Li et al., 2014a; Yuan et al., 2015).
Although fructose has been mainly explored, it is clear that the large-scale sustainable use of EMF will require cellulosic biomass, ie, glucose, as the feedstock. However, catalytic studies report that no 5-HMF or EMF were detected, and the glucose instead reacts with ethanol to produce ethyl glucopyranoside (EDGP). Brönsted acid catalysts on their own appear not to be able to drive the transformation of glucose into EMF. Consequently, to produce EMF from glucose, isomerization of glucose into fructose appears as a particularly relevant reaction. Lew et al. (2012) have reported the one-pot synthesis of EMF from glucose using a combination of Sn-BEA (for the isomerization step) and Amberlyst 131 (for the dehydration-etherification steps) with acceptable yields.

13.2.3. Ester of 5-HMF: acetoxymethyfurfural

In a similar way, the use of carboxylic acids or their anhydrides instead of ethanol can give 5-HMF esters, also with interesting properties as biofuels or fuel additives. The acetyl ester (acetoxymethylfurfural or AMF) has high energy content. AMF can be produced by esterification of biomass-derived 5-HMF with acetic acid or anhydride, using biocatalysts (lipases) (Krystof et al., 2013) as well as different homogeneous and heterogeneous catalysts such as sulfuric acid, metal chloride, and transition metals. However, very few works have been focused on the production of AMF from carbohydrates. As an example, Bicker et al. (2005) reported the production of AMF from fructose in subcritical acetic acid (180°C, 20 MPa). Although acetic acid may also act as a catalyst, the authors used a small amount of sulfuric acid as catalyst, leading to 38% AMF yield in a continuous high-pressure reactor.

13.3. Levulinic acid as platform molecule to oxygenated fuels: alkyl levulinates and valeric biofuels

Levulinic acid (LA) is a compound derived from 5-HMF and listed among the top 12 most promising value-added chemicals from biomass. This platform molecule is formed by dehydration in acidic media of hexoses to 5-HMF and subsequent hydration produces LA, formic acid, along with other unwanted polymerized products (humins) (Mukherjee et al., 2015). Likewise, LA can also be obtained by the hydrolysis of furfuryl alcohol (see Fig. 13.1).
One of the most promising processes for the large-scale continuous conversion of lignocellulosic biomass to levulinic acid is the Biofine process. This process involves a two-stage high-temperature homogeneous acid-catalyzed hydrolysis process with a yield of 75% from lignocellulosic waste biomass. The use of inexpensive lignocellulosic wastes allows competitive production costs of LA (0.06–0.18 euros per kg) to be used as precursor for the production of biofuels. In this section, we will review the different processes to upgrade levulinic acid into biofuels.

13.3.1. Esterification: alkyl levulinates

Low-alkyl levulinates, and particularly ethyl levulinate (EL), can be advantageously used as oxygenated biofuels. They can be easily obtained by esterification of levulinic acid with a lower alcohol such as ethanol. Such reaction occurs even at room temperature, but the rate is very low and needs to be accelerated by either increasing temperature or using a catalyst. Although traditional homogeneous acid catalysts, such as sulfuric, polyphosphoric or p-toluenesulfonic acids have shown good performances in this reaction, the use of solid (heterogeneous) acid catalysts is more desirable. A wide range of Brönsted solid acid catalysts (mesoporous H4SiW12O40–SiO2; acid zeolites; sulfated oxides; sulfonic acid-functionalized SBA-15 silicas; sulfated mesoporous zirconosilicates; sulfonated hydrothermal carbons; bimodal micro-mesoporous H/BEA zeolite; Zr-containing MOFs; dodecatungstophosphoric acid supported on desilicated H-ZSM-5; sulfated carbon nanotubes, etc.) have been successfully used in the esterification of levulinic acid with ethanol (Pascuale et al., 2012; Fernandes et al., 2012; Melero et al., 2013; Yan et al., 2103; Nandiwale et al., 2013; Kuwahara et al., 2014; Pileidis et al., 2014; Patil et al., 2014; Oliveira and da Silva, 2104; Cirujano et al., 2015).
However, the above-discussed routes would necessitate a source of high-purity levulinic acid (nontrivial to obtain due to contamination by polymeric humins). Therefore, it would be more advantageous if levulinates could be obtained in high yields from C6 carbohydrate-based biomass, ie, fructose, glucose, sucrose, cellulose and the like, in alcohol medium without the need for first isolating levulinic acid. As an additional advantage, the use of a nonaqueous alcoholic reaction medium for the treatment of lignocellulosic biomass minimizes the formation of undesired byproducts, since polymeric humins formation greatly diminishes in alcohol. Although homogeneous catalytic systems, such as sulfuric acid, have been employed, the research interest is currently focused on heterogeneous catalysts that can avoid the known problems of homogeneous acids as well as improve the selectivity to EL. Thus, Peng et al. (2011) reported the production of EL from glucose using different sulfated catalysts, including SO4/ZrO2, SO4/TiO2, SO4/ZrO2-TiO2, and SO4/ZrO2-Al2O3. They showed that different components of the sulfated metal oxides have markedly different catalytic effects on the ethanolysis of glucose. An optimized EL yield of above 30 mol% was obtained for sulfated zirconia at 200°C, leading to the side formation of diethyl ether (coming from the autoetherification of ethanol). On the other hand, SO3H-SBA-15 catalyst has shown a better catalytic activity than zeolites and sulfated zirconia catalysts for the conversion of fructose to ethyl levulinate (Saravanamurugan and Riisager, 2012). However, this type of Brönsted acid catalyst cannot drive the direct conversion of glucose to EL, instead leading to ethyl glucopyranoside (EDGP). This evidences that the activity of sulfated zirconia reported by Peng et al. (2011) for glucose should be attributed not only to the presence of sulfate groups but also of zirconium oxide, which is able to isomerize glucose into fructose. In the same line, a work by Tominaga et al. (2011) has showed that mixed-acid systems consisting of both Lewis (metal triflates such as In(OTf)3) and Brönsted (sulfonic) acids is an efficient combination for the direct synthesis of methyl levulinate from cellulose. In this case, the reaction proceeds in two steps; cellulose is first solvolyzed to sugars, which are readily converted to methyl levulinate. The former step is mainly catalyzed by sulfonic acids, and the latter by the metal triflates.
Current efforts are focused on the preparation of bifunctional catalysts capable of directing the Lewis acid-catalyzed isomerization of alkyl glucoside intermediates to alkyl fructosides, and their subsequent Brönsted acid-catalyzed dehydration to 5-alkoxymethylfurfural and esterification to form alkyl levulinate and formate (Fig. 13.2).
image
Figure 13.2 Reaction pathway for the acid-catalyzed conversion of glucose to ethyl levulinate in ethanol.
The correct balance of Lewis and Brönsted acid sites is critical to the success of this complex tandem transformation. In this context, we have recently reported the preparation of conformal sulfated zirconia (SZ) monolayers throughout an SBA-15 architecture that confers efficient acid-catalyzed one-pot conversion of glucose to ethyl levulinate (Morales et al., 2014). In this work, we have shown that conformal SZ monolayers with tunable surface acid strength and site density can be dispersed over a mesoporous SBA-15 framework through a simple wet chemical grafting/hydrolysis protocol. A bilayer SZ/SBA-15 material exhibits the maximum surface acidity and balance of Lewis:Brönsted sites, and exhibits good performance in the one-pot conversion of glucose to alkyl levulinates under mild conditions. A clear advantage of these catalysts is the capability of reaching relatively high EL yields under moderate temperatures, while avoiding ethanol losses as diethyl ether. Furthermore, the sulfated ZrO2 monolayers grafted on SBA-15 are highly stable, overcoming the extended leaching problems of commercial sulfated zirconias.

13.3.2. γ-Valerolactone and valeric biofuels

The hydrogenation of biomass-derived levulinic acid (LA) with either heterogeneous (Hengne and Rode, 2012; Wright and Palkovits, 2012) or homogeneous (Chalid et al., 2011; Li et al., 2012) catalysts is one of the most effective potential methods for the preparation of GVL. Such a reduction of LA to GVL generally takes place in the presence of molecular H2; though the use of in situ generated H2 coming from the decomposition of formic acid is also a promising alternative. Catalysts used for this transformation, either heterogeneous or homogeneous, typically consist of noble metals such as ruthenium, platinum, iridium, etc., providing excellent yields to GVL from LA.
However, as previously discussed for alkyl levulinates, this approach would necessitate a source of high-purity levulinic acid. Therefore, research studies have lately focused on obtaining GVL from carbohydrates in a one-pot approach through the combined action of acid and hydrogenation catalysts, avoiding the need for isolating LA. Heeres et al. (2009) reported a one-step strategy for producing γ-valerolactone directly from glucose and fructose, as well as sucrose and cellobiose, by combining a homogeneous acid catalyst (trifluoroacetic acid) with a heterogeneous hydrogenation catalyst (Ru/C). Reactions were performed in water at 180°C in the presence of hydrogen or formic acid as the hydrogen source. However, this one-pot approach still faces a serious problem of deactivation of the hydrogenation metal catalyst due to the presence of strong acids in the reaction medium. This deactivation is more severe when levulinic acid comes from the sulfuric acid-catalyzed deconstruction of cellulose. In order to increase the stability of the catalyst, the addition of Re as an alloy with Ru improved the catalyst stability in the presence of sulfuric acid, although the TOF was comparatively low. Another way to minimize this problem is the use of different systems for extracting levulinic acid from the acid aqueous solutions. For instance, alkylphenol solvents are able to extract up to 80% of levulinic acid from these aqueous feedstocks (Sen et al., 2012a). Also, the reactive extraction with different alcohols or with olefins; to produce levulinate esters which can be easily separated from the aqueous feedstock has been proposed (Gurbuz et al., 2011). These extracting systems allow obtaining levulinic acid (or an ester thereof) pure enough to be hydrogenated to GVL and the recycling of the mineral acid. Going a step further, the use of two heterogeneous catalysts can avoid the deactivation associated with mineral acids. Thus, the direct catalytic conversion of cellulose to levulinic acid (LA) by niobium-based solid acids and further upgrading to γ-valerolactone (GVL) on a Ru/C catalyst were performed through sequential reactions in the same reactor (Ding et al., 2014). Firstly, using aluminum-modified mesoporous niobium phosphate as a catalyst, cellulose can be directly converted to LA with as high as a 52.9% yield in aqueous solution, even in the presence of the Ru/C catalyst. Then, after replacing N2 with H2, the generated LA in the reaction mixture can be directly converted to γ-valerolactone through hydrogenation over the Ru/C catalyst without further separation of LA.
As a recent attractive alternative to the reduction of LA to GVL using H2 or formic acid over metal catalysts, the group of Dumesic has explored the reduction of LA by catalytic transfer hydrogenation (CTH) through the Meerwein–Ponndorf–Verley (MPV) reaction. In this reaction, a sacrificing secondary alcohol such as 2-butanol is used as a hydrogen source (Chia and Dumesic, 2011; Assary et al., 2013). This approach offers important advantages, such as an increased chemoselectivity for the reduction of carbonyl groups under milder reaction conditions, even in the presence of other functional groups; or the fact that the MPV reaction does not require precious metal heterogeneous catalysts. Moreover, the MPV hydrogen donor, usually 2-propanol, can be recycled after hydrogenation over base-metal catalysts such as nickel or copper, or even sold as a commodity chemical in its oxidized form (ie, ketones). Dumesic's group demonstrate in their work that CTH via MPV reaction is a viable means for the hydrogenation of LA and its esters (levulinates) over inexpensive, metal oxide heterogeneous catalysts that are easily recovered (ZrO2 was demonstrated to be a highly active material), with attainment of close to quantitative yields of GVL under appropriate reaction conditions. In the same direction, Román-Leshkov's group (Luo et al., 2014) have investigated the reaction kinetics of the MPV reduction of methyl levulinate (ML) to 4-hydroxypentanoates and subsequent lactonization to GVL catalyzed by Lewis acid Beta zeolites (modified with Hf, Ti, Zr, and Sn metal species). All catalysts generated GVL with selectivities >97%, with Hf-Beta exhibiting the highest activity. Likewise, Geboers et al. (2014) have demonstrated the feasibility of the CTH approach to convert levulinates into hydroxypentonaotes and GVL, using Raney Ni as catalyst and 2-propanol as H-donor and solvent.
On the other hand, GVL can also be used as an intermediate for the production of valeric biofuels. The process consists of the catalytic hydrogenation of GVL to valeric acid (VA) over hydrogenation metal catalysts and subsequent acid-catalyzed esterification to alkyl (mono/di) valerate esters. Lange et al. (2010) showed that although GVL is a relatively stable product under hydrogenation conditions, it can be hydrogenated to VA in the presence of bifunctional catalysts containing both hydrogenation and acidic functions. Using a continuous high-pressure plug-flow reactor they identified Pt-loaded SiO2-bound H-ZSM-5 as a very effective catalyst for the production of VA, though other zeolites and hydrogenation metals also gave good yields. The process can be intensified by converting LA to ethyl valerate (EV), the most interesting valeric biofuel, in a single step (going through the intermediate sequential production of GVL and VA). Thus, cofeeding ethanol with LA as a physical or chemical mixture (in the form of ethyl levulinate) over a Pt-modified zeolite-based catalyst leads to the efficient coproduction of VA and EV. More recently, Chan-Thaw et al. (2013) have proposed the production of EV and pentyl valerate (PV) in a one-pot one-step reaction from GVL. The bifunctional catalyst used consisted of Cu supported on an amorphous weakly acidic material, therefore representing an interesting alternative to Pt/zeolite catalysts.

13.4. Oxygenated fuels via furfural: furan derivatives

Furfural is industrially produced from the dehydration of C5 sugars (mainly xylose) using mineral acids as homogeneous catalysts. However, this process generates highly pollutant effluents, requires a lot of energy and gives reduced yields. Over the last few years, many studies have searched for new sustainable ways of producing furfural using heterogeneous catalysts, tuning the temperature, pressure, and solvent and exploiting different extracting techniques (Dutta et al., 2012). Much chemical, catalysis and engineering research is still needed to realize the potential of the furfural platform for biofuel manufacture. In this section of the chapter, we will review the present status of this platform molecule in producing second-generation biofuels (Lange et al., 2012).

13.4.1. Furfural hydrogenation toward oxygenated biofuels

Hydrogenation is an important reaction to transform furfural and its derivatives into potential biofuels with enhanced fuel properties. Hydrogenation of furfural includes the hydrogenation of the carbonyl group to hydroxymethyl or methyl and the hydrogenation of the furan ring (Climent et al., 2014). Different products are obtained depending on the type of catalyst and the reaction conditions (furfuryl alcohol (FA); 2-methylfuran (MF); 2-methyl-tetrahydrofuran (MTHF); tetrahydrofuran (THF)).
Furfuryl alcohol is an important chemical intermediate in the transformation of furfural into energy-dense oxygenated biofuels. It is one of the most common products in the hydrogenation of furfural. It has been estimated that 62% of the furfural produced globally each year is converted into FA (Yan et al., 2014). The hydrogenation of furfural to FA is relatively easy to achieve and become more mature over the last several decades development. However, at present, furfuryl alcohol may only be produced from xylose by a two-step process based on catalysts with different features and operating at different conditions. Recently, Perez and Fraga (2014), studied the one-pot production of furfuryl alcohol via xylose dehydration followed by furfural hydrogenation over a dual catalyst system composed of Pt/SiO2 and sulfated ZrO2 as metal and acid catalysts, respectively. They found that the presence of both acid and metal sites is compulsory in order to promote both reaction steps. Likewise, selectivity toward furfuryl alcohol is strongly dependent on the solvent, which can inhibit its polymerization to some extent.
The production of MF (sylvan) by furfural hydrogenation through the FA as an intermediate has been reported over various supported noble metal and bimetallic catalysts. Cu-based catalysts such as Raney-Cu, Cu/alumina, and carbon-supported Cu chromite have shown selective conversion of furfural to MF but often operated at high temperature and low pressure. However, catalyst deactivation is an important drawback of using those catalysts. Hence, increasing the catalyst stability and the development of an effective regeneration procedure is required (Yan et al., 2014). Some authors have recently performed the hydrogenation of furfural in different solvents under milder reaction conditions using supported Pd complex, achieving 100% yield of MF after 1 h of reaction (Climent et al., 2014).
The conversion of furfural into MTHF is achieved by a multistep process which includes hydrogenation–deoxygenation of furfural to MF and further hydrogenation of MF to MTHF in separated reaction systems. High H2 pressure is needed and the multistep conversion requires multicomponent catalysts, at least two different reactors, and the isolation systems for the intermediates. This limits the large-scale production of MTHF. There are few reports dealing with the one-step conversion of furfural to MTHF. Recently, one-step direct conversion of furfural to MTHF was carried out under atmospheric pressure over a dual solid catalyst based on two-stage-packed Cu–Pd in a reactor. This strategy provides a successive hydrogenation process, which avoids high H2 pressure, uses the reactor efficiently, and eliminates the product-separation step. Therefore, it could enhance the overall efficiency because of low cost and high yield of MTHF (97.1%) (Dong et al., 2015).
Finally, tetrahydrofuran (THF) can be obtained by decarbonylation of the carbonyl group of furfural under reductive conditions using Pd-based catalysts followed by hydrogenation of furan formed in the presence of a variety of metal catalysts (Sitthisa and Resasco, 2011).

13.4.2. Esters and ethers from furfuryl alcohol

Intrinsic energy of FA can also be upgraded by means of etherification with short-chain alcohols, transformation of alkyl-furfuryl ethers into alkyl levulinates and esterification with short alkyl carboxylic acids. In the following, we will discuss the most relevant processes.
Etherification of furfuryl alcohol is performed in the presence of strong acid catalysts in an alcoholic medium. In this transformation, it is important to control the reaction time and conditions. Severe reaction conditions (increasing of catalyst acidity and temperature) leads to heavy byproducts production coming from condensation reactions of furfuryl alcohol. Likewise, although alkyl levulinates can be easily obtained by direct esterification of levulinic acid with the corresponding alkyl alcohol in the presence of an acid catalyst (see Section 13.3.1), they can be advantageously prepared from furfuryl alcohol, without the need for isolating levulinic acid. The reaction mechanism involves the quickly etherification of furfuryl alcohol into the alkyl-furfuryl ether followed by a slowly transformation into alkyl levulinate. It is important to take into account that the transformation also involves the formation of byproducts such as humins and ethers (dialkyl ethers from the alcohol used as reaction media). Lange et al. (2009) reported a decrease in catalytic activity in the following order: H2SO4 > macroreticular resins > gel resins > zeolites, finding that the two major factors that settle the activity of catalysts are the acid strength of active sites and their accessibility.
Another approach is the esterification of furfuryl alcohol in the presence of carboxylic acids. Particularly, a process that can be considered a promising route is the direct production of furfuryl acetate in a one-step hydrogenation-esterification of furfural with acetic acid. Under moderate reaction conditions, a bifunctional catalytic system composed of an acid functionality for esterification reaction and a hydrogenating functionality (5% Pd/Al2 (SiO3)3 and 5% Pd/Al-SBA-15) has been proposed with good selectivity to the desired products (Yu et al., 2011a,b).

13.4.3. γ-Valerolactone from furfural

Direct synthesis of GVL from furfural involves hydrogenation steps and acid-driven transformations. However, for large-scale production of GVL, catalytic systems that maximize yield without the use of precious metals, high H2 pressure and excessive number of unit operations are highly required. Catalytic transfer hydrogenation offers an alternative to molecular hydrogen. In this sense, Bui et al. (2013) have reported an integrated catalytic process for the efficient production of GVL from furfural in a one-pot process using a combination of Lewis and Brönsted catalysts. Furfural was firstly converted into furfuryl alcohol through a transfer hydrogenation reaction promoted by a Lewis acid catalyst (Zr-Beta zeolite) and using 2-butanol as the hydrogen donor. Next, a Brönsted acid catalyst (Al-MFI zeolite) converted furfuryl alcohol into a mixture of levulinic acid and butyl levulinate through hydrolytic ring-opening reactions. Finally, both levulinic acid and butyl levulinate underwent a second transfer-hydrogenation step to produce GVL with a yield close to 80% (see Fig. 13.3). Note that a multifunctional catalyst which enables promotion of both type of reactions is a challenge for the future.
image
Figure 13.3 Cascade reaction for the production of GVL from xylose by the combination of Lewis and Brönsted acid catalyst.

13.5. Blending effect of oxygenated biofuels with conventional fuels

Oxygenated fuels from lignocellulosic biomass can be used as blend components in gasoline, diesel, and biodiesel. In this section, we will discuss the works described in the literature dealing with the blending effect of these advanced oxygenated biofuels with conventional fuels (see Table 13.1).
Christensen et al. (2011a) blended some oxygenate compounds derived from lignocellulosic biomass (MF, DMF, MTHF, ML, EL, BL, and MV) with three different gasoline blend-stocks at levels up to 3.7 wt.% oxygen. Chemical and physical properties of the blends were compared to the requirements of ASTM specification D4814 for spark-ignited engine fuels to determine their utility as gasoline extenders. With the exception of MF, all other oxygenates reduced vapor pressure. This may indicate an economic and environmental benefit by eliminating the need to remove light-end components to meet maximum vapor pressure limits set to reduce evaporative emissions. The density and viscosity of blends with some of the oxygenates increased, although the impact that these changes will have on vehicle fuel delivery systems is unclear. BL was found to raise distillation temperatures (distillation end point exceeded 225°C, thus failing the specification) which may cause excessive combustion chamber deposits and lube oil dilution. Distillation parameters for the other oxygenated compounds were within the specification limits. All oxygenates tested except MTHF increased octane rating. However, oxygenates other than MF and DMF did not have a sufficient blending octane number to raise the antiknock index (AKI) (AKI = [RON + MON]/2) above the 87 minimum requirement. ML is fully miscible with water and can separate from gasoline under cold temperatures. Concluding, MF and DMF appear to have good potential because of their favorable properties, MV and EL may also have potential as gasoline blend-stocks, while MTHF appears to have less potential because of its low octane number and high water solubility.
DMF is considered superior to ethanol in several important ways: it has an energy content of 31.5 MJ/L, similar to that of regular gasoline (31.7 MJ/L) and 40% greater than that of ethanol (23 MJ/L); DMF (bp 92–94°C) is less volatile than ethanol (bp 78)C); it blends more easily with petroleum; and, contrary to ethanol, it is immiscible with water, so it does not absorb water from the atmosphere. Rothamer and Jennings (2012) blended DMF with gasoline at volume concentrations of 5, 10, and 15%. The knocking propensity of these mixtures was compared to the performance of E10 and gasoline. The results indicated that for direct-injection operation, ethanol is potentially more effective at reducing engine knock than DMF at the same blend percentage. However, due to the attractive energy density and much lower water solubility of DMF, it is a potentially competitive blending additive. Moreover, analysis of combustion emissions showed that DMF mixtures gave the lowest total carbonyl emissions and, more significantly, the lowest emissions of the more harmful formaldehyde and acetaldehyde among the four oxygenated fuels (n-butanol, ethanol, methanol, and DMF) and gasoline (Daniel el at., 2012).

Table 13.1

Oxygenated biofuels from biomass reported in literature blended with conventional fuels

BiofuelChemical structureOxygen (wt%)Blended with
GLNDSLBDSL
2-Methylfuran (MF)icon19.5X
2-Methyl-tetrahydrofuran (MTHF)icon18.6X
Methyl levulinate (ML)icon36.9X
Ethyl levulinate (EL)icon33.3XXX
Butyl levulinate (BL)icon27.9XX
γ-Valerolactone (GVL)icon32.0X
2,5-Dimethylfuran (DMF)icon16.6X
Methyl valerate (MV)icon27.5X
Ethyl valerate (EV)icon24.6X
Ethyl tetrahydrofurfuryl ether (ETE)icon24.6X
Furfuryl ethyl ether (FEE)icon25.4X
5-Methoxy-methyl furfural (MMF)icon34.3X
5-Ethoxymethyl furfural (EMF)icon31.1X
5-Buthoxymethyl furfural (BMF)icon26.4X

image
image

GLN, gasoline; DSL, diesel; BDSL, biodiesel.

Horvath et al. (2008) compared the blended properties of 10 v/v % mixtures of GVL or ethanol with 95 octane gasoline. All the data for GVL were similar to those obtained with ethanol, but its lower vapor pressure leads to improved performance. Derived GVL oxygenates such as “valeric esters” have also shown even better properties than GVL as gasoline extenders (Lange et al., 2010). The EV blends (10% and 20% v/v in gasoline) showed a favorable increase in octane number (RON and MON) without deterioration of properties such as corrosion and gum formation. EV blending increased the gasoline density and reduced its volatility and lowered its content of aromatics, olefins, and sulfur. Moreover, the presence of EV in gasoline showed no measurable impact on engine wear, oil degradation, vehicle durability, engine deposits, or regulated tailpipe emissions (EURO 4 and 5 specifications). The mixtures were stable over the 4-month period of the test and had no negative impact on the fuel storage and dispensing equipment (tanks, pipes, pumps, and filters).
Levulinate esters of ethanol (EL) and higher-molecular-weight alcohols (BL) have been shown as potential diesel-blending components (Christensen et al., 2011b). Both esters improved the lubricity and conductivity of the diesel fuel. Nevertheless, the low cetane number of both esters and poor solubility in diesel fuel at low temperatures limits partially their commercial application as diesel blend components. EL has also been explored as a blend component for biodiesel (Joshi et al., 2011). The mixtures of EL with biodiesel (2.5, 5, 10 and 20% v/v in biodiesel) showed better cold properties with a gradual decrease of cloud, pour and filter plugging points upon addition of EL. Recently, BL has been examined for blending with jet aviation kerosene (Chuck and Donnelly, 2014). Although the miscibility of BL in kerosene is good in comparison with EL, this oxygenated fuel showed the worst performance of the fuels under investigation.
The performance of furan derivatives (FEE and ETE) in diesel engines has also been assessed (de Jong et al., 2012). Smoke and particulates, as well as sulfur content, decreased significantly with increasing ETE blending concentrations. Fuel consumption increases with increasing ETE amount, but is completely in line with the calculated lower energy content of ETE. The CO, CO2, NO2 exhaust percentages, and THC content appeared to be independent of ETE concentrations. NOX only shows a slight increase at higher blending percentages (>10%). Hydrogenated furanics (ETE) gave slightly better engine performance than nonhydrogenated ones (FEE).
5-HMF ethers such as MMF, EMF, and BMF are also interesting blending compounds for fuels (Gruter and de Jong, 2009). EMF is the main representative of the 5-alkoxymethyl furfural ethers family and it is considered to be an excellent additive for diesel. It has a high energy density of 31.3 MJ/L, which is similar to regular gasoline (31.7 MJ/L), nearly as good as diesel (34.9 MJ/L) and significantly higher than ethanol (23 MJ/L). With favorable blending properties, EMF has been used mixed with commercial diesel in engine tests, leading to promising results with a significant reduction in soot (fine particulates), and a reduction in the SOx emissions. Although MMF and EMF are useful as fuel additives, these ethers show at high concentrations phase separation problems. In contrast, di-ethers coming from the hydrogenation in alcohol medium of EMF and MMF are miscible with commercial diesel in all blend ratios. Moreover, the ring hydrogenated products of these di-ethers have been shown to be good candidates for aviation fuel formulation (Gruter and de Jong, 2009).

13.6. Catalytic conversion of γ-valerolactone to liquid hydrocarbon fuels

γ-Valerolactone (GVL) is a starting point (raw material) in numerous transformations. One of these processes is the production of biomass-derived hydrocarbons, with the same properties as regular fuels obtained from conventional feedstock/procedures (Fig. 13.4).
The transformation of GVL into hydrocarbons can be achieved in multiple ways, for instance, by its transformation into valeric acid (pentanoic acid) through the hydrogenolysis of the lactone cycle, as previously discussed in Section 13.3.2 (Pham et al., 2011; Du et al., 2012). From this point, once the valeric acid is produced, the most obvious and direct way to obtain hydrocarbons is the direct hydrogenation of valeric acid. However, this alternative would involve high hydrogen consumption and the final product (pentene) would not fulfill the requirements of conventional transportation fuels, such as the boiling point. As an alternative to this option, the construction of larger carbon chains, followed by a hydrogenation step, is preferred. In this way, the final products show carbon chains in the range of those shown by regular fuels, leading to similar physicochemical properties in the so-obtained hydrocarbons. Carbon chain enlargement has been reported to be easily obtained from valeric acid through a ketonization route, yielding 5-nonanone as a final product, both in presence of CeO2/ZrO2 (Serrano-Ruiz et al., 2010a; Martin-Alonso et al., 2010; Zaytseva et al., 2013) and Pd/Al2O3 catalysts (Serrano-Ruiz et al., 2010b; Pham et al., 2011), although in the latter case, larger contact times are required. The resultant 5-nonanone can be easily separated from water because of immiscibility, which leads to an important energy saving as compared to other alternatives. 5-Nonanone can be later hydrogenated to the corresponding alcohol and submitted to hydrogenation/dehydration to provide C9 hydrocarbons, which can finally be isomerizated to achieve the required properties for the desired fuel.
image
Figure 13.4 Conversion of GVL into hydrocarbons.
A different alternative for the transformation of GVL into hydrocarbons is the direct decarboxylation of this platform molecule (Bond et al., 2010a,b) to produce a mixture of butenes (mostly 1-butene), which can be fed as starting raw-material to conventional alkylation units, such as the UOP Butamer process, to produce large hydrocarbon fuels. This option is quite interesting, because of the low requirements of the decarboxylation step, both in terms of reactants—the use of hydrogen, unlike hydrogenolysis, is not required—as well as in term of technology—decarboxylation of GVL can be achieved in presence of mild-acid catalysts such as SiO2-Al2O3 gels operating above 250°C (Bond et al., 2011). In addition, the final products, the mixture of butenes can be processed in already-present alkylation units in standard refinery units. In this way, the production of regular hydrocarbon fuels can be easily achieved, by combination of the use of biomass-derived feedstock and conventional refinery units. This procedure has been calculated to provide profits for selling prices for the final butene oligomers in the range 4.40–4.92 $/gallon (Sen et al., 2012a,b).
Although the transformation of GVL into hydrocarbons has been mainly reported through the reaction pathways involving valeric acid and butenes formation, there is still another one, recently reported, that, due to its simplicity, shows enormous potential to be carried out at an industrial scale. This is the catalytic pyrolysis of GVL to yield aromatic hydrocarbons (Zhao et al., 2012). In this case, several heterogeneous acid catalysts were tested, including different zeolites and mesostructured materials. Catalytic assays revealed a very high catalytic activity and selectivity toward aromatic hydrocarbon in the case of the HZSM-5 zeolite (Si/Al = 25), which provided more than 55% of carbon yield, being fully recyclable for several consecutive catalytic assays. This work opens a new possibility for the inclusion of biomass-derived feedstock in conventional oil refinery units, a highly desirable alternative in the substitution of fossil fuels by renewable energy sources such as lignocellulosic biomass.

13.7. Furan derivatives as platform molecules for liquid hydrocarbon fuels

Furan platform molecules (5-HMF and furfural) can be efficient converted to liquid alkanes with a high number of carbons, which can be used as gasoline, diesel, and jet fuels, by means of CC coupling reactions whereas oxygen is removed by dehydration, hydrogenation, and hydrogenolysis reactions (see Fig. 13.5). These kinds of processes will be discussed in this section.
image
Figure 13.5 Pathways to convert 5-HMF into alkanes. Reprinted with permission of Walter de Gruyter (Ed.), Biorefinery: From Biomass to Chemicals and Fuels.

13.7.1. 5-HMF upgrading via CC coupling reactions

In order to obtain diesel fuels of high quality, West et al. (2008) proposed a process involving the aldol condensation of 5-HMF with acetone in a biphasic reactor system catalyzed by aqueous NaOH, followed by hydrogenation/dehydration/ring opening in the presence of a bifunctional catalyst such as Pd/Al2O3 and Pt/NbPO5 producing a mixture of linear C9 and C15 alkanes with a yield of 73%. Similar protocol was followed by Chatterjee et al. (2010) but using Pd/Al-MCM-41 catalyst for the second step in supercritical carbon dioxide, achieving a 99% selectivity of C9 linear alkanes. In an attempt to coupling aqueous phase aldol-condensation of 5-HMF with acetone and hydrogenation/dehydration reactions, a bifunctional base-metal catalyst based on Pd supported over different mixed oxides (MgO, ZrO2, CaO, and Al2O3) has been reported. For instance, using Pd/MgO-ZrO2 in that process produce C12 alkanes from 5-HMF (Faba et al., 2011).
Recently, Liu and Chen (2014) have developed an integrated catalytic process for the conversion of 5-HMF into alkane fuels. The integrated catalytic process involves three different steps: (1) 5-HMF production from fructose and glucose; (2) self-coupling of 5-HMF catalyzed by n-heterocyclic carbine (NHC) to yield furoin intermediates; and (3) linear alkanes production by hydrodeoxygenation using metal-acid tandem as catalysts system consisting of Pd/C + La(OTf)3 + acetic acid. Alkanes were produced in 78% yield with a 64% selectivity to n-C12H26.

13.7.2. Furfural upgrading via CC coupling reactions

Similar to 5-HMF, furfural can also undergo aldol-condensation with external carbonyl-containing molecules using base or acid catalyst. Further hydrogenation of aldol products can produce high-quality longer-chain alkanes.
Opposite to 5-HMF, high yields of single and double condensation products are achieved in the aldol-condensation of furfural with acetone in the presence of an aqueous phase with NaOH catalyst. Mixed oxides with different basic strength have also been used for this reaction, getting more activity with those catalysts with higher concentration of strength basic sites, ie, Mg-Zr > Mg-Al > Ca-Zr (Faba et al., 2012). Moreover, the basic site distribution can be improved supporting the Mg-Zr mixed oxide on mesoporous carbons which leads to a higher interaction of the reactants with the carbon surface achieving 96% conversion of furfural with 88% selectivity for C13 and C8 adducts (Faba et al., 2013).
More interesting is the sequential strategy developed by the Dumesic group. Similar to 5-HMF, the production of C10 alkanes is carried out by a cascade reaction aldol-condensation of furfural with acetone followed by hydrogenation/dehydration using a bifunctional catalyst Pd/MgO-ZrO2 with high overall carbon yield (>80%) (Barret et al., 2006). Another sequential process of two consecutive steps to obtain a mixture of long-chain alkanes with excellent properties as a diesel fuel (cetane number and flow properties at low temperature) was recently reported by Corma et al. (2012) named the “Sylvan process.” The first step consists of a hydroxyalkylation/alkylation of three molecules of 2-methylfuran (sylvan) or hydroxylation of sylvan with aldehydes or ketones catalyzed by organic and inorganic acids to yield oxygenated intermediate molecules (butanal is considered to be the most promising molecular linker). In the second step, a complete hydrodeoxygenation of the previous products catalyzed by platinum metal supported on nonacidic materials leads to the desired mixture of alkanes within the diesel range namely 6-alkylundecane.
More recently, different solid acid catalysts were studied for the alkylation of MF with mesityl oxide (Li et al., 2014b). Among the investigated candidates, Nafion-212 resin exhibited the highest catalytic efficiency, which can be explained by its higher acid strength. For the second step of hydrodeoxygenation, Ni–Mo2C/SiO2 exhibited an evident advantage at the cost and the selectivity to diesel range alkanes (77% yield).

13.8. Sugars to hydrocarbon fuels: aqueous phase reforming process

Hydrodeoxygenation reactions are an effective alternative for the removal of the oxygen atoms of selected biomass-derived compounds in order to obtain biofuels (Furimsky, 2013; Chaudhari et al., 2013; Nakagawa et al., 2015). However, this alternative is quite expensive for several reasons: the consumption of hydrogen and the usually harsh reaction conditions, in terms of hydrogen pressure and temperature conditions, needed to drive the desired chemical transformations. The development of highly efficient, selective heterogeneous catalysts enables the promotion of these transformations to partially overcome these latter drawbacks (De et al., 2015). In contrast, the consumption of hydrogen is still a concern, because of the high cost associated with the production and purification of this chemical.
Hydrogen is conventionally obtained from a process starting from fossil fuels—typically by methane steam reforming. However, during the last decade great efforts have been applied to develop alternative hydrogen production techniques to the conventional ones or in the adjustment of the already-existing methods to the use of alternative feedstock. An interesting alternative in this last sense is the use of biomass as feedstock for hydrogen production (Kalinci et al., 2009; Balat and Kirtay, 2010; Tanksale et al., 2010; Uddin and Daud, 2014), closing the cycle in the use of renewable materials for the production of both the structural carbon chains of the final pursued chemicals and the required hydrogen used for the removal of oxygen.
The production of both hydrogen and alkanes can be achieved using the same procedure in a single step, aqueous phase reforming (APR) of biomass-derived oxygenated hydrocarbons coming from renewable biomass sources (Davda et al., 2005). APR takes advantage of the ability of several noble metal-based catalysts with hydrogenating activity, including Pd, Pt, Ru, Rh, or Ir, and their mixtures (Huber and Dumesic, 2006; Shabaker and Dumesic, 2004; Tanksale et al., 2007), to favor the water gas shift reaction (WGS), starting from oxygenated hydrocarbons, under aqueous phase conditions to yield H2 + CO or H2 + CO2 gas mixtures, depending on the substrate and the reaction conditions.
The reactions taking place in APR processes include hydrolysis, dehydration, reforming, aldol-condensation, and hydrogenolysis transformations (Benson et al., 2013), starting from polysaccharides and involving a whole collection of oxygenated reaction intermediates. These intermediates react in contact with the surface of the aforementioned catalysts—based on metals with hydrogenating/dehydrogenating capability, yielding CO. Subsequent transformation of CO into CO2 through the WGS reaction leads to the formation of hydrogen, as the main hydrogen production pathway. Depending on the starting oxygenated hydrocarbon, the reaction mechanism is more or less complicated, but, in any case, these can be summarized as CC and CO bonds cleavage reactions, dehydration, hydrogenation, and dehydrogenation reactions. In this way, the transformation of oxygenated hydrocarbons by APR allows obtaining of multiple possible products, ranging from hydrogen to alkanes (Fig. 13.6). Depending on the objective of the APR transformation, this can be used either for the production of hydrogen (first stage) or looking for maximizing the production of larger alkanes (if reaction is allowed to proceed further). This can be achieved by tuning the catalytic activity of the hydrogenation/dehydrogenation catalyst, promoting or depressing the different reaction pathways, thus favoring hydrogen or alkane production.
image
Figure 13.6 APR reaction mechanism. Reprinted with permission of the Royal Society of Chemistry. Energy Environmental Science 2012, 5, 7393–7420.
This knowledge, the ability to tune the catalytic selectivity of the heterogeneous catalysts which play the crucial role in the transformation of biomass-derived water-soluble oxygenated chemicals into hydrogen/alkanes, is the basis of the BioForming Process (Virent Energy Systems) (Dumesic and Roman–Leshkov, 2009, Fig. 13.7). The process consists of two different reaction stages in which, starting from sugars or lignocellulosic biomass hydrolysates, hydrogen is produced in the first step. Together with H2, several low-molecular-weight oxygenated compounds, including alcohols, acids, ketones, and aldehydes are also produced. These are the basis for the production of larger alkanes since, in the second reaction step, the oxygenated chemicals can be transformed, through multiple reaction pathways (condensation, hydrodeoxygenation, dehydration, oligomerization, etc.) into regular fuels, including diesel, gasoline, or kerosene, which can economically compete with petroleum fuels at crude oils prices greater than 60 $ per barrel (Blommel et al., 2008).
image
Figure 13.7 Virent's bio-forming process. Reprinted with permission of the Royal Society of Chemistry. Energy Environmental Science 2012, 5, 7393–7420.

13.9. Final remarks and future outlook

Commercial processes for the conversion of biomass to biofuels are now based mainly on the production of bioethanol (from sugar cane and corn, and recently some commercial plants processing lignocellulose feedstock have been set up) and biodiesel by processing of triglycerides molecules. In this chapter, the conversion of lignocellulose toward liquid biofuels has been demonstrated through the formation of several platform molecules (5-HMF, levulinic acid, and furfural). But, unfortunately, these kind of processes are still far from being commercial large scale operations.
We have seen that in most cases the catalytic processes involve a large number of reaction steps. An integrated development of catalytic cascade processes and adapted separation steps will be necessary for the future. Likewise, the design of multifunctional catalysts than can perform cascade-type reactions in less reaction steps and avoiding intermediate product separation and purification will facilitate implementation of sustainable lignocellulose-based production processes. Likewise, heterogeneous catalysts must have an outstanding role to substitute homogeneous mineral acids and bases. On the other hand, a great number of the approaches reported in this chapter need a high amount of hydrogen in order to remove the oxygen and yield high-energy-density biofuels which will have a great impact on the final cost. Hopefully, the transformation of carbohydrates to hydrogen using APR processes might be a good alternative to the current fossil-based hydrogen sources and supplying renewable hydrogen. Hence, much catalysis and engineering research are still needed to achieve the potential of these platform molecules for biofuel production.
Other aspects to have in mind are the improvement of feedstock sustainability and availability, and acceleration of the market deployment of the most promising advanced biofuels. Of course, commercial deployment of such fuels will require significant effort in the areas of registration, specification, and legislation.

Acknowledgments

Financial support from the Spanish Ministry of Economy and Competitiveness through the project CTQ- 2014–52907-R and Regional Government of Madrid through the project S2013/MAE-2882 RESTOENE-2 are kindly acknowledged.

References

Assary R.S, Curtiss L.A, Dumesic J.A. Exploring Meerwein – Ponndorf – Verley reduction chemistry for biomass catalysis using a first-principles approach. ACS Catalysis. 2013;3:2694–2704.

Balakrishnan M, Sacia E.R, Bell A.T. Etherification and reductive etherification of 5-(hydroxymethyl)furfural: 5-(alkoxymethyl)furfurals and 2,5-bis(alkoxymethyl) furans as potential bio-diesel candidates. Green Chemistry. 2012;14:1626–1634.

Balat H, Kirtay E. Hydrogen from biomass – present scenario and future prospects. International Journal of Hydrogen Energy. 2010;35:7416–7426.

Barrett C.J, Chheda J.N, Huber G.W, Dumesic J.A. Single-reactor process for sequential aldol-condensation and hydrogenation of biomass-derived compounds in water. Applied Catalysis B: Environmental. 2006;66:111–118.

Benson T.J, Daggolu P.R, Hernandez R.A, Liu S, White M.G. Catalytic deoxygenation chemistry: upgrading of liquid derived from biomass processing. Advances in Catalysis. 2013;56:187–353.

Bicker M, Kaiser D, Ott L, Vogel H. Dehydration of D-fructose to hydroxymethylfurfural in sub- and supercritical fluids. Journal of Supercritical Fluids. 2005;36:118–126.

Blommel P.G, Keenan G.R, Rozmiarek P.T, Cortright R.D. Catalytic conversion of sugar into conventional gasoline, diesel, jet fuel, and other hydrocarbons. International Sugar Journal. 2008;110(1319):672–679.

Bohre A, Dutta S, Saha B, Mahdi M.A. Upgrading furfural to drop-in biofuels: an overview. ACS Sustainable Chemistry & Engineering. 2015 doi: 10.1021/acssuschemeng.5b00271.

Bond J.Q, Martín-Alonso D, Wang D, West R.M, Dumesic J.A. Integrated catalytic conversion of γ-valerolactone to liquid alkanes for transportation fuels. Science. 2010;327:1110–1114.

Bond J.Q, Martín-Alonso D, West R.M, Dumesic J.A. γ-valerolactone ring-opening and decarboxylation over SiO2/Al2O3 in the presence of water. Langmuir. 2010;26(21):16291–16298.

Bond J.Q, Wang D, Martín-Alonso D, Dumesic J.A. Interconversion between γ-valerolactone and pentenoic acid combined with decarboxylation to form butene over silica/alumina. Journal of Catalysis. 2011;281:290–299.

Brown D.W, Floyd A.J, Kinsman R.G, Roshan-Ali Y. Dehydration reactions of fructose in nonaqueous media. Journal of Chemical Biotechnology. 1982;32:920–924.

Bui L, Luo H, Gunther W.R, Roman-Leshkov Y. Domino reaction catalyzed by zeolites with Brönsted and Lewis acid sites for the production of γ-valerolactone from furfural. Angewandte Chemie International Edition. 2013;52:8022–8025.

Chalid M, Broekhuis A.A, Heeres H.J. Experimental and kinetic modeling studies on the biphasic hydrogenation of levulinic acid to γ-valerolactone using a homogeneous water-soluble Ru–(TPPTS) catalyst. Journal of Molecular Catalysis A: Chemical. 2011;341(1):14–21.

Chan-Thaw C.E, Marelli M, Psaro R, Ravasio N, Zaccheria F. New generation biofuels: γ-valerolactone into valeric esters in one pot. RSC Advances. 2013;3:1302–1306.

Chatterjee C, Pong F, Sen A. Chemical conversion pathways for carbohydrates. Green Chemistry. 2015;17:40–71.

Chatterjee M, Matsushima K, Ikushima Y, Sato M, Yokoyama T, Kawanami H, Suzuki T. Production of linear via hydrogenative ring opening of a furfural-derived compound in supercritical. Green Chemistry. 2010;12:779–782.

Chaudhari R.V, Torres A, Jin X, Subramaniam B. Multiphase catalytic hydrogenolysis/hydrodeoxygenation processes for chemicals from renewable feedstocks: kinetics, mechanism, and reaction engineering. Industrial & Engineering Chemistry Research. 2013;52:15226–15243.

Chia M, Dumesic J.A. Liquid-phase catalytic transfer hydrogenation and cyclization of levulinic acid and its esters to γ-valerolactone over metal oxide catalysts. Chemical Communications. 2011;47:12233–12235.

Chidambaram M, Bell A.T. A two-step approach for the catalytic conversion of glucose to 2,5-dimethylfuran in ionic liquids. Green Chemistry. 2010;12:1253–1262.

Christensen E, Yanowitz J, Ratcliff M, McCormick R.L. Renewable oxygenate blending effects on gasoline properties. Energy Fuels. 2011;25(10):4723–4733.

Christensen E, Williams A, Paul S, Burton S, McCormick R.L. Properties and performance of levulinate esters as diesel blend components. Energy Fuels. 2011;25(11):5422–5428.

Chuck C.J, Donnelly J. The compatibility of potential bio-derived fuels with Jet A-1 aviation kerosene. Applied Energy. 2014;118:83–91.

Cirujano F.G, Corma A, Llabrés i Xamena F.X. Conversion of levulinic acid into chemicals: synthesis of biomass derived levulinate esters over Zr-containing MOFs. Chemical Engineering Science. 2015;124:52–60.

Climent M.J, Corma A, Iborra S. Conversion of biomass platform molecules into fuel additives and liquid hydrocarbon fuels. Green Chemistry. 2014;16:516–547.

Corma A, de la Torre O, Renz M. Production of high-quality diesel from cellulose and hemicellulose by the Sylvan process: catalysts and process variables. Energy & Environmental Science. 2012;5:6328–6344.

Daniel R, Wei L, Xu H, Wang C, Wyszynski M.L, Shuai S. Speciation of hydrocarbon and carbonyl emissions of 2,5-dimethylfuran combustion in a DISI engine. Energy Fuels. 2012;26(11):6661–6668.

Davda R.R, Shabaker J.W, Huber G.W, Cortright R.D, Dumesic J.A. A review of catalytic issues and process conditions for renewable hydrogen and alkanes by aqueous-phase reforming of oxygenated hydrocarbons over supported metal catalysts. Applied Catalysis B: Environmental. 2005;56:171–186.

De S, Saha B, Luque R. Hydrodeoxygenation processes: advances on catalytic transformations of biomass-derived platform chemical into hydrocarbon fuels. Bioresource Technology. 2015;178:108–118.

de Jong E, Vijlbrief T, Hijkoop R, Gruter G.-J.M, van der Waal J.C. Promising results with YXY diesel components in an ESC test cycle using PACCAR diesel engine. Biomass Bioenergy. 2012;36:151–159.

Delidovich I, Leonhard K, Palkovits R. Cellulose and hemicellulose valorization: an integrated challenge of catalysis and reaction engineering. Energy & Environmental Science. 2014;7:2803–2830.

Ding D, Wang J, Xi J, Liu X, Lu G, Wang Y. High-yield production of levulinic acid from cellulose and its upgrading to γ-valerolactone. Green Chemistry. 2014;16:3846–3853.

Dong F, Zhu Y, Ding G, Cui J, Li X, Li Y. One-step conversion of furfural into 2-methyl-tetrahydrofuran under mild conditions. ChemSusChem. 2015;8:1534–1537.

Du X.L, Bi Q.Y, Liu Y.M, Cao Y, He H.Y, Fan K.N. Tunable copper-catalyzed chemoselective hydrogenolysis of biomass-derived gamma-valerolactone into 1,4-pentanediol or 2-methyltetrahydrofuran. Green Chemistry. 2012;14(4):935–939.

Dumesic J.A, Roman–Leshkov Y. Production of Liquid Alkanes in the Jet Fuel Range (C8C15) from Biomass–derived Carbohydrates. 2009 US 2009/0124839 A1.

Dutta S, De S, Saha B, Alam Md I. Advances in conversion of hemicellulosic biomass to furfural and upgrading to biofuels. Catalysis Science & Technology. 2012;2:2015–2036.

Faba L, Diaz E, Ordonez S. Performance of bifunctional Pd/MxNyO (M = Mg, Ca; N = Zr, Al) catalysts for aldolization-hydrogenation of furfural-acetone mixtures. Catalysis Today. 2011;164:451–456.

Faba L, Diaz E, Ordonez S. Aqueous-phase furfural-acetone aldol condensation over basic mixed oxides. Applied Catalysis B: Environmental. 2012;2012(113):201–211.

Faba L, Diaz E, Ordonez S. Improvement on the catalytic performance of MgZr mixed oxides for furfural acetone aldol condensation by supporting on mesoporous carbons. ChemSusChem. 2013;6:463–473.

Fernandes D.R, Rocha A.S, Mai E.F, Mota C.J.A, da Silva V.T. Levulinic acid esterification with ethanol to ethyl levulinate production over solid acid catalysts. Applied Catalysis A: General. 2012;425–426:199–204.

Furimsky E. Hydroprocessing in aqueous phase. Industrial & Engineering Chemistry Research. 2013;52:17695–17713.

Geboers J, Wang X, de Carvalho A.B, Rinaldi R. Densification of biorefinery schemes by H-transfer with Raney Ni and 2-propanol: a case study of a potential avenue for valorization of alkyl levulinates to alkyl γ-hydroxypentanoates and γ-valerolactone. Journal of Molecular Catalysis A: Chemical. 2014;388–389:106–115.

Gruter G.-J, de Jong E. Furanics: novel fuel options from carbohydrates. Biofuels Technology. 2009;1:11–17.

Gurbuz E.I, Alonso D.M, Bond J.Q, Dumesic J.A. Reactive extraction of levulinate esters and conversion to γ-valerolactone for production of liquid fuels. ChemSusChem. 2011;4:357–361.

Heeres H, Handana R, Chunai D, Rasrendra C.B, Girisuta B, Heeres H.J. Combined dehydration/(transfer)-hydrogenation of C6-sugars (D-glucose and D-fructose) to γ-valerolactone using ruthenium catalysts. Green Chemistry. 2009;11:1247–1255.

Hengne A.M, Rode C.V. Cu-ZrO2 nanocomposite catalyst for selective hydrogenation of levulinic acid and its ester to gamma-valerolactone. Green Chemistry. 2012;14(4):1064–1072.

Horváth I.T, Mehdi H, Fábos V, Boda L, Mika L.T. γ-Valerolactone-a sustainable liquid for energy and carbon-based chemicals. Green Chemistry. 2008;10(2):238–242.

Huang Y.-B, Chen M.-Y, Yan L, Guo Q.X, Fu Y. Nickel–Tungsten carbide catalysts for the production of 2,5-dimethylfuran from biomass-derived molecules. ChemSusChem. 2014;7:1068–1072.

Huber G.W, Dumesic J.A. An overview of aqueous-phase catalytic processes for production of hydrogen and alkanes in a biorefinery. Catalysis Today. 2006;111:119–132.

Joshi H, Moser B.R, Toler J, Smith W.F, Walker T. Ethyl levulinate: a potential bio-based diluent for biodiesel which improves cold flow properties. Biomass Bioenergy. 2011;35(7):3262–3266.

Kalinci Y, Hepbasli A, Dincer I. Biomass-based hydrogen production: a review and analysis. International Journal of Hydrogen Energy. 2009;34:8799–8817.

Kraus G.A, Guney T. A direct synthesis of 5-alkoxymethylfurfural ethers from fructose via sulfonic acid-functionalized ionic liquids. Green Chemistry. 2012;14:1593–1596.

Krystof M, Perez-Sanchez M, de Maria P.D. Lipase-catalyzed (trans)esterification of 5-Hydroxy-methylfurfural and separation from 5-HMF esters using deep-eutectic solvents. ChemSusChem. 2013;6:630–634.

Kuwahara Y, Kaburagi W, Nemoto K, Fujitani T. Esterification of levulinic acid with ethanol over sulfated Si-doped ZrO2 solid acid catalyst: study of the structure–activity relationships. Catalysis Today. 2014;237:18–28.

Lange J.-P, van de Graaf W.D, Haan R.J. Conversion of furfuryl alcohol into ethyl levulinate using solid acid catalysts. ChemSusChem. 2009;2:437–441.

Lange J.-P, Price R, Ayoub P.M, Louis J, Petrus L, Clarke L, Gosselink H. Valeric biofuels: a platform of cellulosic transportation fuels. Angewandte Chemie International Edition. 2010;49(26):4479–4483.

Lange J.P, van de Heide E, van Buijtenen J, Price R. Furfural – a promising platform for lignocellulosic biofuels. ChemSusChem. 2012;5:150–166.

Lanzafame P, Temi D, Perathoner S, Centi G, Macario A, Aloise A, Giordano G. Etherification of 5-hydroxymethyl-2-furfural (HMF) with ethanol to biodiesel components using mesoporous solid acidic catalysts. Catalysis Today. 2011;175(1):435–441.

Lew C.M, Rajabbeigi N, Tsapatsis M. One-pot synthesis of 5-(Ethoxymethyl) furfural from glucose using Sn-BEA and amberlyst catalysts. Industrial Engineering Chemistry Research. 2012;51:5364–5366.

Li H, Govind K.S, Kotni R, Shunmugavel S, Riisager A, Yang S. Direct catalytic transformation of carbohydrates into 5-ethoxymethylfurfural with acid–base bifunctional hybrid nanospheres. Energy Conversion Management. 2014;88:1245–1251.

Li S, Li N, Li G, Wang A, Cong Y, Wang X, Zhang T. Synthesis of diesel range alkanes with 2-methylfuran and mesityloxide from lignocelluloses. Catalysis Today. 2014;234:91–99.

Li W, Xie J.H, Lin H, Zhou Q.L. Highly efficient hydrogenation of biomass-derived levulinic acid to γ-valerolactone catalyzed by iridium pincer complexes. Green Chemistry. 2012;14:2388–2390.

Liu B, Zhang Z, Deng K. Efficient one-pot synthesis of 5-(ethoxymethyl) furfural from fructose catalyzed by a novel solid catalyst. Industrial Engineering Chemistry Research. 2012;51:15331–15336.

Liu D, Chen E.Y.-X. Integrated catalytic process for biomass conversion and, upgrading to C12 furoin and alkane fuel. ACS Catalysis. 2014;4:1302–1310.

Luo H.Y, Consoli D.F, Gunther W.R, Román-Leshkov Y. Investigation of the reaction kinetics of isolated Lewis acid sites in Beta zeolites for the Meerwein – Ponndorf – Verley reduction of methyl levulinate to γ-valerolactone. Journal of Catalysis. 2014;320:198–207.

Martin-Alonso D, Bond J.Q, Serrano-Ruiz J.C, Dumesic J.A. Production of liquid hydrocarbon transportation fuels by oligomerization of biomass-derived C9 alkenes. Green Chemistry. 2010;12:992–999.

Melero J.A, Morales G, Iglesias J, Paniagua M, Hernandez B, Penedo S. Efficient conversion of levulinic acid into alkyl levulinates catalyzed by sulfonic mesostructured silicas. Applied Catalysis A: General. 2013;466:116–122.

Morales G, Osatiashtiani A, Hernández B, Iglesias J, Melero J.A, Paniagua M, Brown R, Granollers M, Lee A.F, Wilson K. Conformal sulfated zirconia monolayer catalysts for the one-pot synthesis of ethyl levulinate from glucose. Chemical Communications. 2014;50:11742–11745.

Mukherjee A, Dumont M.J, Raghavan V. Review: sustainable production of hydroxymethylfurfural and levulinic acid: challenges and opportunities. Biomass and Bioenergy. 2015;72:143–183.

Nakagawa Y, Liu S.B, Tamura M, Tomishige K. Catalytic total hydrodeoxygenation of biomass-derived polyfunctionalized substrates to alkanes. ChemSusChem. 2015;8(7):1114–1132.

Nakagawa Y, Takada K, Tamura M, Tomishige K. Total hydrogenation of furfural and 5-hydroxymethylfurfural over supported Pd–Ir Alloy catalyst. ACS Catalysis. 2014;4:2718–2726.

Nandiwale K.Y, Sonar S.K, Niphadkar P.S, Joshi P.N, Deshpande S.S, Patil V.S, Bokade V.V. Catalytic upgrading of renewable levulinic acid to ethyl levulinate biodiesel using dodecatungstophosphoric acid supported on desilicated H-ZSM-5 as catalyst. Applied Catalysis A: General. 2013;460–461:90–98.

Nishimura S, Ikeda N, Ebitani K. Selective hydrogenation of biomass derived 5-hydroxymethylfurfural (HMF) to 2,5-dimethylfuran (DMF) under atmospheric hydrogen pressure over carbon supported PdAu bimetallic catalyst. Catalysis Today. 2014;232:89–98.

Oliveira B.L, da Silva V.T. Sulfonated carbon nanotubes as catalysts for the conversion of levulinic acid into ethyl levulinate. Catalysis Today. 2014;234:257–263.

Pasquale G, Vázquez P, Romanelli G, Baronetti G. Catalytic upgrading of levulinic acid to ethyl levulinate using reusable silica-included Wells-Dawson heteropolyacid as catalyst. Catalysis Communications. 2012;18:115–120.

Patil C.R, Niphadkar P.S, Bokade V.V, Joshi P.N. Esterification of levulinic acid to ethyl levulinate over bimodal micro–mesoporous H/BEA zeolite derivatives. Catalysis Communications. 2014;43:188–191.

Peng L, Lin L, Zhang J, Shi J, Liu S. Solid acid catalyzed glucose conversion to ethyl levulinate. Applied Catalysis A: General. 2011;397:259–265.

Perez R.F, Fraga M.A. Hemicellulose-derived chemicals: one-step production of furfuryl alcohol from xylose. Green Chemistry. 2014;16:3942–3950.

Pham H.N, Pagan-Torres Y.J, Serrano-Ruiz J.C, Wang D, Dumesic J.A, Datye A.K. Improved hydrothermal stability of niobia-supported Pd catalysts. Applied Catalysis A: General. 2011;397(1–2):153–162.

Pileidis F.D, Tabassum M, Coutts S, Titirici M.M. Esterification of levulinic acid into ethyl levulinate catalysed by sulfonated hydrothermal carbons. Chinese Journal of Catalysis. 2014;35:929–936.

Quian Y, Zhu L, Wang Y, Lu X. Recent progress in the development of biofuel 2,5-dimethylfuran. Renewable and Sustainable Energy Reviews. 2015;41:633–646.

Roman-Leshkov Y, Barrett C.J, Liu Z.Y, Dumesic J.A. Production of dimethylfuran for liquid fuels from biomass-derived carbohydrates. Nature. 2007;447:982–985.

Rothamer D.A, Jennings J.H. Study of the knocking propensity of 2,5-dimethylfuran–gasoline and ethanol–gasoline blends. Fuel. 2012;98:203–212.

Saravanamurugan S, Riisager A. Solid acid catalysed formation of ethyl levulinate and ethyl glucopyranoside from mono- and disaccharides. Catalysis Communications. 2012;14:71–75.

Sen S.M, Martín-Alonso D, Wettstein S.G, Gürbüz E.I, Henao C.A, Dumesic J.A, Maravelias C.T. A sulfuric acid management strategy for the production of liquid hydrocarbon fuels via catalytic conversion of biomass derive levulinic acid. Energy & Environmental Science. 2012;5:9690–9697.

Sen S.M, Gürbüz E.I, Wettstein S.G, Martín-Alonso D, Dumesic J.A, Maravelias C.T. Production of butene oligomers as transportation fuels using butene for esterification of levulinic acid from lignocellulosic biomass: process synthesis and technoeconomic evaluation. Green Chemistry. 2012;14:3289–3294.

Serrano-Ruiz J.C, Braden D.J, West R.M, Dumesic J.A. Conversion of cellulose to hydrocarbon fuels by progressive removal of oxygen. Applied Catalysis B: Environmental. 2010;100(1–2):184–189.

Serrano-Ruiz J.C, Wang D, Dumesic J.A. Catalytic upgrading of levulinic acid to 5-nonanone. Green Chemistry. 2010;12(4):574–577.

Shabaker J.W, Dumesic J.A. Kinetics of aqueous–phase reforming of oxygenated hydrocarbons: Pt/Al2O3 and Sn–Modified Ni catalysts. Industrial & Engineering Chemistry Research. 2004;43:3105–3112.

Sitthisa S, Resasco D.E. Hydrodeoxygenation of furfural over supported metal catalysts: a comparative study of Cu, Pd and Ni. Catalysis Letters. 2011;141(6):784–791.

Tanksale A, Wong Y, Beltramini J.N, Lu G.Q. Hydrogen generation from liquid phase catalytic reforming of sugar solutions using metal–supported catalysts. International Journal of Hydrogen Energy. 2007;32:717–724.

Tanksale A, Beltramini J.N, Lu G.M. A review of catalytic hydrogen production processes from biomass. Renewable and Sustainable Energy Reviews. 2010;14:166–182.

Thananatthanachon T, Rauchfuss T.B. Efficient production of the liquid fuel 2,5-dimethylfuran from fructose using formic acid as a reagent. Angewandte Chemie International Edition. 2010;49:6616–6618.

Tominaga K, Mori A, Fukushima Y, Shimada S, Sato K. Mixed-acid systems for the catalytic synthesis of methyl levulinate from cellulose. Green Chemistry. 2011;13:810–812.

Uddin M.N, Daud W.M.A.W. Technological diversity and economics: coupling effects on hydrogen production from biomass. Energy & Fuels. 2014;28(7):4300–4320.

Upare P.P, Hwang D.W, Hwang Y.K, Lee U.-H, Hong D.-Y, Chang J.-S. An integrated process for the production of 2,5-dimethylfuran from fructose. Green Chemistry. 2015;17:3310–3313.

Wang G.H, Hilgert J, Richter F.H, Wang F, Bongard H.J, Spliethoff B, Weidenthaler C, Schuth F. Platinum–cobalt bimetallic nanoparticles in hollow carbon nanospheres for hydrogenolysis of 5-hydroxymethylfurfural. Nature Materials. 2014;13:293–300.

Wang L, Xiao F.S. Nanoporous catalysts for biomass conversion. Green Chemistry. 2015;17:24–39.

West R.M, Liu Z.Y, Peter M, Dumesic J.A. Liquid alkanes with targeted molecular weights from biomass-derived carbohydrates. ChemSusChem. 2008;1(5):417–424.

Wright W.R.H, Palkovits R. Development of heterogeneous catalysts for the conversion of levulinic acid to γ-valerolactone. ChemSusChem. 2012;5(9):1657–1667.

Yan K, Wu G, Wen J, Chen A. One-step synthesis of mesoporous H4SiW12O40-SiO2 catalysts for the production of methyl and ethyl levulinate biodiesel. Catalysis Communications. 2013;34:58–63.

Yan K, Wu G, Lafleur T, Jarvis C. Production, properties and catalytic hydrogenation of furfural to fuel additives and value-added chemicals. Renewable and Sustainable Energy Reviews. 2014;38:663–676.

Yang P, Cui Q, Zu Y, Liu X, Lu G, Wang Y. Catalytic production of 2,5-dimethylfuran from 5-hydroxymethylfurfural over Ni/Co3O4 catalyst. Catalysis Communications. 2015;66:55–59.

Yu W, Tang Y, Mo L, Chen P, Lou H, Zheng X. One-step hydrogenation-esterification of furfural and acetic acid over bifunctional Pd catalysts for bio-oil upgrading. Bioresource Technology. 2011;102(17):8241–8246.

Yu W, Tang Y, Mo L, Chen P, Lou H, Zheng X. Bifunctional Pd/Al-SBA-15 catalyzed one step hydrogenation-esterification of furfural and acetic acid: a model reaction for catalytic upgrading of bio-oil. Catalysis Communications. 2011;13(1):35–39.

Yuan Z, Zhang Z, Zheng J, Lin J. Efficient synthesis of promising liquid fuels 5-ethoxymethylfurfural from carbohydrates. Fuel. 2015;150:236–242.

Zaytseva Y.A, Panchenko V.N, Simonov M.N, Shutilov A.A, Zenkovets G.A, Renz M, Simakova I.L, Parmon V.N. Effect of gas atmosphere on catalytic behavior of zirconia, ceria, and ceria-zirconia catalysts in valeric acid ketonization. Topics in Catalysis. 2013;56:846–855.

Zhao Y, Fu Y, Guo Q.-X. Production of aromatic hydrocarbons through catalytic pyrolysis of γ-valerolactone from biomass. Bioresource Technology. 2012;114:740–744.

Zu Y.H, Yang P.P, Wang J.J, Liu X.H, Ren J.W, Lu G.Z, Wang Y.Q. Efficient production of the liquid fuel 2,5-dimethylfuran from 5-hydroxymethylfurfural over Ru/Co3O4 catalyst. Applied Catalysis B: Environmental. 2014;146:244–248.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset