+ All Categories
Home > Documents > Synthesis of 1-Octanol and 1,1-Dioctyl Ether from Biomass...

Synthesis of 1-Octanol and 1,1-Dioctyl Ether from Biomass...

Date post: 04-Apr-2018
Category:
Upload: buinga
View: 224 times
Download: 2 times
Share this document with a friend
17
Biogenic Alcohols DOI: 10.1002/anie.201203669 Synthesis of 1-Octanol and 1,1-Dioctyl Ether from Biomass-Derived Platform Chemicals** Jennifer Julis and Walter Leitner* As a consequence of diminishing fossil resources and global endeavors to reduce anthropogenic carbon dioxide emissions, biomass-derived substrates are receiving increasing attention in the effort to establish renewable supply chains for trans- portation fuels and chemical products. [1] Carbohydrates con- stitute the largest fraction of biomass feedstock. The con- version of carbohydrates from a set of platform molecules into tailor-made products can be envisaged through selective catalytic transformation steps. [1a, 2] Primary alcohols of medium chain length are very important industrial products, as they are valuable com- pounds for the production of detergents and surfactants, in perfumery, and as flavors. 1-Octanol is of particular impor- tance and is also used for the synthesis of 1-octene, an important co-monomer for polyethylene. 1-Octanol is pre- dominantly synthesized either by the reaction of ethylene with triethylaluminum (Alfen process) or by oxo synthesis starting from n-heptene, which are both petrochemical processes. [3] Aliphatic alcohols from biomass are accessible by the reduction of fatty acids, but this is commercially exploited almost exclusively for long carbon chains ( C 12 ). [1a] Carbo- hydrate-based alcohols are currently limited to short carbon chains ( C 4 ) and are obtained through fermentation. [1a, 4] In contrast, the formation of alcohols of medium chain length from lignocellulosic platform chemicals is described in only very few cases and is not yet synthetically exploited. [5] 1- Pentanol has been observed as a by-product, for example, in the hydrogenolysis of tetrahydrofurfuryl alcohol and in the selective transformation of levulinic acid into 2-methyltetra- hydrofuran. [2, 6] Herein, we describe the highly selective catalytic synthesis of the linear primary C 8 alcohol products 1-octanol and dioctyl ether from the biomass-derived platform molecule furfural [7] and acetone, which is also accessible from carbohy- drates, [1c, 4] at least in principle. This opens a general strategy for the synthesis of medium-chain-length alcohols from carbohydrate feedstock. Recently, we proposed the concept of synthetic pathway design for biomass-derived products in analogy to the retrosynthetic analysis used in modern organic synthesis. [2] Scheme 1 shows how 1-octanol can be traced back to furfural and acetone as starting materials using this approach. These compounds are readily converted into furfuralacetone (FFA) by an aldol condensation ; [8] FFA can then be hydrogenated to 4-(2-tetrahydrofuryl)-2-butanol (THFA), [9] which might be converted into 1-octanol (1-OL) by selective deoxygenation and ring opening, provided that over-hydrogenation to the alkane can be avoided. [5c, 10] Therefore, the challenge in establishing this pathway lies in the development of a selective catalytic system that can give access to 1-octanol from THFA by deoxygenation of the secondary alcohol function coupled with the selective ring-opening of the tetrahydrofuryl ring by hydrogenolysis. Scheme 2 shows the possible products resulting from hydrogenation and dehydration of THFA using a multifunc- tional catalytic system that provides both transition-metal- based hydrogenation activity and Brønsted acidity. [1d] 2- Butyltetrahydrofuran (BTHF) is obtained by removal of the secondary hydroxy group. BTHF is an interesting molecule in its own right, for example, as a potential fuel additive. [5c, 11] Full deoxygenation of THFA leads to n-octane. [5c, 10a] The targets of the present study are the linear C 8 alcohol products Scheme 1. Retrosynthetic analysis for a pathway to 1-octanol using platform chemicals derived from lingocellulosic feedstock. Scheme 2. Possible C 8 products accessible by catalytic conversion of THFA by dehydration/hydrogenation reactions. [*] Dipl.-Chem. J. Julis, Prof. Dr. W. Leitner Institut fɒr Technische und Makromolekulare Chemie, RWTH Aachen University Worringerweg 1, 52074 Aachen (Germany) E-mail: [email protected] Homepage: http://www.itmc.rwth-aachen.de Prof. Dr. W. Leitner Max-Planck-Institut fɒr Kohlenforschung 45470 Mɒlheim an der Ruhr (Germany) [**] This work was performed as part of the Cluster of Excellence “Tailor- Made Fuels from Biomass”, which is funded by the Excellence Initiative of the German federal and state government to promote science and research at German universities. Supporting information for this article is available on the WWW under http://dx.doi.org/10.1002/anie.201203669. A ngewandte Chemi e 8615 Angew. Chem. Int. Ed. 2012, 51, 8615 –8619 # 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Transcript

Biogenic AlcoholsDOI: 10.1002/anie.201203669

Synthesis of 1-Octanol and 1,1-Dioctyl Ether from Biomass-DerivedPlatform Chemicals**Jennifer Julis and Walter Leitner*

As a consequence of diminishing fossil resources and globalendeavors to reduce anthropogenic carbon dioxide emissions,biomass-derived substrates are receiving increasing attentionin the effort to establish renewable supply chains for trans-portation fuels and chemical products.[1] Carbohydrates con-stitute the largest fraction of biomass feedstock. The con-version of carbohydrates from a set of platform moleculesinto tailor-made products can be envisaged through selectivecatalytic transformation steps.[1a, 2]

Primary alcohols of medium chain length are veryimportant industrial products, as they are valuable com-pounds for the production of detergents and surfactants, inperfumery, and as flavors. 1-Octanol is of particular impor-tance and is also used for the synthesis of 1-octene, animportant co-monomer for polyethylene. 1-Octanol is pre-dominantly synthesized either by the reaction of ethylenewith triethylaluminum (Alfen process) or by oxo synthesisstarting from n-heptene, which are both petrochemicalprocesses.[3]

Aliphatic alcohols from biomass are accessible by thereduction of fatty acids, but this is commercially exploitedalmost exclusively for long carbon chains (�C12).[1a] Carbo-hydrate-based alcohols are currently limited to short carbonchains (�C4) and are obtained through fermentation.[1a, 4] Incontrast, the formation of alcohols of medium chain lengthfrom lignocellulosic platform chemicals is described in onlyvery few cases and is not yet synthetically exploited.[5] 1-Pentanol has been observed as a by-product, for example, inthe hydrogenolysis of tetrahydrofurfuryl alcohol and in theselective transformation of levulinic acid into 2-methyltetra-hydrofuran.[2, 6]

Herein, we describe the highly selective catalytic synthesisof the linear primary C8 alcohol products 1-octanol anddioctyl ether from the biomass-derived platform moleculefurfural[7] and acetone, which is also accessible from carbohy-

drates,[1c,4] at least in principle. This opens a general strategyfor the synthesis of medium-chain-length alcohols fromcarbohydrate feedstock.

Recently, we proposed the concept of synthetic pathwaydesign for biomass-derived products in analogy to theretrosynthetic analysis used in modern organic synthesis.[2]

Scheme 1 shows how 1-octanol can be traced back to furfuraland acetone as starting materials using this approach. Thesecompounds are readily converted into furfuralacetone (FFA)by an aldol condensation;[8] FFA can then be hydrogenated to4-(2-tetrahydrofuryl)-2-butanol (THFA),[9] which might beconverted into 1-octanol (1-OL) by selective deoxygenationand ring opening, provided that over-hydrogenation to thealkane can be avoided.[5c,10] Therefore, the challenge inestablishing this pathway lies in the development of a selectivecatalytic system that can give access to 1-octanol from THFAby deoxygenation of the secondary alcohol function coupledwith the selective ring-opening of the tetrahydrofuryl ring byhydrogenolysis.

Scheme 2 shows the possible products resulting fromhydrogenation and dehydration of THFA using a multifunc-tional catalytic system that provides both transition-metal-based hydrogenation activity and Brønsted acidity.[1d] 2-Butyltetrahydrofuran (BTHF) is obtained by removal of thesecondary hydroxy group. BTHF is an interesting molecule inits own right, for example, as a potential fuel additive.[5c,11]

Full deoxygenation of THFA leads to n-octane.[5c,10a] Thetargets of the present study are the linear C8 alcohol products

Scheme 1. Retrosynthetic analysis for a pathway to 1-octanol usingplatform chemicals derived from lingocellulosic feedstock.

Scheme 2. Possible C8 products accessible by catalytic conversion ofTHFA by dehydration/hydrogenation reactions.

[*] Dipl.-Chem. J. Julis, Prof. Dr. W. LeitnerInstitut f�r Technische und Makromolekulare Chemie, RWTHAachen UniversityWorringerweg 1, 52074 Aachen (Germany)E-mail: [email protected]: http://www.itmc.rwth-aachen.de

Prof. Dr. W. LeitnerMax-Planck-Institut f�r Kohlenforschung45470 M�lheim an der Ruhr (Germany)

[**] This work was performed as part of the Cluster of Excellence “Tailor-Made Fuels from Biomass”, which is funded by the ExcellenceInitiative of the German federal and state government to promotescience and research at German universities.

Supporting information for this article is available on the WWWunder http://dx.doi.org/10.1002/anie.201203669.

AngewandteChemie

8615Angew. Chem. Int. Ed. 2012, 51, 8615 –8619 � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

lixl
高亮
lii
高亮
lii
文本注释
发酵
lii
文本注释
综合利用

(C8-OL), which can be directly formed as free 1-octanol (1-OL) by hydrogenolytic ring opening, or as dioctyl ether(DOE) upon reversible etherification.

Scheme 3 gives an overview of the components that wereselected for the generation of the catalytic systems in thepresent study. Ruthenium was introduced as the metalcomponent for hydrogenation because of its proven activityin transformations of FFA, THFA, and related mole-cules.[2,8, 12] Ruthenium nanoparticles stabilized in ionic liquids(Ru@IL) were investigated for this application, along withcommercially available heterogeneous catalysts. Brønstedacid additives, including functional ionic liquids (ILs) werechosen to control the acidity required for dehydration.[13] Inthe first series of experiments, the transformation of THFAunder hydrogen with ruthenium nanoparticles (particle size2–3 nm) was investigated. The nanoparticles were preparedby hydrogenation of [(cod)Ru(h3-C4H7)2] in the presence ofdifferent ILs.[8,13c]

As seen in Table 1, the Ru@[BCO2BIM][NTf2] catalystshowed no activity in the hydrogenolysis of THFA, indicatingthat the acidity of the carbonic acid function in the ionic liquidis too weak for the dehydration (Table 1, entry 1). In contrast,the performance of Ru@[BSO3BIM][NTf2] was dominated bythe acidity of the SO3H function, resulting mainly in ether-ification and isomerization of THFA (entry 2). However, itwas possible to reduce the acid-catalyzed side reactions by

diluting the Ru@[BSO3BIM][NTf2] catalyst with the inert IL[EMIM][NTf2], which resulted in significant formation of C8-OL (28 % overall yield of 1-OL and DOE) and highselectivity towards 1-OL (entry 3). Under the same condi-tions, the use of Ru@[BSO3BIM][OTf] or Ru@[BSO3N444]-[NTf2] resulted in a combination of hydrogenation and acidcatalysis that led to a remarkable BTHF yield of up to 75 %,with isomerization and etherification products of THFA asthe main side products (entries 4 and 5). Therefore, thesystem of entry 3 was chosen as first lead structure for furtherinvestigation.

Monitoring the reaction (Figure 1) revealed that acid-catalyzed self-etherification and isomerization of the sub-strate also occur with this system and dominate the con-version during the first two hours (Scheme 4). However, thehydrogenolysis of the secondary alcohol group in the sidechain of THFA reverses the etherification equilibrium, andthe deoxygenation product, BTHF, is accumulated witha maximum of 80% conversion in the reaction mixture aftereight hours. Selective ring opening of the tetrahydrofuryl ringin BTHF then yields 1-OL, which again subsequently under-goes partial etherification under the acidic conditions toproduce DOE. Further dehydration to n-octane remains

Scheme 3. Metal components and acidic additives, including ionicliquids (ILs), used for the multifunctional catalytic systems.

Table 1: Hydrogenolysis of THFA with Ru@IL.

Entry Ionic liquid Conv.[%]

BTHF[%]

1-OL[%]

DOE[%]

Other[a]

[%]

1[b] [BCO2BIM][NTf2] 0 – – – –2[b] [BSO3BIM][NTf2] >99 5.0 – – 953[c] [BSO3BIM][NTf2] >99 66.6 25 2.8 5.64[c] [BSO3BIM][OTf ] >99 69.0 9.7 1.0 21.25[c] [BSO3N444][NTf2] >99 74.4 9.6 3.2 12.86[c] [BSO3N888][NTf2] >99 75.3 3.2 – 21.5

[a] Other = isomers and etherification products of THFA. [b] 120 8C, H2

(120 bar), 15 h, Ru (0.016 mmol), acidic IL (0.11 mmol), THFA(1.57 mmol). [c] 120 8C, H2 (120 bar), 15 h, Ru (0.016 mmol), acidic IL(0.11 mmol), THFA (1.57 mmol), [EMIM][NTf2] (2.9 mL).

Figure 1. Reaction monitoring of the hydrogenolysis of THFA withRu@[BSO3BIM][NTf2]. Conversion (c&c), BTHF (c*c),1-OL (c*c), DOE (c~c), condensate (c*c). Forreaction conditions, see Table 1, entry 3. For structures, see Scheme 4.

Scheme 4. Consecutive hydrogenation/deyhdration pathway of THFA.

.AngewandteCommunications

8616 www.angewandte.org � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2012, 51, 8615 –8619

lixl
高亮

insignificant under these conditions, reflecting the greaterresistance of the primary alcohol versus secondary alcohols toacid-catalyzed hydrogenolysis.

The monitoring of the hydrogenolysis of THFA allowedelucidation of the complex network of consecutive reactions,which involve uni- and bimolecular transformations of thesubstrate and product, respectively. Therefore, the reactionparameters of temperature, reaction time, and substrateconcentration were selected for further optimization usingthe simplex algorithm, with the yield of 1-OL as the targetvalue (Figure 2). The parameter limits were set to T= 120–160 8C, t = 10–25 h and c0(THFA) = 0.2–0.8 molL�1, whilehydrogen pressure and the substrate-to-catalyst ratio werekept constant.

After 16 iteration steps, the alteration of the commandvariables reached a stable output within error tolerance.Under the optimized reaction conditions of 150 8C, 15 h, and

0.5 molL�1 substrate concentration, a maximum yield of 45%1-OL and an overall C8-OL yield of 71.4% was obtained. Inparticular, no over-hydrogenation to n-octane was observed.

Using the optimized conditions, a series of combinationsof Ru-based catalysts and acidic additives were evaluated forthe hydrogenolysis of THFA (Table 2, entries 1–5; see alsothe Supporting Information). Ru@[N444BSO3] gave a slightlyhigher combined yield for the C8-OL products with a lower 1-OL/DOE-ratio and small amounts of n-octane, as comparedto the imidazolium ionic liquid (entries 1 and 2). Commer-cially available Ru/C and Ru/Alox (Alox = aluminum oxide)catalysts also performed well (entries 3 and 4), providing upto 75 % yield of C8-OL with outstanding selectivity for theremoval of the secondary hydroxy function. Results variedsignificantly with the source of the catalysts (see theSupporting Information for details) and the data presentedin Table 2 were obtained with Ru/C (5%) obtained from thesupplier abcr. They were found to be reproducible within� 5.0% over three independent experiments and varied� 7.0% between two different catalyst batches. In combina-tion with p-toluenesulfonic acid (p-TsOH), a 77 % overallyield of C8-OL was achieved within 15 h, whereby the largercontent of DOE and the formation of significant amounts ofn-octane reflect the higher acidity of the reaction mixture(entry 5).

Once the transformation of THFA into C8-OL had beensuccessfully demonstrated, further development of the syn-thetic pathway shown in Scheme 1 was attempted. If thereaction sequence was performed in individual steps withintermediate isolation of the products, the cumulative yieldsof the individual steps (97%, 98 %, 77%) resulted in anoverall yield of 73%, based on furfural. A direct one-stepconversion of FFA into C8-OL using the multifunctionalcatalyst systems was not possible, because furfuralacetone(FFA) is very sensitive towards acidic conditions and formshumin-type products in the presence of p-TsOH or[BSO3BIM][NTf2].

However, a two-step one-pot process could successfullybe achieved when the acidic additive was introduced directlyinto the reaction vessel after full hydrogenation of FFA toTHFA (entry 6). After FFA was hydrogenated at 120 8C withH2 (120 bar) for two hours in the presence of Ru/C, the acidic

Figure 2. Optimization of reaction conditions for Ru@[BSO3BIM][NTf2]by the simplex algorithm (size of data points correspond to yields of 1-OL).

Table 2: One-pot catalytic synthesis of linear C8 alcohol products (C8-OL) from biomass-derived platform chemicals.[a]

Entry Substrate Steps Catalyst Additive BTHF[%]

1-OL[%]

DOE[%]

Octane[%]

Other[b]

[%]C8-OL

[%]

1 THFA 1[c] Ru@[BSO3BIM][NTf2] – 25.8 44.9 26.1 – 3.2 71.02 THFA 1[c] Ru@[BSO3N444][NTf2] – 11.7 41.5 34.5 2.4 9.9 76.03 THFA 1[c] Ru/C (5 wt % Ru) [BSO3BIM][NTf2] 47.8 20.5 19.4 – 12.3 39.44 THFA 1[c] Ru/Alox (5 wt % Ru) [BSO3BIM][NTf2] 23.0 43.6 31.8 – 1.6 75.45 THFA 1[c] Ru/C (5 wt % Ru) p-TsOH 8.7 24.7 51.9 8.5 6.2 76.66 FFA 2[d] Ru/C [BSO3BIM][NTf2] 5.4 48.8 44.2 – 1.8 93.07 FFA 2[d] Ru/C p-TsOH 2.1 18.1 53.0 16.9 10.8 71.18 furfural 3[e] Ru/C [BSO3BIM][NTf2] 23.5 32.5 19.8 – 24.2[f ] 52.3

[a] Mass distribution of the product mixture after pentane extraction according to GC analysis using n-tetradecane as internal standard; >99%conversion was observed in all cases. [b] Other =other isomers, mainly 2-propyltetrahydropyran. [c] THFA (1.57 mmol), Ru (0.016 mmol), acidicadditive (0.11 mmol) in [EMIM][NTf2] (2.9 mL), 150 8C, H2 (120 bar), 15 h. [d] 1st step: 120 8C, H2 (120 bar), 2 h, neat FFA (1.57 mmol), Ru(0.016 mmol); 2nd step: as in [c], 60 h. [e] 1st step: furfural/acetone 1:10, RT, 15 h, NaOH (50 mL, 1.0m); 2nd and 3rd step: as in [c] and [d],respectively. [f ] Isomers of THFA. Alox =aluminum oxide.

AngewandteChemie

8617Angew. Chem. Int. Ed. 2012, 51, 8615 –8619 � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org

lixl
高亮
lixl
高亮
lixl
高亮
lixl
高亮
一步法由糠醛合成C8-OL不可能,形成胡敏素
lii
高亮

IL [BSO3BIM][NTf2] and the [EMIM]-[NTf2] solvent were added, and thereaction mixture was treated further at150 8C with H2 (120 bar) for 45 h toobtain 1-OL and DOE in excellentcombined yields of 93%. Considerableamounts of n-octane were formed whenp-TsOH was used as an acidic additive toRu/C (entry 7). The ruthenium nanopar-ticle catalysts could not be used in thisprocedure, as they were found to bedeactivated by agglomeration during thefirst part of the sequence.

It is even possible to conduct this reaction sequence asa one-pot procedure starting from furfural as a platformchemical (Scheme 5). Furfural was transformed to FFA by analdol condensation in the presence of excess acetone usingsodium hydroxide under standard conditions (see Table 2 fordetails). After neutralization of the reaction mixture withaqueous HCl and evaporation of acetone, Ru/C was addedand the hydrogenation was carried out as above. After 2 h, theacidic IL [BSO3BIM][NTf2] and the solvent IL were intro-duced directly for the final step. The overall yield for C8-OLwas 54% with a higher 1-OL content, corresponding toa remarkable 33 % yield of the free alcohol. The somewhatlower overall yield can be attributed at least partly to thepresence of the salt resulting from the neutralization process.Using a base-catalyzed protocol for the aldol condensation[11]

and/or conducting the one-pot procedure in a flow system[14]

are possibilities for further development towards a fullyintegrated reaction sequence.

In summary, we have demonstrated for the first time theselective conversion of tetrahydrofurfurylacetone (THFA)and furfuralacteone (FFA) into 1-octanol and dioctyl ether bydehydration/hydrogenation using a multifunctional catalystsystem comprised of a Ru hydrogenation catalyst togetherwith an acidic additive, including functional ionic liquids. Upto 93% yield of the linear C8 alcohol products were obtainedand the new transformation was integrated into a completesequence starting from furfural and acetone, which gavea combined yield (of 1-octanol and dioctyl ether) of 73%overall in a step-wise procedure, and 54% overall in a one-potprocedure. Using the retrosyntheic approach[2] shown inScheme 1, other primary alcohols can be readily envisagedto be produced by analogous pathways using the correspond-ing ketones, RCOCH3. This opens a new general route frombiomass-derived platform molecules to medium-chain pri-mary alcohols, again demonstrating the viability of rationalpathway design for the exploration of lignocellulosic supplychains.

Experimental SectionAll reactions were carried out in a 10 mL stainless-steel high-pressurereactor with a glass inlet. Metal catalysts and ionic liquids werehandled under an argon atmosphere. Ru@ILs were freshly preparedbefore use by suspending [(cod)Ru(h3-C4H7)2] (5.0 mg, 0.016 mmol)in the corresponding ionic liquid (0.11 mmol), followed by hydro-genation at 60 8C with H2 (60 bar) for two hours. Ru/C (5 wt %, abcr)

was activated at 80 8C with H2 (100 bar) for 10 h prior to use. Ina typical hydrogenolysis reaction, 4-(2-tetrahydrofuryl)-2-butanol(THFA; 225.7 mg, 1.57 mmol), 1-ethyl-3-methylimidazolium bis(tri-fluoromethylsulfonyl)imide (2.9 mL) and the acidic additive(0.11 mmol) were added to the metal catalyst and the reactionmixture was stirred for 15 h at 150 8C with H2 (120 bar). Aftercarefully venting the reactor, the reaction mixture was extracted withpentane (3 � 20 mL). The colorless solution was concentrated underreduced pressure and the molar composition of the product mixturewas analyzed by GC with n-tetradecane as an internal standard. Fulldetails of the experimental and analytical procedures are provided inthe Supporting Information.

Received: May 11, 2012Published online: July 6, 2012

.Keywords: biomass conversion · homogeneous catalysis ·hydrogenation · linear alcohols · multifunctional catalysts

[1] a) A. Corma, S. Iborra, A. Velty, Chem. Rev. 2007, 107, 2411 –2502; b) J. J. Bozell, G. R. Petersen, Green Chem. 2010, 12, 539;c) Biorefineries—Industrial Processes and Products Status Quoand Futures Directions (Eds.: B. Kamm, P. R. Gruber, M.Kamm), Wiley-VCH, Weinheim, 2006 ; d) A. M. Ruppert, K.Weinberg, R. Palkovits, Angew. Chem. 2012, 124, 2614 – 2654;Angew. Chem. Int. Ed. 2012, 51, 2564 – 2601.

[2] F. M. A. Geilen, B. Engendahl, A. Harwardt, W. Marquardt, J.Klankermayer, W. Leitner, Angew. Chem. 2010, 122, 5642 – 5646;Angew. Chem. Int. Ed. 2010, 49, 5510 – 5514.

[3] W. K. R. Dittmeyer, G. Kreysa, A. Oberholz, Winnacker-K�ch-ler: Chemische Technik, Vol. 5, Wiley-VCH, Weinheim, 2003 –2005.

[4] a) T. Liu, C. Khosla, Annu. Rev. Genet. 2010, 44, 53 – 69; b) T. G.Lenz, A. R. Morelra, Ind. Eng. Chem. Prod. Res. Dev. 1980, 19,478 – 483.

[5] a) S. Koso, I. Furikado, A. Shimao, T. Miyazawa, K. Kunimori,K. Tomishige, Chem. Commun. 2009, 2035 – 2037; b) T. Buntara,S. Noel, P. H. Phua, I. Meli�n-Cabrera, J. G. de Vries, H. J.Heeres, Angew. Chem. 2011, 123, 7221 – 7225; Angew. Chem. Int.Ed. 2011, 50, 7083 – 7087; c) W. Xu, Q. Xia, Y. Zhang, Y. Guo, Y.Wang, G. Lu, ChemSusChem 2011, 4, 1758 – 1761; d) Thefollowing reference appeared while the manuscript was underreview: R. Alamillo, M. Tucker, M. Chia, Y. Pagan-Torres, J.Dumesic, Green Chem. 2012, 14, 1413 – 1419.

[6] M. Chatterjee, H. Kawanami, T. Ishizaka, M. Sato, T. Suzuki, A.Suzuki, Catal. Sci. Technol. 2011, 1, 1466 – 1471.

[7] a) C. D. Hurd, L. L. Isenhour, J. Am. Chem. Soc. 1932, 54, 317 –330; b) M. J. Antal, T. Leesomboon, W. S. Mok, G. N. Richards,Carbohydr. Res. 1991, 217, 71 – 85.

[8] C. J. Barrett, J. N. Chheda, G. W. Huber, J. A. Dumesic, Appl.Catal. B 2006, 66, 111 – 118.

Scheme 5. Integrated multi-step synthesis of linear C8 alcohol products (C8-OL) from furfuraland acetone as biogenic platform chemicals.

.AngewandteCommunications

8618 www.angewandte.org � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2012, 51, 8615 –8619

[9] J. Julis, M. Hçlscher, W. Leitner, Green Chem. 2010, 12, 1634 –1639.

[10] a) M. Chatterjee, K. Matsushima, Y. Ikushima, M. Sato, T.Yokoyama, H. Kawanami, T. Suzuki, Green Chem. 2010, 12,779 – 782; b) G. W. Huber, J. N. Chheda, C. J. Barrett, J. A.Dumesic, Science 2005, 308, 1446 – 1450; c) A. Hinz, G. Meyer,G. Sch�cking, Ber. Dtsch. Chem. Ges. (A and B Series) 1943, 76,676 – 689.

[11] a) J.-P. Lange, R. Price, P. M. Ayoub, J. Louis, L. Petrus, L.Clarke, H. Gosselink, Angew. Chem. 2010, 122, 4581 – 4585;Angew. Chem. Int. Ed. 2010, 49, 4479 – 4483; b) A. Brassat, M.Thewes, M. M�ther, S. Pischinger, Motortech. Z. 2010, 12, 922 –928.

[12] a) N. Yan, C. Zhao, C. Luo, P. J. Dyson, H. Liu, Y. Kou, J. Am.Chem. Soc. 2006, 128, 8714 – 8715; b) P. S. Campbell, C. C.

Santini, F. Bayard, Y. Chauvin, V. Colli�re, A. Podgorsek, M. F.Costa Gomes, J. S�, J. Catal. 2010, 275, 99 – 107; c) S. Crossley, J.Faria, M. Shen, D. E. Resasco, Science 2010, 327, 68 – 72.

[13] a) N. Yan, Y. Yuan, R. Dykeman, Y. Kou, P. J. Dyson, Angew.Chem. 2010, 122, 5681 – 5685; Angew. Chem. Int. Ed. 2010, 49,5549 – 5553; b) A. C. Cole, J. L. Jensen, I. Ntai, K. L. T. Tran,K. J. Weaver, D. C. Forbes, J. H. Davis, J. Am. Chem. Soc. 2002,124, 5962 – 5963; c) M. A. Liauw, S. Winterle, Chem. Ing. Tech.2010, 82, 1211 – 1214.

[14] a) U. Hintermair, T. Hçfener, T. Pullmann, G. Franci�, W.Leitner, ChemCatChem 2010, 2, 150 – 154; b) M. Ruta, G.Laurenczy, P. J. Dyson, L. Kiwi-Minsker, J. Phys. Chem. C2008, 112, 17814 – 17819.

AngewandteChemie

8619Angew. Chem. Int. Ed. 2012, 51, 8615 –8619 � 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org

Supporting Information

� Wiley-VCH 2012

69451 Weinheim, Germany

Synthesis of 1-Octanol and 1,1-Dioctyl Ether from Biomass-DerivedPlatform Chemicals**Jennifer Julis and Walter Leitner*

anie_201203669_sm_miscellaneous_information.pdf

1

 

 

 

Safety  Warning High-­‐pressure   experiments   with   compressed   hydrogen   must   be   carried   out   only   with   appropriate  

equipment  and  under  rigorous  safety  precautions.  

 

General  If  not  stated  otherwise,  the  syntheses  of  ionic  liquids  and  nanoparticle  solutions  were  carried  out  under  

argon   inert   gas   atmosphere  using   standard   Schlenk   technique.   Catalyst   solutions   and   substrates  were  

handled  under  air,  but  were  flushed  with  hydrogen  prior  starting  catalysis.    

 

Analytics  Conversion   and   selectivity   of   catalytic   reactions   were   determined   via   GC   using   a   Thermo   Scientific  

Chromatograph  Tace  GC  Ultra  equipped  with  a  FID  detector  and  a  CP-­‐WAX-­‐52CB  column  (60  m,  50   °C-­‐

180   °C,   5   min   iso,   12   °C/min,   He).   GCs   were   measured   in   dichlormethan   wit   tetradecan   as   internal  

standard.   Signals   were   assigned   via   NMR,   GC-­‐MS   and   pure   substance   calibration.   NMR   spectra   were  

recorded  on  a  Bruker  AV  400  or  a  Bruker  DPX  300  spectrometer  at  400  MHz  for  1H  and  100  MHz  for  13C,  

respectively   at   300   MHz   for   1H   and   75   MHz   for   13C.   Chemical   shifts   are   reported   relative   to  

Tetramethylsilan  and  solvent  residual  protons  or  carbon  signals  as  internal  reference. 1.  Hydrogenation/Dehydration  reaction A.  Hydrogenation/dehydration  of  4-­‐(2-­‐tetrahydrofuryl)-­‐2-­‐butanol  (THFA) Ru/C  and  Ru/Alox  were  activated  by  hydrotreatment  at  80  °C  and  100  bar  H2-­‐pressure  for  10  h  prior  use.  

In  a  typical  experiment  Ru/C,  Ru/alox  or  IL-­‐stabilisied  nanoparticles  (0.016  mmol  Ru)  were  placed  in  a  10  

mL   stainless-­‐steel   high-­‐pressure   reactor   with   a   glass   inlet.   THFA   (225.7   mg,   1.565   mmol),   2.9   mL  

[EMIM][NTf2]   and,   in   case   of   Ru/C   and   Ru/alox,   the   acidic   additive   (0.114   mmol)   were   added   to   the  

catalyst  and  the  reactor  was  pressurized  to  120  bar  with  hydrogen.  The  reaction  mixture  was  stirred  for  

15   h   at   150   °C.   The   reactor  was   cooled   to   ambient   temperature   and  was   carefully   vented   by   using   a  

cooling   trap  to  retain  any  volatile  organic  products.  For  GC  and  NMR  analysis  the  reaction  mixture  was  

extracted  with   pentane   (3x   20  mL)   and   the  pentane   phase   and   the  products   in   the   cooling   trap  were  

combined.   After   drying   with   MgSO4   and   evaporation   of   pentane   the   products   were   obtained   as   a  

colorless  solution.  For  some  selected  experiments  mass  balance  was  calculated  via  GC  with  tetradecan  as  

2

internal  standard  and  the  weight  of  the  isolated  product  mixture.  The  values  of  the  mass  balance  were  

85-­‐95  %. B.  Hydrogenation/dehydration  of  furfuralacetone  (FFA) Ru/C  wasactivated  by  hydrotreatment  at  80   °C  and  100  bar  H2-­‐pressure   for  10  h  prior  use.   In  a   typical  

experiment   Ru/C   (0.016  mmol   Ru)   was   placed   in   a   10  mL   stainless-­‐steel   high-­‐pressure   reactor  with   a  

glass  inlet.  After  addition  of  FFA  (213.1  g,  1.565  mmol)  the  reaction  mixture  was  stirred  at  120°C  and  120  

bar  hydrogen  pressure  for  2h.  The  reactor  was  cooled  to  ambient  temperature  and  was  carefully  vented.  

After  addition  of  2.9  mL  [EMIM][NTf2]  and  the  acidic  additive  (0.114  mmol)  the  reactor  was  pressurized  

again  and  the  reaction  mixture  was  stirred  at  150  °C  and  120  bar  H2  pressure  for  60  h.  The  reactor  was  

cooled  to  ambient  temperature  and  carefully  vented  by  using  a  cooling  trap  to  retain  any  volatile  organic  

products.  For  GC  and  NMR  analysis  the  reaction  mixture  was  extracted  with  pentane  (3x  20  mL)  and  the  

pentane   phase   and   the   products   in   the   cooling   trap   were   combined.   After   drying   with   MgSO4   and  

evaporation   of   pentane   the   products   were   obtained   as   a   colorless   solution.   For   some   selected  

experiments  mass  balance  was  calculated  via  GC  with  tetradecan  as  internal  standard  and  the  weight  of  

the  isolated  product  mixture.  The  values  of  the  mass  balance  were  85-­‐95  %. C.  Hydrogenation/dehydration  of  furfural  (FF) Furfural  was  distilled  and  stored  under  argon  at  -­‐20  °C.  Ru/C  was  activated  by  hydrotreatment  at  80  °C  

and  100  bar  H2-­‐pressure  for  10  h  prior  use. Furfural  (150.4  g,  1.565  mmol)  and  acetone  (1.15  mL,  15.652  mmol,  10  eq)  were  placed  in  a  glas  inlet.  To  

start   the   reaction  50  �L  of  0.1  M  NaOH  were  slowly  added   to   the  solution,  which   immediately   turned  

yellow.  After  stirring  the  reaction  mixture  for  16  h  at  room  temperature  50  �L  of  0.1  M  HCl  were  added  

and  the  excess  of  acetone  was  evaporated.  The  glas  inlet  was  transferred  to  a  high-­‐pressure  reactor  and  

after  addition  of  Ru/C  (0.016  mmol)  the  hydrogenation  reaction  was  started  by  pressurizing  the  reactor  

with  120  bar  H2.  For  the  following  reaction  and  work-­‐up  procedure  see  B.

3

2.  GC  chromatograms THFA,  isomers  of  THFA  and  self-­‐etherification  products   (Ru@[BSO3BIM][NTf2]+THFA,  120  °C,  120  bar  H2,  15  h)

4

2-­‐Butyltetrahydrofuran  (BTHF),  1-­‐octanol  (1-­‐OL)  and  dioctylether  (DOE) (Ru@[BSO3BIM][NTf2]+THFA+[EMIM][NTf2],  120  °C,  150  bar  H2,  15  h)

5

Furfural  and  Furfuralaceton (Furfural+Aceton+NaOH,  20  °C,  18  h)

6

2.  Screening  of  selected  commercially  available  Ru/C[a| No. Catalyst Conv.[%] Selectivity  [%]

BTHF 1-­‐OL DOE Self-­‐etherification Others[b] C8-­‐OL 1 Abcr  (5  wt%  Ru,  dry) 81.4 52.5 -­‐ -­‐ 37.7 9.8 -­‐ 2 Abcr  (5  wt%,  reduced) 84.0 79.5 3.5 -­‐ 10.8 6.2 3.5 3 Abcr  (5  wt%,  12-­‐25  Å) ≥99 56.6 27.1 9.1 -­‐ 7.2 36.2 4 Strem  (Escat™)[c] ≥99 46.6 31.9 16.2 -­‐ 5.3 48.1

[a]:  150  °C,  120  bar  H2,  15  h,  0.016  mmol  Ru,  1.565  mmol  THFA,  0.114  mmol  acidic  additive,  2.9  mL  [EMIM][NTf2]  [b]:  others  are  

mainly  2-­‐propyltetrahydropyran;  [c]:  dried  at  80  °C  under  reduced  pressure  prior  use. 3.  Comparison  of  catalytic  activity  between  two  different  batches  of  Ru/C  (abcr,  5  wt%,  12-­‐25  

Å,  batch  A  and  B)

No. substrate steps Selectivity  [%]

BTHF 1-­‐OL DOE Others[b] C8-­‐OL 1A THFA 1[a] 56.6 27.1 9.1 7.2 36.2 1B THFA 1[a] 47.8 20.5 19.4 12.3 39.4 2A FFA 2[b] 2.6 48.8 44.2 4.4 93.0 2B   FFA 2[b]   14.6   42.8   35.4   7.2   78.2  

[a]:  0.016  mmol  Ru,  1.565  mmol  THFA,  0.114  mmol  [BSO3BIM][NTf2],  2.9  mL  [EMIM][NTf2],  150  °C,  120  bar  H2,  15  

h  [b]:  1st  step:  0.016  mmol  Ru,  1.565  mmol  FFA,  120  °C,  120  bar  H2,  2  h;  2nd  step:  0.114  mmol  [BSO3BIM][NTf2],  

2.9  mL  [EMIM][NTf2],  150  °C,  120  bar  H2,  60  h;  [c]:  others  are  mainly  2-­‐propyltetrahydropyran. 4.  Reproducibility  of  the  dehydration/hydrogenation  step  of  THFA  with  Ru@[BSO3BIM][NTf2]

[a]    

Batch  No. conv.  [%] Selectivity  [%]

BTHF 1-­‐OL DOE Others[b] C8-­‐OL 1 ≥99 26.9 35.9 30.2 7.0 66.1 2 ≥99 21.8 44.9 26.1 7.2 71.0 3 ≥99 19.1 42.5 28.6 9.8 71.1

[a]:  0.016  mmol  Ru,  1.565  mmol  THFA,  0.114  mmol  [BSO3BIM][NTf2],  2.9  mL  [EMIM][NTf2],  150  °C,  

120  bar  H2,  15  h  [c]:  others  are  mainly  2-­‐propyltetrahydropyran

7

5.  Synthesis  of  Ru@IL[1]  

 

Ruthenium   nanoparticles   were   prepared   by   chemical   reduction   of   bis(methylallyl)(1,5-­‐

cyclooctadiene)ruthenium(II)  with   hydrogen   in   presence  of   an   ionic   liquid.   In   a   typical   experiment   the  

precursor  (5.0  mg,  0.016  mmol)  was  dispersed  in  the  ionic   liquid  (0.114  mmol)  and  the  suspension  was  

placed  in  a  10  mL  stainless-­‐steel  high-­‐pressure  reactor  with  a  glass  inlet.  After  pressurising  with  hydrogen  

to   60   bar,   the   reaction   mixture   was   stirred   for   2   h   at   60   °C.   The   reactor   was   cooled   to   ambient  

temperature  and  was  carefully  vented.  A  dark  brown  solution  was  obtained,  which  was  used  directly   in  

the  hydrogenation/dehydration  reaction.Size  and  size  distribution  of  the  nanoparticles  were  analysed  via  

Transmission   Electron   Microscopy   using   a   Hitachi-­‐HF-­‐200.   The   samples   were   prepared   by   dilution   of  

Ru@IL  with  acetone  and  deposition  on  a  carbon  coated  copper  grid.

Figure  1.  Ru@[BSO3BIM][NTf2]:  TEM  image  (left  side)  and  size  distribution  (right  side). 6.  Synthesis  of  ionic  liquids  

 1-­‐Butyl-­‐3-­‐(3-­‐carboxypropyl)-­‐imidazolium  bis(trifluoromethylsulfonyl)imid  [BCO2BIM][NTf2]

[2]  

4-­‐chlorobutanoic  acid  (2.52  g,  0.02  mol)  was  slowly  added  to  1-­‐butylimidazol  (2.55  g,  0.02  mol)  and  the  

reaction  mixture  was  stirred  for  6  h  at  120  °C.  A  viscous  solution  was  formed,  which  was  diluted  with  10  

mL  MilliQ   H2O.   After   addition   of   Lithiumbis(trifluoromethylsulfonyl)imid   (5.86   g,   0.02   mmol)   in   10   mL  

MilliQ   H2O   the   mixture  was   stirred   for   another   10   h   at   room   temperature.   Two   layers   were   formed,  

which   were   separated   and   the   aqueous   layer   was   extracted   with   dichlormethan   (3x15   mL).   The  

combined  organic  layers  were  washed  with  MilliQ  H2O.  Dichlormethan  was  evaporated  and  the  resulting  

ionic  liquid  was  dried  for  10  hours  at  50  °C  under  reduced  pressure.  1H-­‐NMR  (400  MHz,  CDCl3):  δ 1.27  (t,  3H,  J3=7.4  Hz,  CH3),  1.68  (tq,  2H,  J3=7.4  Hz,  CH2CH3),  2.19  (tt,  J3=7.4  

Hz,  CH2),  2.61  (tt,  2H,  J3=7.4  Hz,  CH2),  2.84  (t,  2H,  J3=8.1  Hz,  CH2COOH),  4.53  (t,  2H,  J3=7.4  Hz,  NCH2),  4.70  

(t,  2H,  J3=7.4  Hz,  NCH2),  7.67  (s,  1H,  NCHCHN),  7.71  (s,  1H,  NCHCHN),  8.86  (s,  1H,  NCHN),  11.21  (s,  1H,  

COOH)  ppm.  

8

13C-­‐NMR  (100  MHz,  CDCl3):  δ 13.3  (s,  CH3),  19.4  (s,  CH2),  22.2  (s,  CH2),  28.0  (s,  CH2),  32.0  (s,  CH2COOH),  

49.8  (s,  NCH2),  69.0  (s,  NCH2),  119.8  (q,  2C,  CF3),  120.  6  (s,  NCHCHN),  121.8(s,  NCHCHN),  134.2  (s,  NCHN),  

178.8  (s,  COOH)  ppm.  

 

1-­‐Butyl-­‐3-­‐(4-­‐sulfobutyl)-­‐imidazolium  bis(trifluoromethylsulfonyl)imid  [BSO3BIM][NTf2][3]  

In   a   Schlenk   roundflask   n-­‐butylimidazol   (10.0   g,   0.08   mol)   was   diluted   with   10   mL   dry   and   degassed  

toluene.  1,4-­‐butansulton  (1.47  g,  0.08  mol)  and  15  mL  toluene  were  added  and  the  mixture  was  stirred  

at   50   °C   for   24   h.   The   colourless   solution   turned   yellow   and   a   white   precipitate   was   formed.   The  

precipitate  was  filtrated  of   the  solution.  After  washing  with  toluene  and  acetone,  the  white  solid  4-­‐(N-­‐

butylimidazolium)butane-­‐1-­‐sulfonat   was   dried   under   reduced   pressure.   The   filtrate   was   stirred   for  

another   24   h   and   the   formed   precipitate   was   separated   by   filtration   as   before.   This   procedure   was  

repeated  until  the  conversion  of  n-­‐butylimidazol  was  complete.  4-­‐(N-­‐butylimidazolium)butane-­‐1-­‐sulfonat  

(3.81   g,   0.01   mol)   was   dissolved   in   6   mL   MilliQ   H2O.   An   aqueous   solution   of  

bis(trifluoromethane)sulfonimide  (80  %,  3.79  mL,  0.01  mmol)  was  added  and  the  solution  was  stirred  for  

2  h  at   room  temperature.  After   evaporation  of  water   the  viscous   ionic   liquid  was  dried  under   reduced  

pressure.  1H-­‐NMR  (400  MHz,  D2O):  δ0.77  (t,  3H,  J3=7.5  Hz,  CH3),  1.17  (tq,  2H,  J3=7.5  Hz,  CH2),  1.64  (tt,  2H,  J3=7.5  Hz,  

CH2),  1.70  (tt,  2H,   J3=7.5  Hz,  CH2),  1.90  (tt,  2H,   J3=7.5  Hz,  CH2),  2.81  (t,   J3=7.5  Hz,  CH2SO3H),  4.04  (t,  2H,  

J3=7.3  Hz,  CH2N),  4.11  (t,  2H,  J3=7.3  Hz,  CH2N),  7.34  (s,  1H,  NCHCHN),  7.38  (s,  1H,  NCHCHN),  8.62  (s,  1H,  

NCHN)  ppm.  13C-­‐NMR  (100  MHz,  D2O):  δ12.5  (s,  CH3),  18.7  (s,  CH2),  21.0  (s,  CH2),  28.2  (s,  CH2),  31.2  (s,  CH2),  48.9  (s,  

CH2N),   49.3   (s,   CH2SO3H),   50.1   (s,   CH2N),   119.3   (q,   J1=325   Hz),   122.3   (s,   NCHCHN),   122.4   (s,  

NCHCHN),135.0  (NCHN)  ppm.  

 

1-­‐Butyl-­‐3-­‐(4-­‐sulfobutyl)-­‐imidazolium  trifluoromethanesulfonate  [BSO3BIM][OTf]  

In   a   Schlenk   roundflask   n-­‐butylimidazol   (10.0   g,   0.08   mol)   was   diluted   with   10   mL   dry   and   degassed  

toluene.  1,4-­‐butansulton  (1.47  g,  0.08  mol)  and  15  mL  toluene  were  added  and  the  mixture  was  stirred  

at   50   °C   for   24   h.   The   colourless   solution   turned   yellow   and   a   white   precipitate   was   formed.   The  

precipitate  was  filtrated  of   the  solution.  After  washing  with  toluene  and  acetone,  the  white  solid  4-­‐(N-­‐

butylimidazolium)butane-­‐1-­‐sulfonat   was   dried   under   reduced   pressure.   The   filtrate   was   stirred   for  

another   24   h   and   the   formed   precipitate   was   separated   by   filtration   as   before.   This   procedure   was  

repeated  until  the  conversion  of  n-­‐butylimidazol  was  complete.  4-­‐(N-­‐butylimidazolium)butane-­‐1-­‐sulfonat  

(3.81   g,   0.01   mol)   was   dissolved   in   6   mL   MilliQ   H2O.   Trifluoromethanesulfonic   acid   (0.01   mmol)   was  

added  and  the  solution  was  stirred  for  2  h  at  room  temperature.  After  evaporation  of  water  the  viscous  

ionic  liquid  was  dried  under  reduced  pressure.  1H-­‐NMR  (400  MHz,  D2O):  δ 0.76   (t,  3H,   J3=7.5  Hz,  CH3),  1.16   (tq,  2H,   J3=7.5  Hz,  CH2),  1.60   (m,  2H,  CH2),  

1.70  (tt,  2H,  J3=7.5  Hz,  CH2),  1.88  (tt,  2H,  J3=7.5  Hz,  CH2),  2.79  (t,  J3=7.5  Hz,  CH2SO3H),  4.04  (t,  2H,  J3=7.1  

Hz,  CH2N),  4.10  (t,  2H,  J3=7.1  Hz,  CH2N),  7.36  (m,  2H,  NCHCHN),  8.65  (s,  1H,  NCHN)  ppm.  

9

13C-­‐NMR  (100  MHz,  D2O):  δ12.5  (s,  CH3),  18.7  (s,  CH2),  20.9  (s,  CH2),  28.0  (s,  CH2),  31.1  (s,  CH2),  48.9  (s,  

CH2N),   49.3   (s,  CH2SO3H),   50.0   (s,  CH2N),   119.6   (q,   J1=319  Hz),   122.4   (s,   NCHCHN),   122.4   (s,  NCHCHN),  

135.1  (NCHN)  ppm.  

N,N,N-­‐tributly-­‐N-­‐(4-­‐sulfobutyl)ammonium  bis(trifluoromethylsulfonyl)imide  [N444BSO3][NTf2]

[4] Tributlyamine  (3.57  g,  0.02  mol)  and  1,4-­‐Butansulton   (2.66  g,  0.02  mol)  were  stirred  for  24  h  at  130   °C  

forming  a  pale  yellow  viscous  liquid.  After  addition  of  20  mL  dry  and  degassed  ethyl  acetate  the  solution  

was  stirred  for  another  2h  at  90  °C.  The  reaction  was  cooled  down  to  0  °C  and  a  white  solid  precipitated.  

Filtration  of  the  precipitate,  washing  with  ethyl  acetate  and  drying  under  reduced  pressure  gave  access  

to  4-­‐(N,N,N-­‐tributylammonium)butane-­‐1-­‐sulfonat  as  a  white  powder. In  a  Schlenk  roundflask  4-­‐(N,N,N-­‐tributylammonium)butane-­‐1-­‐sulfonat  (2.88  g,  0.01  mol)  was  dissolved  

in  10  mL  MilliQ  H2O.  An  aqueous  solution  of  bis(trifluoromethan)sulfonimid  (80  %,  2.30  mL,  0.01  mol))  

was   added   and   the   solution  was   stirred   for   2   h   at   room   temperature.   After   evaporation  of   water   the  

viscous  ionic  liquid  was  dried  under  reduced  pressure.  1H-­‐NMR  (400  MHz,  DMSO):  δ0.93  (t,  9H,  J3=7.2  Hz,  CH3),  1.30  (tq,  6H,  J3=7.2  Hz,  CH2),  1.58  (m,  8H,  CH2),  

1.72  (m,  2H,  CH2),  2.57  (m,  2H,  CH2SO3H),  3.02  (m,  2H,  CH2N),  4.10  (m,  6H,  CH2N)  ppm.  13C-­‐NMR  (100  MHz,  D2O):  δ 13.4  (s,  3C,  CH3),  19.2  (s,3C,  CH2),  19.3  (s,  1C,  CH2),  23.0  (s,  3C,  CH2),  25.0  (s,  

1C,  CH2),  50.1  (s,  3C,  CH2N),  51.7  (s,  1C,  CH2N),  57.5  (s,  CH2SO3H)  ppm. 7.  Analytic  Data 4-­‐(2-­‐tetrahydrofuryl)-­‐2-­‐butanol  (THFA)

1H-­‐NMR  (400  MHz,  CDCl3):  δ 1.08  (dd,  3H,  J3=  x  Hz,  CH3),  1.32-­‐1.58  (m,  5H,  2  x  CH2,  tetrahydrofurylring:  

CH(4)),  1.71-­‐1.92  (m,  3H,  tetrahydrofruylring:  CH2(3),CH(4)),  3.60-­‐3.65  (m,  1H,  CHOH),  3.65-­‐3.79  (m,  3H,  

tetrahydrofurylring:  CH(2)  CH2(5))  ppm. 13C-­‐NMR   (100   MHz,   CDCl3):   δ 23.3   (d,   CH3),   25.6   (d,   tetrahydrofurylring:   CH2),   31.4   (d,   CH2),   31.9   (d,  

tetrahydrofurylring:   CH2),   36.0   (d,   CH2),   67.5   (d,   CHOH),   67.6   (s,   tetrahydrofurylring:   CH2),   79.5   (d,  

tetrahydrofurylring:  CH)  ppm. MS  (CI):  m/z  145  ([M++H],44),  143  (11),  127  (81),  109  (46),  71  (100),  67  (10). Correction  factor  GC:  1.81

10

2-­‐butyltetrahydrofuran  (BTHF)

1H-­‐NMR  (400  MHz,  CDCl3):  δ 0.80  (t,  3H,  J3=  6.8  Hz,  CH3),  1.28  (m,  7H,  3  x  CH2,  tetrahydrofurylring:  CH(4)),  

1.80  (m,  3H,  tetradhydrofurylring:  CH2(3),  CH(4)),  3.66  (m,  3H,  tetrahydrofurylring:  CH(2),  CH2(5))  ppm. 13C-­‐NMR  (100  MHz,  CDCl3):  δ 13.9  (s,  CH3)  ,  22.8  (s,  CH2),  25.7  (s,  CH2),  28.5  (s,  CH2),  31.3  (s,  CH2),  35.4  (s,  

CH2),  67.5  (s,  tetrahydrofurylring:  CH2),  79.4  (s,  tetrahydrofurylring:  CH)  ppm. MS  (EI):  m/z  71  (100),  70  (10),  43  (25),  42  (17),  41  (37),  39  (17). Correction  factor  GC:  1.28 1-­‐octanol  (1-­‐OL)

1H-­‐NMR   (400  MHz,   CDCl3):  δ 0.87   (t,   3H,   J3=  6.7  Hz,   CH3),   1.26   (m,   10H,   CH2),   1.55   (tt,   2H,   J3=  6.9  Hz,  

CH2CH2OH),  3.39  (t,  2H,  J3=  6.7  Hz,  CH2OH)  ppm. 13C-­‐NMR  (100  MHz,  CDCl3):  δ 14.1  (s,  CH3),  22.6  (s,  CH2),  25.7  (s,  CH2),  29.3  (s,  CH2),  29.4  (s,  CH2),  31.8  (s,  

CH2),  32.8  (s,  CH2),  63.0  (s,  CH2OH)  ppm. MS  (EI):  m/Z  84  (10),  83  (10),  70  (30),  69  (25),  56  (45),  55  (50),  43  (70),  43  (55),  41  (100),  39  (62).  Correction  factor  GC:  1.22 1,1-­‐dioctylether  (DOE)

1H-­‐NMR  (400  MHz,  CDCl3):  δ 0.88  (m,  6H,  CH3),  1.27  (m,  20H,  CH2),  1.56  (tt,  4H,  J3=  6.8  Hz,  CH2CH2OH),  

3.39  (t,  4H,  J3=  6.8  Hz,  CH2OH)  ppm. 13C-­‐NMR  (100  MHz,  CDCl3):  δ14.2  (s,  CH3),  22.8  (s,  CH2),  26.4  (s,  CH2),  29.4  (s,  CH2),  29.6  (s,  CH2),  29.9  (s,  

CH2),  32.0  (s,  CH2),  71.1  (s,  CH2OH)  ppm. MS  (EI):  m/Z  84  (10),  83  (10),  71  (37),  69  (25),  57  (73),  55  (35),  43  (100),  41  (85). Correction  factor  GC:  1.19

11

 

 

[1]   J.  Julis,  M.  Hölscher,  W.  Leitner,  Green  Chemistry  2010,  12,  1634-­‐1639. [2]   R.  Abu-­‐Reziq,  D.  Wang,  M.  Post,  H.  Alper,  Advanced  Synthesis  &  Catalysis  2007,  349,  2145-­‐2150. [3]   M.  A.  Liauw,  S.  Winterle,  Chemie  Ingenieur  Technik  2010,  82,  1211-­‐1214. [4]   Y.  Gu,  C.  Ogawa,  S.  Kobayashi,  Chemistry  Letters  2006,  35,  1176-­‐1177.


Recommended