Synthesis of Diesel and Jet Fuel Range Alkanes ... - ACS Publications

Jul 26, 2017 - ABSTRACT: A route was developed for the synthesis of diesel and jet fuel range C9 and C10 alkanes with furfural and angelica lactone, w...
3 downloads 0 Views 823KB Size
Subscriber access provided by BOSTON UNIV

Article

Synthesis of Diesel and Jet Fuel Range Alkanes with Furfural and Angelica Lactone Jilei Xu, Ning Li, Xiaofeng Yang, Guangyi Li, Aiqin Wang, Yu Cong, Xiaodong Wang, and Tao Zhang ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.7b01992 • Publication Date (Web): 26 Jul 2017 Downloaded from http://pubs.acs.org on July 28, 2017

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

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

Page 1 of 20

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

ACS Catalysis

Synthesis of Diesel and Jet Fuel Range Alkanes with Furfural and Angelica Lactone Jilei Xu,† Ning Li,†,‡,* Xiaofeng Yang,† Guangyi Li,† Aiqin Wang,†,‡ Yu Cong,† Xiaodong Wang,† and Tao Zhang†,‡,* †

State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences,

Dalian 116023, China. ‡

iChEM (Collaborative Innovation Center of Chemistry for Energy Materials), Dalian Institute of

Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China

ABSTRACT: A route was developed for the synthesis of diesel and jet fuel range C9 and C10 alkanes with furfural and angelica lactone which can be obtained from hemicellulose and cellulose. It was found that angelica lactone is more reactive than levulinic acid or its other derivates in the aldol condensation with furfural. Among the investigated catalysts, Mn2O3 was found to be the most active and very stable for the aldol condensation of furfural and angelica lactone. Over it, high carbon yield of C 10 oxygenates (about 96%) can be achieved by the aldol condensation of furfural and angelica lactone under mild conditions (353 K, 4 h). By the hydrogenation and hydrodeoxygenation of the aldol condensation product over the Pd/C and Pd-FeOx/SiO2 catalysts, high carbon yields (~96%) of C9 and C10 alkanes were obtained.

KEYWORDS:

lignocellulose,

diesel

and

jet

fuel,

angelica

hydrodeoxygenation

ACS Paragon Plus Environment

lactone,

aldol

condensation,

ACS Catalysis

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

Page 2 of 20

Introduction With the increasing of social concern about the energy and environmental problems, the catalytic conversion of renewable and inedible biomass to fuels1 and chemicals2 has become a hot research topic. Lignocellulose is the main component of agricultural and forestry wastes. Compared with other forms of biomass, lignocellulose is much cheaper and more abundant. Diesel and jet fuel are two kinds of often used transportation fuels. Following the pioneering work of Dumesic,3 Huber4 and Corma,5 the synthesis of diesel and jet fuel range hydrocarbons with the lignocellulose derived platform compounds has drawn tremendous attention.6 Furfural and levulinic acid are two important platform compounds which have been produced in industrial scale by the hydrolysis/dehydration of the hemicellulose and cellulose.7 Angelica lactone is the dehydration product of levulinic acid.6g, 6j In this work, a C10 oxygenates (i.e. compound 1A in Scheme 1) was obtained by the solvent-free aldol condensation of furfural and angelica lactone. Compared with levulinc acid, methyl levulinate and γ-valerolactone, angelica lactone is more reactive in the aldol condensation with furfural. Among the invesigated catalysts, Mn2O3, an amphoteric metal oxide which has low acidity and low basity, was found to be the most active and stable for the aldol condensation reaction. The C10 oxygenate as obtained was further hydrogenated and hydrodeoxygenated to C9 and C10 alkanes. The stragey of this work is proposed in Scheme 1. Scheme 1. Stragey for the Synthesis of C9 and C10 Alkanes with Furfural and Angelica Lactone which can be Derived from Hemicellulose and Cellulose.

ACS Paragon Plus Environment

Page 3 of 20

Results and Discussion In the first part of this work, we investigated the solvent-free aldol condensation of furfural and angelica lactone over a series of metal oxides. To facilitate the comparsion, we fixed the ratio of metal atom in the catalyst to molecule of furfural (or angelica lactone) in the feedstock as 1: 5. According to the analysis of HPLC and NMR spectra (see Figure S1 and Figure S2 in supporting information), 3-(furan-2ylmethylene)-5-methylfuran-2(3H)-one (i.e. compound 1A in Scheme 1) was identified as the main product from the solvent-free aldol condensation of furfural and angelica lactone. This compound has the carbon chain length of diesel and jet fuel range. Therefore, it can be used as a potential precursor in the synthesis of

renewable diesel and jet fuel. Among the investigated metal oxide catalysts, Mn2O3

exhibited the highest acitivity and selectivity for the aldol condensation of furfural and angelica lactone (see Figure 1). Over this catalyst, furfural was completely converted and high carbon yield of 1A (96%) was acheived after the reaction was carried out at 353 K for 4 h. 100

100

Selectivity of 1A (%)

Furfural conversion (%) Carbon yield of 1A (%)

80

80

Selectivity of 1A (%)

Conversion or yield (%)

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

ACS Catalysis

60 40 20

60

40

20

0

0 O O 3 gO M Mn 2 La 2 3

O3 r O3 O O 5 TiO 2 nO 2 Zn S Fe 2 C 2 Nb 2

O3 O 3 gO M Mn 2 La 2

Zn

O

Mn

O3 r O3 O 5 TiO 2 nO 2 S Fe 2 C 2 Nb 2 O 2 a 2O 3 gO M L

Zn

O

O3 r O3 O5 Fe 2 C 2 Nb 2

Figure 1. Conversions of furfural, carbon yields and selectivities of 1A over different metal oxide

catalysts. Reaction conditions: 10 mmol angelica lactone, 10 mmol furfural, 2 mmol MgO, ZnO, TiO2, SnO2 or 1 mmol Mn2O3, La2O3, Fe2O3, Cr2O3, Nb2O5; 353 K, 4 h. Besides the Mn2O3 catalyst, the La2O3, MgO and ZnO catalysts were also found to be highly active for the aldol condensation of furfural and angelica lactone. Over them, furfural was compeletely converted.

ACS Paragon Plus Environment

ACS Catalysis

However, the selectivities of 1A over these catalysts were slightly lower than that over the Mn2O3 catalyst, which leads to the lower carbon yields of 1A over these catalysts. Compared with the La2O3, MgO and ZnO catalysts, the Fe2O3 catalyst is more selective for the aldol condensation of furfural and angelica lactone, but this catalyst is still less active and selective than the Mn2O3 catalyst. Subsequently, we also compared the catalytic performances of different manganese oxides for the aldol condensation of furfural and angelica lactone. To do this, we chose MnO, Mn2O3 and MnO2 as representatives for manganese oxides with different valence states. From the results illustrated in Figure 2, we can see that all of these manganese oxides are active for the aldol condensation of furfural and angelica lactone. Under the investigated reaction conditions, furfural was compeletly converted, high carbon yields (> 65%) of 1A were acheived over these catalysts. The carbon yields (or selectivity) of 1A over these manganese oxides decreases in the order of decrease in the order of Mn2O3 > MnO2 > MnO. From the XRD patterns of the used catalysts (see Figure S3 in supporting information), it was noticed that the Mn2O3 and MnO kept same but a part of MnO2 was reduced to Mn2O3 (which was labeled by black diamond in Figure S3c) after being used in the activity tests. Based on these results, we think that the Mn2O3 is the most active sites for the aldol condensation of furfural and angelica lactone.

Furfural conversion (%)

Carbon yield of 1A (%)

100

Conversion or yield (%)

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

Page 4 of 20

80 60 40 20 0

MnO2

Mn2O3

MnO

Figure 2. Conversions of furfural, carbon yields of 1A over different manganese oxide catalysts. Reaction conditions: 10 mmol angelica lactone, 10 mmol furfural, 2 mmol MnO, MnO2 or 1 mmol Mn2O3; 353 K, 4 h.

ACS Paragon Plus Environment

Page 5 of 20

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

ACS Catalysis

As we know, levulinic acid is the feedstock in the production of angelica lactone. Methyl levulinate and γ-valerolactone can be obtained by the esterification or hydrogenation of levulinic acid.8 In this work, we also studied the solvent-free aldol condenstaion of these compounds with furfural. The reactions were carried out over the Mn2O3 catalyst under the same reaction conditions (353 K, 4 h) as what we used for the aldol condensation of furfural and angelica lactone. From the analysis of HPLC, no aldol condensation product was identified, which means that these compounds do not react with furfural under the investigated conditions. Based on these results, we can see that the reactivity of angelica lactone is evidently higher than levulinic acid, methyl levulinate and γ-valerolactone, which is advantagous in the industrialized production of diesel and jet fuel precursor. The relatively higher reactivity of angelica lactone can be rationalized by the higher stability of carbanion which is caused by its conjugate chemical structure. According to the DFT calculation results illustrated in Figure 3 and the Table S1 of supporting information, the energy difference between angelica lactone and its carbanion is evidently lower than those between the levulinic acid, methyl levulinate, γ-valerolactone and corresponding carbanions, which may be the reason for the higher reactivity of angelica lactone.

Figure 3. Energy differences between γ-valerolactone, angelica lactone, methyl levulinate and their corresponding carbanions. The detail information for the DFT calculation was given in supporting information. The white balls, grey balls and red balls in the figure account for hydrogen atoms, carbon atoms and oxygen atoms, respectively.

ACS Paragon Plus Environment

ACS Catalysis

In the previous work of Wang et al.9 and our group10, C15 and C10 diesel and jet fuel precursors have been synthesized by the aldol condensation of isopropyl levulinate (or sodium levulinate) and furfural. Compared with these routes, the new route reported in this work has disadvantange because it contaions an additional levulinic acid dehydration step for the production of angelica lactone. However, the new route doesn’t need to neutralization the aldol condensation product with acid before HDO process. This will reduced the cost of acid and the waste of sodium salt which will generated as a by-product.

60

Conversion or yield (%)

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

Page 6 of 20

40 Furfural conversion (%) Carbon yield of 1A (%)

20

0 0

1

2 Recycle time

3

4

Figure 4. Conversion of furfural and carbon yield of 1A over the Mn2O3 catalyst as a function of recycle time. Reaction conditions: 10 mmol angelica lactone, 10 mmol furfural, 1 mmol Mn2O3; 353 K, 2 h. The stability of Mn2O3 catalyst in the solvent-free aldol condensation of furfural and angelica lactone was investigated. To do this, the Mn2O3 catalyst was repeatedly used in the solvent-free aldol condensation of furfural and angelica lactone under the reaction conditions when ~55% furfural conversion was acheived over the fresh Mn2O3 catalyst. After each usage, the catalyst was throughly washed with ethanol and dried at 333 K for 8 h. According to Figure 4, the Mn2O3 catalyst was stable under the investigated conditions. No evident deactivation was observed during 5 usages. According to ICP result, small amount of Mn was identified in the aldol condensation product. Based on these results, we believe that there is leaching of Mn during the reaction. However, the leaching of Mn species during the reaction has no evident influence the activity of the Mn2O3 catalyst. In the most of previous work

ACS Paragon Plus Environment

Page 7 of 20

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

ACS Catalysis

about the aldol condensation of furfural over heterogenous catalysts,11 the activity of catalyst decreased due to the Cannizzaro reaction (or polymerization) of furfural. To overcome this problem, the catalysts must be regnerated at higher temperature in oxidative atomophere to remove the organic compounds which were strongly adsorbed on the catalysts. Compared with those reaction systems, the regenaration of catalyst in this reaction system is much easier. This is advantageous in future application. To figure out the reason for the excellent performance of the Mn2O3 catalyst, we characterized it by CO2-TPD and NH3-TPD. From the results shown in the Figure S4 of the supporting information, we can see that the Mn2O3 is an amphoteric metal oxide which has weak acid sites and weak base sites at the same time. This may be the reason for the higher activity and selectivity of the Mn2O3 catalyst in the aldol condensation of furfural and angelica lactone. As we know, furfural will polymerize in the presence of strong acid sites, which will lead to the lower activity and selectivty to 1A over strong acid catalysts (such as Nb2O5 and Cr2O3). Analogously, the metathetical reaction between angelica lactone and strong base sites (or the Cannizzaro reaction of furfural catalyzed by strong base sites) will also decrease the activity and selectivty to 1A over strong base catalysts (such as MgO and La2O3). In this work, the Mn2O3 catalyst has proper acid strength and base strength which are effective for the aldol condensation of furfural and angelica lactone but not enough for the polymerization of furfural and the metathetical reaction of angelica lactone (or the Cannizzaro reaction of furfural). This may be one reason for the excellent performance of Mn2O3 in the aldol condensation of furfural and angelica lactone. Furthermore, as what has been suggested in literature,6m, 12 the synergism effect between acid sites and base site is benefical for the aldol condensation reactions. Therefore, the simultaneous presence of acid site and base site on the surface of Mn2O3 may be another reason for the high activity of Mn2O3 in the aldol condensation of furfural and angelica lactone. The reaction mechanism for the aldol condensation of furfural and angelica lactone was proposed in Scheme 2. In the first step, the angelica lactone is adsrobed and deprotonated to carbanion on the base sites of Mn2O3. In the second step, the carbanion attacks the carbonyl group of furfural which is adsorbed

ACS Paragon Plus Environment

ACS Catalysis

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

Page 8 of 20

on the acid sites of Mn2O3. As Corma et al. suggested previously,13 weak acid site has activation effect on the carbonyl group (by polarizing the C=O bond and increasing the eletrophilicity of the carbonyl group), which will decrease the energy barrier for the condensation of this molecule with a carbanion. In the third step, the product generated in the second step was protonated and dehydrated to 1A. Scheme 2. Reaction Mechanism for the Aldol Condensation of Furfural and Angelica Lactone over the Mn2O3 Catalyst.

Subsequently, we also explored the solvent-free aldol condensation of angelica lactone and other lignocellulose derived carbonyl compounds. Among them, the 5-hydroxylmethylfurfural can be obtained from the hydrolysis/dehydration of cellulose14. Benzene formaldehyde is a representative of the aldehydes which can be obtained from lignin.15 Acetone can be produced from the ABE fermentation of lignocellulose.16 Butanal can be produced from the partial oxidation of dehydrogenation of the biobutanol. Cyclopentanone can be obtained from the aqueous phase selective hydrogenation of furfural. 17 From the results listed in Table 1, we can see that angelica lactone can react with the investigated aldehydes under mild conditions (353 K, 4 h). From these reactions, several C9-C12 oxygenates can be obtained. As a potential application, these compounds can be used as the precursors for the synthesis of renewable diesel and jet fuel. Compared with aldehydes, ketones have lower reactivity. Under the same reaction conditions, no aldol condensation products were obtained over the Mn2O3 catalyst when acetone and cyclopentanone were used as the feedstocks. This phenomena can be explained by two reasons: 1) As we know, the carbonyl groups of aldehydes are connected with a alkyl group and a hydrogen atom, while

ACS Paragon Plus Environment

Page 9 of 20

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

ACS Catalysis

the carbonyl groups of ketones are connected with a alkyl groups. Because the alkyl groups have higher electronic effect than hydrogen atom, the carbonyl groups in aldehydes are more electrophilic and easier to be attacked by the carbanion generated from the deprotonation of angelica lactone than those in ketones. 2) The alkyl groups are larger than hydrogen atom. From the pointview of steric effect, the carbonyl groups in aldehydes are also easier to be attacked by the carbanion than those in ketones. Due to these reasons, 5-hydroxylmethylfurfural, butanal, and benzene formaldehyde are more reactive in the aldol condensation with angelica lactone. Analogously, we also investigated the solvent-free aldol condensation of furfurl and its partial oxidization prouct 2(5H)-furanone18 (which has similar chemical structure as angelica lactone). It was found that this compound is less reactive than angelica lactone. Under the investigated conditions, the furfural conversion (1.8%) over the Mn2O3 catalyst is very low. No C9 oxygenate from the aldol condenation of furfural and 2(5H)-furanone was obtained. Table 1. Conversions of carbonyl compounds and the carbon yields of aldol condensation products from the reactions of angelica lactone with various lignocellulosic carbonyl compounds. Reaction conditions: 10 mmol angelica lactone, 10 mmol carbonyl compound, 1 mmol Mn2O3; 353 K, 4 h. Reactant

Product

Conversion (%)

Carbon yield (%)

100

75.7

100

77.9

100

68.9

0

0

0

0

ACS Paragon Plus Environment

ACS Catalysis

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

Page 10 of 20

As the final aim of this work, we investigated the catalytic conversion of 1A to diesel and jet fuel range alkanes. Because 1A is a solid at room temperature (see Figure S5a in supporting information), it can’t be delivered by HPLC pump. To increase the fluidity of 1A, we hydrogenated it before the solvent-free hydrodeoxygenation (HDO) tests (the detail information of the hydrogenation step was given in support information). From the analysis of the hydrogenation products with GC-MS (see Figure S6 and Figure S7 in supporting information), it was found that 1A was totally converted to a mixture of 3-(5hydroxypentyl)-5-methyldihydrofuran-2(3H)-one and 3-(2-hydroxypentyl)-5-methyldihydrofuran-2(3H)one (i.e. 1B and 1C in Scheme 3) after being hydrogenated over the Pd/C catalyst at 433 K. The hydrogenation products exist as liquid at room temperature and can be directly used for the solvent-free HDO (see Figure S5b). Scheme 3. Reaction Pathway for the Generation of 1B and 1C from the Hydrogenation of 1A.

The solvent-free HDO tests of hydrogenated 1A (i.e. a mixture of 1B and 1C) were carried out over the Pd/SiO2, Pd-FeOx/SiO2 and FeOx/SiO2 catalysts at 623 K. From the analysis of the HDO products, it was noticed that the 1B and 1C were completely converted over these catalysts. High carbon yields of diesel and jet fuel range C8-C10 alkanes were achieved over the Pd/SiO2 and Pd-FeOx/SiO2 catalysts. These alkanes have jet fuel range carbon chain length and low freezing points (see Table S2). Therefore, they can be directly blended into conventional jet fuel without hydroisomerization. In contrast, the 1B and 1C were only partially hydrodeoxygenated over the FeOx/SiO2 catalyst (see Figure S8 in supporting information).

ACS Paragon Plus Environment

Page 11 of 20

100 C1-C4 light alkanes

Carbon yields of alkanes (%)

C5-C7 gasoline range

80

alkanes C8-C10 diesel and jet fuel range alkanes

60

40

20

0

Pd-FeOx/SiO2

Pd/SiO2

FeOx/SiO2

Figure 5. Carbon yields of different alkanes from the solvent-free HDO of hydrogenated 1A over the Pd/SiO2, Pd-FeOx/SiO2 and FeOx/SiO2 catalysts. Reaction conditions: 623 K, 6 MPa, 1.8 g catalyst; liquid

Carbon yield (%)

feedstock flow rate 0.033 mL min-1, hydrogen flow rate: 120 mL min-1. 40

(a)

Pd/SiO2

30 20 10 0 C1

C2

C3

C4

C5

C6

C7

C8

C9

C10

C7

C8

C9

C10

100 Carbon yield (%)

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

ACS Catalysis

(b)

Pd-FeOx/SiO2

80 60 40 20 0 C1

C2

C3

C4

C5

C6

Figure 6. Carbon distributions of the alkane products from the HDO of hydrogenated 1A over the Pd/SiO2 (a) and Pd-FeOx/SiO2 (b) catalysts. Reaction conditions: 623 K, 6 MPa, 1.8 g catalyst; liquid feedstock flow rate 0.033 mL min-1, hydrogen flow rate: 120 mL min-1.

ACS Paragon Plus Environment

ACS Catalysis

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

Page 12 of 20

It is interesting that the carbon yield of C8-C10 alkanes over the Pd-FeOx/SiO2 catalyst is higher than that over the Pd/SiO2 catalyst (see Figure 5). From Figure 6, it is noticed that there is also big difference between the carbon distributions of the alkane products over the Pd/SiO2 and Pd-FeOx/SiO2 catalysts. Over the Pd/SiO2 catalyst, the HDO products were mainly composed of C8 and C9 alkanes. In contrast, C10 alkanes was obtained as the predominate HDO product over the Pd-FeOx/SiO2 catalyst. According to our previous work on the HDO of other furan compounds,19 these results can be explained becuase the presence of Fe species not only promoted the HDO reaction, but also restrained the decarbonylation and the decarboxylation which will lead to the generation of C8 and C9 alkanes (see Scheme 4). Scheme 4. Reaction Pathways for the Generation of C8 and C9 Alkanes during the HDO of 1B and 1C.

ACS Paragon Plus Environment

Page 13 of 20

100

Undecane

Dodecane

80

Carbon yield (%)

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

ACS Catalysis

60

40

20

0 Pd/SiO2

Pd-FeOx/SiO2

Figure 7. Carbon yields of alkanes from the HDO of ethyl laurate over the Pd/SiO2 and Pd-FeOx/SiO2 catalysts. Reaction conditions: 623 K, 6 MPa, 1.8 g catalyst; liquid feedstock flow rate 0.033 mL min-1, hydrogen flow rate: 120 mL min-1. In the previous work of Resasco et al. and our group,19-20 the restraining effect of Fe species on the decarbonylation over Ni and Pd catalysts has been proved. In this work, the inhibition effect of iron species on the decarboxylation over the Pd catalyst was further verified by the solvent-free HDO of ethyl laurate. According to the Figure 7 and Figure S9 in supporting information, the main HDO product over the Pd/SiO2 catalyst was undecane, while the major HDO product over the Pd-FeOx/SiO2 catalyst was dodecane. This result indicated that the modificiation of Pd/SiO2 catalyst with Fe species can evidently restrain the decarboxylation during the HDO process. Conclusion In summary, the solvent-free aldol condensation of furfural and angelica lactone followed by hydrogenation and hydrodeoxygenation (HDO) is a promising way for the production of renewable diesel and jet fuel range alkanes. Compared with levulinic acid and its other derivates, angelica lactone is much more reactive, which is advantageous in real application. Mn2O3 was found to be a highly active and stable catalyst for the solvent-free aldol condensation of furfural and angelica lactone, which can be

ACS Paragon Plus Environment

ACS Catalysis

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

Page 14 of 20

explained by its proper acid and base strengths. Over the Mn2O3 catalyst, a series of diesel and jet fuel precursors can be obtained by the solvent-free aldol condensation of angelica lactone and lignocellulosic aldehydes. The aldol condensation product of furfural and angelica lactone was further hydrogenated and hydrodeoxygenated to diesel and jet fuel range alkanes over Pd catalysts. The modification of Pd/SiO2 catalyst with iron species increases its HDO activity and restrains the unexpected decarbonylation and decarboxylation during the HDO process. Both effects lead to the high carbon yield of diesel and jet fuel range alkanes over the Pd-FeOx/SiO2 catalyst.

ACS Paragon Plus Environment

Page 15 of 20

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

ACS Catalysis

ASSOCIATED CONTENT Supporting Information. The following files are available free of charge. Experimental details. Characterization of the catalysts. Copies of HPLC, GC chromatograms and NMR spectra. (PDF) AUTHOR INFORMATION Corresponding Author *T.Z.: [email protected]; *N.L.: [email protected]. Notes The authors declare no competing financial interests. ACKNOWLEDGMENT This work was supported by the National Natural Science Foundation of China (no. 21603221; 21690082; 21690084; 21506213; 21476229; 21277140), Dalian Science Foundation for Distinguished Young Scholars (no. 2015R005), the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB17020100), Department of Science and Technology of Liaoning Province (under contract of 2015020086-101). Dr. Xu appreciates the Postdoctoral Science Foundation of China (2015M580235) and the dedicated grant for methanol conversion from DICP for funding this work.

ACS Paragon Plus Environment

ACS Catalysis

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

Page 16 of 20

REFERENCES (1)

(a) Huber, G. W.; Iborra, S.; Corma, A., Chem. Rev. (Washington, DC, U. S.) 2006, 106,

4044-4098; (b) Alonso, D. M.; Bond, J. Q.; Dumesic, J. A., Green Chem. 2010, 12, 1493-1513; (c) Matson, T. D.; Barta, K.; Iretskii, A. V.; Ford, P. C., J. Am. Chem. Soc. 2011, 133, 1409014097; (d) Alonso, D. M.; Wettstein, S. G.; Dumesic, J. A., Chem. Soc. Rev. 2012, 41, 80758098; (e) Barta, K.; Ford, P. C., Acc. Chem. Res. 2014, 47, 1503-1512. (2)

(a) Corma, A.; Iborra, S.; Velty, A., Chem. Rev. (Washington, DC, U. S.) 2007, 107,

2411-2502; (b) Climent, M. J.; Corma, A.; Iborra, S., Green Chem. 2011, 13, 520-540; (c) Gallezot, P., Chem. Soc. Rev. 2012, 41, 1538-1558; (d) Yong, G.; Zhang, Y.; Ying, J. Y., Angew. Chem., Int. Ed. 2008, 47, 9345-9348; (e) Li, X.; Wu, D.; Lu, T.; Yi, G.; Su, H.; Zhang, Y., Angew. Chem., Int. Ed. 2014, 53, 4200-4204; (f) Zhao, Z.; Arentz, J.; Pretzer, L. A.; Limpornpipat, P.; Clomburg, J. M.; Gonzalez, R.; Schweitzer, N. M.; Wu, T.; Miller, J. T.; Wong, M. S., Chem. Sci. 2014, 5, 3715-3728; (g) Wang, Y.; De, S.; Yan, N., Chem. Commun. (Cambridge, U. K.) 2016, 52, 6210-6224; (h) Chen, L.; Zhao, J.; Pradhan, S.; Brinson, B. E.; Scuseria, G. E.; Zhang, Z. C.; Wong, M. S., Green Chem. 2016, 18, 5438-5447. (3)

(a) Huber, G. W.; Chheda, J. N.; Barrett, C. J.; Dumesic, J. A., Science 2005, 308, 1446-

1450; (b) Bond, J. Q.; Alonso, D. M.; Wang, D.; West, R. M.; Dumesic, J. A., Science 2010, 327, 1110-1114. (4)

(a) Xing, R.; Subrahmanyam, A. V.; Olcay, H.; Qi, W.; van Walsum, G. P.; Pendse, H.;

Huber, G. W., Green Chem. 2010, 12, 1933-1946; (b) Olcay, H.; Subrahmanyam, A. V.; Xing, R.; Lajoie, J.; Dumesic, J. A.; Huber, G. W., Energy Environ. Sci. 2013, 6, 205-216. (5)

(a) Corma, A.; de la Torre, O.; Renz, M.; Villandier, N., Angew. Chem., Int. Ed. 2011, 50,

2375-2378; (b) Corma, A.; de la Torre, O.; Renz, M., Energy Environ. Sci. 2012, 5, 6328-6344.

ACS Paragon Plus Environment

Page 17 of 20

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

ACS Catalysis

(6)

(a) Harvey, B. G.; Quintana, R. L., Energy Environ. Sci. 2010, 3, 352-357; (b) Yan, N.;

Yuan, Y.; Dykeman, R.; Kou, Y.; Dyson, P. J., Angew. Chem., Int. Ed. 2010, 49, 5549-5553; (c) Anbarasan, P.; Baer, Z. C.; Sreekumar, S.; Gross, E.; Binder, J. B.; Blanch, H. W.; Clark, D. S.; Toste, F. D., Nature 2012, 491, 235-239; (d) Wang, X. Y.; Rinaldi, R., Energy Environ. Sci. 2012, 5, 8244-8260; (e) Sutton, A. D.; Waldie, F. D.; Wu, R. L.; Schlaf, M.; Silks, L. A.; Gordon, J. C., Nat. Chem. 2013, 5, 428-432; (f) Wang, X.; Rinaldi, R., Angew. Chem., Int. Ed. 2013, 52, 11499-11503; (g) Mascal, M.; Dutta, S.; Gandarias, I., Angew. Chem., Int. Ed. 2014, 53, 18541857; (h) Xia, Q.-N.; Cuan, Q.; Liu, X.-H.; Gong, X.-Q.; Lu, G.-Z.; Wang, Y.-Q., Angew. Chem., Int. Ed. 2014, 53, 9755-9760; (i) Prasomsri, T.; Shetty, M.; Murugappan, K.; Roman-Leshkov, Y., Energy Environ. Sci. 2014, 7, 2660-2669; (j) Xin, J.; Zhang, S.; Yan, D.; Ayodele, O.; Lu, X.; Wang, J., Green Chem. 2014, 16, 3589-3595; (k) Xu, H.; Wang, K.; Zhang, H.; Hao, L.; Xu, J.; Liu, Z., Catal. Sci. Technol. 2014, 4, 2658-2663; (l) Sacia, E. R.; Balakrishnan, M.; Deaner, M. H.; Goulas, K. A.; Toste, F. D.; Bell, A. T., ChemSusChem 2015, 8, 1726-1736; (m) Sankaranarayanapillai, S.; Sreekumar, S.; Gomes, J.; Grippo, A.; Arab, G. E.; Head-Gordon, M.; Toste, F. D.; Bell, A. T., Angew. Chem., Int. Ed. 2015, 54, 4673-4677; (n) Xu, H.; Yu, B.; Zhang, H.; Zhao, Y.; Yang, Z.; Xu, J.; Han, B.; Liu, Z., Chem. Commun. (Cambridge, U. K.) 2015, 51, 12212-12215. (7)

(a) Lange, J.-P.; van der Heide, E.; van Buijtenen, J.; Price, R., ChemSusChem 2012, 5,

150-166; (b) Xing, R.; Qi, W.; Huber, G. W., Energy Environ. Sci. 2011, 4, 2193-2205; (c) Wettstein, S. G.; Alonso, D. M.; Chong, Y. X.; Dumesic, J. A., Energy Environ. Sci. 2012, 5, 8199-8203; (d) Weingarten, R.; Conner, W. C.; Huber, G. W., Energy Environ. Sci. 2012, 5, 7559-7574.

ACS Paragon Plus Environment

ACS Catalysis

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

(8)

Page 18 of 20

(a) Gurbuz, E. I.; Alonso, D. M.; Bond, J. Q.; Dumesic, J. A., ChemSusChem 2011, 4,

357-361; (b) Alonso, D. M.; Wettstein, S. G.; Dumesic, J. A., Green Chem. 2013, 15, 584-595. (9)

Li, C.; Ding, D.; Xia, Q.; Liu, X.; Wang, Y., ChemSusChem 2016, 9, 1712-1718.

(10)

Liang, G.; Wang, A.; Zhao, X.; Lei, N.; Zhang, T., Green Chem. 2016, 18, 3430-3438.

(11)

(a) Barrett, C. J.; Chheda, J. N.; Huber, G. W.; Dumesic, J. A., Appl. Catal., B 2006, 66,

111-118; (b) Yang, J.; Li, N.; Li, S.; Wang, W.; Li, L.; Wang, A.; Wang, X.; Cong, Y.; Zhang, T., Green Chem. 2014, 16, 4879-4884; (c) Kikhtyanin, O.; Kelbichová, V.; Vitvarová, D.; Kubů, M.; Kubička, D., Catal. Today 2014, 227, 154-162; (d) Kikhtyanin, O.; Kubička, D.; Čejka, J., Catal. Today 2015, 243, 158-162. (12)

(a) Shen, W. Q.; Tompsett, G. A.; Xing, R.; Conner, W. C.; Huber, G. W., J. Catal. 2012,

286, 248-259; (b) Lang, W. Z.; Su, B.; Guo, Y. J.; Chu, L. F., Sci. China: Chem. 2012, 55, 11671174; (c) Shylesh, S.; Hanna, D.; Gomes, J.; Krishna, S.; Canlas, C. G.; Head-Gordon, M.; Bell, A. T., ChemCatChem 2014, 6, 1283-1290; (d) Shylesh, S.; Hanna, D.; Gomes, J.; Canlas, C. G.; Head-Gordon, M.; Bell, A. T., ChemSusChem 2015, 8, 466-472. (13)

Climent, M. A. J.; Corma, A.; Iborra, S.; Velty, A., J. Mol. Catal. A: Chem. 2002, 182–

183, 327-342. (14)

(a) Zhao, H.; Holladay, J. E.; Brown, H.; Zhang, Z. C., Science 2007, 316, 1597-1600; (b)

Cao, F.; Schwartz, T. J.; McClelland, D. J.; Krishna, S. H.; Dumesic, J. A.; Huber, G. W., Energy Environ. Sci. 2015, 8, 1808-1815; (c) Mascal, M.; Nikitin, E. B., Angew. Chem., Int. Ed. 2008, 47, 7924-7926. (15)

Zhou, Z.-Z.; Liu, M.; Li, C.-J., ACS Catal. 2017, 7, 3344-3348.

(16)

Harvey, B. G.; Meylemans, H. A., J. Chem. Technol. Biotechnol. 2011, 86, 2-9.

ACS Paragon Plus Environment

Page 19 of 20

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

ACS Catalysis

(17)

(a) Hronec, M.; Fulajtarova, K., Catal. Commun. 2012, 24, 100-104; (b) Yang, Y.; Du, Z.;

Huang, Y.; Lu, F.; Wang, F.; Gao, J.; Xu, J., Green Chem. 2013, 15, 1932-1940; (c) Zhang, G. S.; Zhu, M. M.; Zhang, Q.; Liu, Y. M.; He, H. Y.; Cao, Y., Green Chem. 2016, 18, 2155-2164. (18)

(a) Xiang, X.; Zhang, B.; Ding, G.; Cui, J.; Zheng, H.; Zhu, Y., Catal. Commun. 2016, 86,

41-45; (b) Li, X.; Lan, X.; Wang, T., Green Chem. 2016, 18, 638-642. (19)

Yang, J.; Li, S.; Zhang, L.; Liu, X.; Wang, J.; Pan, X.; Li, N.; Wang, A.; Cong, Y.; Wang,

X.; Zhang, T., Appl. Catal., B 2017, 201, 266-277. (20)

Sitthisa, S.; An, W.; Resasco, D. E., J. Catal. 2011, 284, 90-101.

ACS Paragon Plus Environment

ACS Catalysis

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

Table of contents entry

ACS Paragon Plus Environment

Page 20 of 20