• No results found

University of Groningen Sustainable pathways to chemicals and fuels from lignocellulose via catalytic cleavage and coupling reactions Sun, Zhuohua

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Sustainable pathways to chemicals and fuels from lignocellulose via catalytic cleavage and coupling reactions Sun, Zhuohua"

Copied!
23
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Sustainable pathways to chemicals and fuels from lignocellulose via catalytic cleavage and coupling reactions

Sun, Zhuohua

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2018

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Sun, Z. (2018). Sustainable pathways to chemicals and fuels from lignocellulose via catalytic cleavage and coupling reactions. Rijksuniversiteit Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

Chapter 6

Catalytic Conversion of 1,Diaminobenzenes and

2-Nitroanilines to Benzimidazoles in Supercritical Methanol

In this chapter the possibility of using supercritical methanol as carbon source for the construction of the benzimidazoles and N-methylbenzimidazole moiety is presented. The starting materials, 1,2-diaminobenzenes were simply heated in supercritical methanol over copper-doped porous metal oxides (Cu-PMO). These catalysts were derived from synthetic hydrotalcites that only contain earth-abundant starting materials. The carbon equivalents needed for the construction of the benzimidazole core originated from the solvent itself, which is known to undergo reforming to hydrogen and carbon monoxide through the formation of formaldehyde intermediate. A variety of 1,2-diaminobenzenes were converted to the corresponding mixtures of benzimidazoles and N-methylated analogues in good yields. Interestingly, the more challenging, but readily available 2-nitroanilines, which require an additional reduction step prior to cyclization, could also be successfully converted to benzimidazoles in high selectivity. Furthermore, various other alcohols were applied besides methanol, to obtain 2-alkyl- and 1,2-dialkylbenzimidazoles. Preliminary mechanistic insights into the origins of N-alkylation as well as the reactivity of the nitro derivatives are discussed.

This chapter has been published as: Z. Sun, G. Bottari & K. Barta, Green Chem., 2015, 17, 5172–5181.

(3)

118

6.1 Introduction

Benzimidazole and its derivatives are frequently targeted in medicinal chemistry research due to their prominent biological activity.1,2 Several pharmaceutically active structures, containing the benzimidazole core, have been found to exhibit significant activity against viruses such as HIV,3 herpes (HSV-1)4 or influenza5. Furthermore, Esomeprazole (Figure 6.1) has been one of the top selling drugs against peptic ulcers and gastro esophageal reflux in the last decade, while Telmisartan (Micardis), containing a 1,2-disubstituted benzimidazole, is a popular antihypertensive.6

Figure 6.1 Commercial drugs containing a benzimidazole core unit.

Scheme 6.1 Comparison of existing methods and methods developed in this paper for the

(4)

119 methods. b) Catalytic acceptorless dehydrogenative coupling. c) The use of copper doped porous metal oxides for the transfer of useful carbon equivalents from methanol.

Classical methods for benzimidazole synthesis involve the coupling of 1,2-diaminobenzene with acids, acid chlorides or anhydrides (Scheme 6.1a), usually requiring strongly acidic conditions.1,7For example, a standard procedure for the synthesis of benzimidazole consists of heating 1,2-diaminobenzene in concentrated formic acid.8 Many of the classical methods suffer from low atom economy and formation of stoichiometric amounts of waste due to substrate leaving groups or additives.7 Recent reports focused on improving reaction conditions using aldehydes as substrates and an appropriate oxidizing agent. For example, an inorganic iodine catalyst in the presence of hydrogen peroxide afforded high benzimidazole yields from 1,2-diaminobenzenes and aldehydes at room temperature.9 Microwave-assisted methodologies have also reduced reaction times.10–13 In all the strategies detailed above, the availability of the coupling partner to 1,2-diaminobenzene may present further limitations.

Alcohols are readily available starting materials that will become accessible by fermentation14 or catalytic conversion of renewable lignocellulose.15 Therefore the development of catalytic methodologies for the synthesis of benzimidazoles directly from alcohols is desired. Recent approaches involve the classical or photochemical oxidation or dehydrogenation of the alcohol to obtain the more reactive aldehyde, which subsequently reacts with the diamine to form the desired benzimidazole ring. For example, one-pot benzimidazole synthesis was carried out using bifunctional supported gold and palladium catalysts under oxygen pressure.16 The use of an iron phthalocyanine catalyst was also reported, which afforded a variety of 2-substituted benzimidazoles from 1,2-diaminobenzenes.17 Elegant approaches using TiO218 and Pt-TiO219,20 allowed for the

photocatalytic (λ > 300 nm) coupling of o-arylenediamines with a variety of alcohols. This catalytic systems could be also successfully applied to 2-nitroaniline substrates that are in situ reduced to 1,2-diaminobenzenes.18,20,21

Acceptorless dehydrogenative coupling (Schemes 6.1b and 6.1c)22,23 has emerged as an attractive strategy for a variety of processes starting from alcohols during which only hydrogen gas and innocuous water molecules are eliminated. Through acceptorless dehydrogenative condensation, important N-heterocycles (cyclic amines, lactames, pyrroles, benzimidazoles, etc.) can be accessed in a clean and highly atom-economic manner.24–26 In 2014, Kempe and coworkers reported a robust catalytic system for the synthesis of benzimidazoles and quinoxalines from aromatic 1,2-diamines using a wide range of alcohols and diols.27 The reported iridium catalyst capable of acceptorless dehydrogenation, is highly active at very low catalyst loadings (0.04–1.4%), and requires the addition of base and diglyme solvent (Scheme 6.1b). Homogeneous ruthenium catalysts are also known to afford a number of benzimidazoles, but only in the presence of an olefin sacrificial acceptor and catalytic amounts of acid.28

(5)

120

The use of copper-doped porous metal oxides (Cu-PMO) for the depolymerization of organosolv lignin29 and lignocellulosic biomass30 in supercritical methanol was introduced by Ford and co-workers. In this unique catalyst system, the reductive equivalents needed for depolymerization and various hydrogenation steps were transferred from the solvent itself, which in part underwent reforming to syngas.31

We have devised a related concept for the synthesis of benzimidazoles using PMO in methanol, whereby in situ formed formaldehyde,31 the primary product of methanol dehydrogenation would serve as source of useful carbon equivalents for the construction of the benzimidazole core upon reaction with 1,2-diaminobenzenes (Scheme 6.1c). Moreover, the hydrogen produced under these reaction conditions would allow for accomplishing an additional reduction step starting from more readily available 2-nitroanilines with no other reagents needed.

In this chapter, I will report the development of a new, additive-free method for the preparation of benzimidazoles from various 1,2-diaminobenzenes as well as 2-nitroanilines simply by heating these starting materials in scMeOH that acts both as solvent and reactant, in presence of Cu-PMO and related catalysts. The methodology is extended to the use of different neat n-alcohols, which offer a route to 1,2-dialkylbenzimidazoles in good yields.

6.2 Results and discussion

6.2.1 Benzimidazole formation from 1,2-diaminobenzenes

The investigations were starting by identifying the products formed upon heating 1,2-diaminobenzenes in supercritical methanol in the presence of a Cu-PMO, which is the same catalyst used in Chapter 2 and the advantage of this catalyst has been discussed in Chapter 1.4.

Scheme 6.2 Formation of benzimidazole and N-methylbenzimidazole from

(6)

121

Table 6.1 Optimization of reaction conditions for the conversion of 1,2-diaminobenzene to

benzimidazole derivatives in methanol using Cu-PMO catalyst. a

Entry T (oC) Time (h) Conversion (%)b Selectivity %

b 1a (%) 1b (%) 1c (%) 1 280 2 100 43 57 0 2 250 2 69 41 21 7 3 250 3 80 46 27 7 4 250 6 100 43 56 1 5 220 2 16 13 1 2 6 150 2 0 0 0 0

a. Reaction conditions: 1,2-diaminobenzene (1 mmol), Cu-PMO (50 mg), methanol (3 mL). b. Determined by GC analysis.

The reaction using 1 mmol 1,2-diaminobenzene and 50 mg Cu-PMO catalyst at 280 °C for 2 h (Table 6.1, Entry 1) afforded full substrate conversion. The two main reaction products were identified as benzimidazole 1a and N-methylbenzimidazole 1b, formed in almost equimolar ratio (Scheme 6.2). GC-MS and GC-FID analysis using authentic standards confirmed the formation of these desired products, and subsequently 1a and 1b were isolated by column chromatography and unambiguously characterized.

Small amounts of N-methylbenzene-1,2-diamine 1c were also detected in the product mixture (see Mechanistic Considerations Chapter 6.2.5). Decreasing the reaction temperature to 250 °C, conversion was 69% and a 2/1 ratio between 1a and 1b was found. When the reaction time was prolonged to 3 h and 6 h higher substrate conversion of 80% and 100% respectively and lower 1a/1b ratio (1.7/1 and 0.8/1) were seen. Little or no product was formed at 220 °C and 150 °C within 2 hours (Table 6.1, Entries 5 and 6).

Table 6.2 Product distribution in the benzimidazole synthesis from 1,2-benzenediamine at

different substrate loading and constant catalyst amount.a

Entry Substrate (mmol) Conversion (%)b Selectivity %b 1a (%) 1b (%) 1c (%) 1 0.6 100 58 42 0 2 0.8 95 53 40 2 3 1.0 79 46 27 6 4 1.5 69 33 23 13 5 2.0 53 27 14 13

a. Reaction conditions: Cu-PMO (50 mg), methanol (3 mL), 250 oC, 3h. b. Determined by GC analysis.

(7)

122

Figure 6.2 GC selectivities of 1a, 1b and 1c at different substrate to catalyst ratios (the

remainder is the unconverted 1,2-diaminobenzene). Reaction conditions: substrate (amount indicated in the bar chart), Cu-PMO (50mg), methanol (3ml), 250oC, 3h.

Next, the catalyst to substrate ratio was varied by gradually increasing substrate amount from 0.6 to 2 mmol at constant catalyst loading (50 mg) and methanol volume (3 mL). These results are summarized in Table 6.2 and Figure 6.2. At low substrate to catalyst ratio (ca. 1.3: 1), full conversion was achieved in 3 h and comparable amounts of 1a and 1b were observed. At higher concentration of 1,2-diaminobenzene, the substrate was partially converted (e.g. 53% conversion when using 2 mmol substrate, 7 mol% Cu to substrate, Table 6.2, Entry 5). In all cases, 1a was the major product and the 1a/1b ratio ranged between 1.3 and 1.9. Thus it can be concluded that lower temperatures and higher substrate concentration favor benzimidazole 1a, over its methylated analogue 1b. N-Methyl-o-phenylenediamine could also be detected in the product mixture up to 13% yield and its formation was favored at high substrate concentration (see Mechanistic considerations).

Table 6.3 Screening of different PMO compositions for the conversion of

1,2-diaminobenzene to benzimidazole derivatives.a

Catalyst Conversion b (%) Selectivity %b 1a (%) 1b (%) 1c (%) Not identified Cu-PMO 80 46 27 7 0 Cu-Ni-PMO 62 45 9 5 1 Cu-Zn-PMO 92 50 35 3 4

a. Reaction conditions: 1,2-diaminobenzene (1 mmol), catalyst (50 mg) , methanol (3 mL), 250 oC, 3h. b. Determined by GC analysis.

The advantage of hydrotalcites is their modular synthesis as subtle changes in catalyst composition are expected to have an influence on catalytic activity.17,32 As shown in Table 6.3, two new catalyst compositions were prepared, additionally introducing 5% of Ni2+ or Zn2+ metal dopants, which did have an influence on both catalyst activity and product

(8)

123 selectivity. With Cu–Ni-PMO 62% conversion was seen (Table 6.3, Entry 2) however 1a selectivity improved compared to the reactions with Cu-PMO. An opposite effect was found with Cu–Zn-PMO, which was the most active among the catalysts tested, and afforded the highest selectivity to benzimidazole 1a (50% at 92% substrate conversion, Table 6.3, Entry 3), but a lower 1a/1b ratio (1.4, compared to 1.7 with Cu-PMO and 5.0 with Cu–Ni-PMO). The amount of unidentified products when the zinc-doped composition was used was 4%, which is more than in the other cases where these amounts were practically insignificant.

6.2.2 Formation of benzimidazoles directly from 2-nitroaniline

Compared to phenylenediamines, 2-nitroanilines are more readily available substrates. Generally, 2-nitroanilines are reduced to 1,2-diaminobenzenes with zerovalent metals (Fe, Sn, Zn) and diluted mineral acids.33 Direct catalytic methods that allow for conversion of 2-nitroanilines to benzimidazoles involve the reduction of the nitro functionality first, followed by cyclization. A number of such systems, mainly using aldehydes and acids as coupling partners are known.34–37 However, the one-pot direct coupling of alcohols with 2-nitroanilines to benzimidazoles has only been accomplished by using photocatalysts.18,20,21

Scheme 6.3 One-pot synthesis of benzimidazole derivatives from 2-nitroaniline. Table 6.4 Synthesis of benzimidazoles 1a and 1b using 2-nitroaniline substrate.a

Entry Catalyst Conversion (%)b Selectivity %

b 1a (%) 1b (%) Other products (%)c 1 Cu-PMO 38 30 4 4 2 Cu-Ni-PMO 46 41 2 3 3 Cu-Zn-PMO 100 82 11 7 4 Cu-Zn10-PMO 25 16 2 2 5 Cu-Ru-PMO 69 60 6 3 6d Cu-PMO 100 79 16 5

a. Reaction conditions: catalyst (50 mg), 2-nitroaniline (1 mmol), methanol (3 mL), 250 oC, 6 h. b. Determined by GC-FID. c. Other products mainly include 1-(methoxymethyl)-benzimidazole. d. Reaction time was 20 h.

It has been previously shown that copper nanoparticles are active in reduction of nitro-aromatics.38–40 Since during methanol reforming both Cu(0) nanoparticles and hydrogen gas are generated in our system,31 we anticipated that Cu-PMO in scMeOH will be uniquely suited for the reduction of 2-nitroanilines. The Cu-PMO catalyst is expected to (a) promote the formation of hydrogen gas via methanol reforming, (b) reduce the nitro group by the in

(9)

124

situ formed Cu(0) species and (c) provide carbon equivalents for the construction of the benzimidazole scaffold.

Indeed, the PMO catalysts showed moderate to excellent activity (38% to full conversion in 6 hours) depending on the metal dopant (see Table 6.4) in the direct conversion of 2-nitroanilines to benzimidazoles (Scheme 6.3). More interestingly, the reaction was shown highly selective towards the formation of benzimidazole 1a (see Table 6.4). The Cu-PMO catalyst afforded the lowest substrate conversion (38%, Table 6.4, Entry 1) while the nickel-doped catalyst showed a moderate 46% conversion and a 1a/1b ratio of 20/1. Especially when small amounts of Zn (1/3 with respect to Cu) were introduced, the substrate was fully converted after 6 hours with 82% selectivity to 1a (Table 6.4, Entry 3). Higher Zn loadings (Cu–Zn10-PMO, with 1/2 Zn to Cu ratio) did not improve the catalyst activity (Table 6.4, Entry 4). Surprisingly, even by introducing ruthenium noble metal in the composition the catalyst activity did not perform as good as the Cu-Zn-PMO catalyst; however, a good 69% conversion and 10/1 ratio between 1a and 1b was obtained (Table 6.4, Entry 5). For comparison, a reaction was also performed with Cu-PMO at longer reaction time (20 h, Table 6.4, Entry 6) in order to ensure full conversion and direct comparison with Cu-Zn-PMO in terms of 1a selectivity. Indeed, a 79% selectivity to 1a was obtained, close to that obtained by Cu-Zn-PMO.

In summary, all PMO catalysts prepared were suitable to carry out the conversion of 2-nitroaniline to benzimidazole in methanol. Among these, significant differences were observed: the performance of Cu-PMO and Cu-Ni-PMO was comparable but considerably lower than the Ru- and Zn-doped PMO catalysts. The same trend was observed in the conversion of 1,2-diaminobenzene, but in this case, the differences were more pronounced. In addition, Figure 6.3 shows the product formation profile during 5 hours using Cu–Zn-PMO. The conversion increased linearly during the first 3 hours (92% conversion), with no significant changes in the composition of the product mixture. Kinetic fitting of the obtained data points, revealed a rate constant of approximately k=0.5 h−1 for the benzimidazole 1a formation (Figure 6.3) and a pseudo-first order regime due to supposed excess of formaldehyde generated.

(10)

125

Figure 6.3 Kinetic profile for the 2-nitroaniline conversion to benzimidazole and

1-methylbenzimidazole. k value refers to the reaction rate constants for the formation of benzimidazole and 1-methylbenzimidazole, respectively. Reaction conditions: Cu-Zn-PMO (50 mg), 2-nitroaniline (1 mmol), methanol (3 ml), 250 oC.

6.2.3 Substrate scope

This new method could be successfully extended to a variety of 1,diaminobenzenes and 2-nitroanilines with various substituents on the aromatic ring. The reactions were carried out at 250 °C for a reaction time ensuring full substrate conversion. The main products were isolated by column chromatography. It was already discussed that 1,2-diaminobenzene affords an approximately 1 : 1 mixture of 1a and 1b. A good 78% combined isolated yield was obtained (Table 6.5, Entry 1). Using 1,2- diamino-4-methoxybenzene resulted in a similar outcome (76% combined yield, Table 6.5, Entry 2). Interestingly, a methyl group in the 4 position favored the formation of the simple benzimidazole (60%, Table 6.5, Entry 3) while the analogous t-butyl substrate favored N-methylbenzimidazole (64%, Table 6.5, Entry 4). An interesting case is offered by substrate 3,3’-diaminobenzidine, containing a diphenyl backbone and two possible reactive sites (Table 6.5, Entry 5). In this case the progress of the reaction was monitored by TLC due to the high molecular weights of the products. Three different products were detected and isolated by chromatography. The major product 5c was isolated in a good 51% yield while lower yields of two N-methylated benzimidazoles were obtained (15% of the bis-N-methylated 5a and 8% of the mono-N-methylated 5b, both present as couple of structural isomers). A diaminopyridine backbone afforded valuable imidazo-pyridine compounds, although a lower yield could be attributed to less basic NH2

(11)

126

groups (Table 6.5, Entry 6). A longer reaction time (15 h) did not afford higher substrate conversion. Electronic effects derived from substituents in the aromatic ring largely influenced the reactivity of 1,2-diaminobenzenes. More electron-rich substrates were more reactive. A deactivating effect was observed for a 4-chloro-1,2-diaminobenzene, which preferentially underwent methylation of the NH2 groups affording only little benzimidazole

yields (Table 6.5, Entry 7), moreover, dehalogenation also took place.

Several 2-nitroanilines were also employed to give higher yields of unsubstituted benzimidazoles, and most substrates were fully converted in only 3 h in presence of the more reactive Cu–Zn-PMO catalyst. When using 2-nitroaniline, 68% benzimidazole yield was achieved. In contrast to the 1,2-diaminobenzenes, alkyl substituents and electron-donating – OMe group in the aromatic ring of 2-nitroanilines did not influence to a large extent the product selectivity. Selectivity to N-methyl benzimidazoles 8b–12b ranged from 3% to 13%. Very good isolated yields of benzimidazoles 8a–12a were obtained (68–80%) and are shown in Table 6.5 (Entries 8–12). The presence of a halogen led to significant catalyst deactivation, and only small amounts of benzimidazole products (12% combined yield, Table 6.5, Entry 13) were obtained.

Table 6.5 shows that benzimidazoles unsubstituted in the 2 position can be readily prepared in methanol solvent from both 1,2-diaminobenzenes as well as 2-nitroanilines in good yields. In order to achieve variations in the 2 position, frequently related to a specific biological activity, these products may be further functionalized by C–H activation.41 Metal-catalyzed alkylation,42–44 arylation,45 acylation46 and annulation reactions47 are known to proceed at this position with excellent chemoselectivity. Nonetheless, we extended the scope of our methodology established with methanol, to the use of other n-alcohols to form 2-alkylbenzimidazole products, since supported copper catalysts48 and copper hydrotalcites are known to generally promote dehydrogenation of a variety of alcohols to the corresponding aldehydes.49 Reaction of 1,2-diaminobenzenes in ethanol gave results very similar to those observed in methanol (75% combined yield with a 1 : 1 product ratio, Table 6.6, Entry 1). The 2-nitroaniline in ethanol afforded higher yield of compound 14a (77%) and smaller fraction of 14b (16%, Table 6.6, Entry 2). When the reaction was carried out in 1-propanol, a good isolated yield of 2-ethylbenzimidazole was obtained from 1,2-diaminobenzene (66%, Table 6.6, Entry 3), and this amount was similar when 2-nitroaniline was used (62%, Table 6.5, Entry 4). 2-Alkylbenzimidazoles were isolated as main products when 1-butanol and 1-pentanol was coupled with 1,2-diaminobenzene (Table 6.6, Entries 5 and 6). The corresponding 2,3-dialkylated products were also isolated in moderate yields (24% and 45%). A reaction between 1,2-di-aminobenzene and benzyl alcohol was performed in toluene affording almost full diamine conversion, but low 2-phenyl-benzimidazole yield (26%, Table 6.6, Entry 7). Interestingly, an equal amount (26%) of benzylated starting material was isolated accounting for a competition between the hydrogenation of the imine intermediate with the cyclization process.

(12)

127

Table 6.5 Scope of methodology in the formation of benzimidazoles and

N-methylbenzimidazoles using Cu-PMO and Cu-Zn-PMO compositions.

Entrya Substrate Time (h) Conversion (%)c Selectivityc (yield)d (%) a b 1 6 100 1a, 56c (42)d 1b, 44 (36) 2 6 100 2a, 51 (42) 2b, 49 (34) 3 6 100 3a, 70 (60) 3b, 28 (21) 4 6 100 4a, 8 4b, 76 (64) 5 15 100 5a (15) 5b (8) 5c (51) 6 6 40 6a (4) 6b, (22) 7 6 100 7a,42 7b, 16 7c, 24 8 b 6 100 8a, 82 (68) 8b, 11 9 b 3 100 9a, 80 (73) 9b, 13 10b 3 100 10a, 80 (73) 10b, 12 11b 3 100 11a, 78 (61) 11b, 7 12b 6 100 12a, 90 (80) 12b, 3 13b 12 86 13a, 7 13b, 5

a. Reaction conditions: substrate (1 mmol), Cu-PMO (50 mg), methanol (3ml), 250 oC. b. Reaction conditions:

substrate (1 mmol), Cu-Zn-PMO (50 mg), methanol (3ml), 250 oC c. Determined by GC-FID. d. Isolated yields by

(13)

128

Table 6.6 Synthesis of 2-alkyl- and 1,2-dialkylbenzimidazoles by using different alcohols.

Entrya Alcohol Time (h) substrate Yield (%)c

1 Ethanol 6

1,2-diaminobenzene 14a 37

14b 38 2 Ethanol 19 2-nitroaniline 14a 77

14b 16 3 1-propanol 15 1,2-diaminobenzene 15a 66 15b 23 4 1-propanol 19 2-nitroaniline 15a 62 15b 27 5 1-butanol 15 1,2-diaminobenzene 16a 44 16b 24 6 1-pentanol 15 1,2-diaminobenzene 17a 50 17b 45 7b Benzyl alcohol 15 1,2-diaminobenzene 18a 26 18b 26

a) Reaction conditions: substrate (1 mmol), alcohol (3 ml), Cu-PMO (50 mg), 250 oC. Full substrate conversion.

b) Reaction conditions: 1,2-diaminobenzene (1 mmol), benzyl alcohol (1 mmol), Cu-PMO (50 mg), toluene (3ml), 250 oC. 80% conversion. c) Isolated yields by column chromatography.

6.2.4 Recycling tests and catalyst stability

Recycling experiments were conducted at 250 °C with 1,2-diaminobenzene and Cu-PMO for 6 h (see Figure 6.4). The catalyst was recovered at the end of each run, washed with methanol and acetone, and reused after drying. The catalyst showed excellent selectivity to benzimidazole and 1-methylbenzimidazole (>80%) for 6 cycles and after the 7th cycle still a good 75% combined selectivity was observed. The catalyst has proven considerable stability after converting more than 0.5 g 1,2-diaminobenzene in relatively harsh conditions. Interestingly, a variation in selectivity to benzimidazole 1a is visible between the first and the second cycle (45% vs. 20% selectivity) while the relative product distribution is constant for

(14)

129 the subsequent cycles. This trend is very likely related to the initial low concentration of active Cu0 species, and which are formed in situ during the first cycle from CuO present in the catalyst.31 Under comparable reaction conditions (see Table 6.7), the Cu–Zn-PMO showed almost no leaching of the metals incorporated in the catalyst structure while a small Mg and Al loss was observed in the analysis of the liquid sample after reaction with the Cu-PMO catalyst.

Figure 6.4 Catalyst recycling experiments. Reaction conditions: 1,2-diaminobenzene (1

mmol), Cu-PMO (100 mg), methanol (3 ml), 250 oC, 6 h.

Table 6.7 Leaching tests for Cu-PMO and Cu-Zn-PMO catalysts.a

Catalyst Cu (mg/L) Mg (mg/L) Al (mg/L) Zn (mg/L)

Cu-PMO <1 15 11

Cu-Zn-PMO <1 6 1 3

a. Reaction conditions: Catalysts (50 mg), 1,2-benzenediamine (1 mmol), methanol (3 mL), 250 oC, 3h.

6.2.5 Mechanistic considerations

Based on previous reports19,20,37–40,50a preliminary mechanistic description can be proposed involving methanol dehydrogenation, imine formation/cyclization and dehydrogenation. The Cu-PMO catalysts have multiple roles, catalyzing methanol reforming which leads to formaldehyde 12a,31 that likely represents a key intermediate.51 Excess of formaldehyde with 1,2-diaminobenzene has been elsewhere shown to give the expected N-methylbenzimidazole.10–13 In addition, we have independently verified the formation of benzimidazole with formaldehyde in presence of a Cu-PMO catalyst.

(15)

130

The more activated 1,2-diaminobenzene substrates may react with one or two molecules of formaldehyde (present in excess in the reaction medium) to give imine intermediates. Related imines with benzyl alcohol have been detected by Ghosh et al. at room temperature.50 However, the imines shown on Scheme 6.4 will be difficult to isolate, as they should undergo rapid cyclization or hydrogenation. In fact, hydrogenation of the imine bond is plausible and would account for the formation of larger amounts of 1c (Scheme 6.4, green arrow). Compound 1a can be originated from cyclization of the mono-imine intermediate, 1,3-hydrogen shift and dehydrogenation of the corresponding benzimidazoline intermediate (Scheme 6.4, blue arrows) analogously to previously proposed pathways. The process should be further driven by the aromatization to the corresponding benzimidazole product, which would be aided by the catalyst. Related dehydrogenation of the CH–NH bond to imines and nitriles with alumina- supported Cu nanoparticles is known.52

Scheme 6.4 Proposed reaction network for the formation of 1a, 1b and 1c from

1,2-diaminobenzene (solid arrows indicate the main pathways).

N-Methylbenzimidazole 1b can be obtained through several possible pathways (Scheme 6.4, solid red arrows). Compounds 1c can be a precursor of 1b, which can be obtained by reaction of 1c with formaldehyde, H shift and dehydrogenation. We have verified that compound 1c is indeed converted into 1b in high 84% yield in 3 h at 250 °C (Scheme 6.5, top). This observation is also in agreement with the relatively low amount of 1c in the product mixtures (1–7%, see Table 6.1). A second, less likely possibility is the formation of a diimine intermediate followed by reduction to a monoimine and subsequent cyclization/dehydrogenation to the desired product 1b. A third way to generate 1b is the methylation of benzimidazole 1a, nevertheless low conversion and product yield (∼10% with

(16)

131 both Cu-PMO and Cu–Zn-PMO, Scheme 6.5, bottom) were obtained using benzimidazole 1a as starting material in scMeOH, also according to the expected lower reactivity of the aromatic substrate 1a.

Regarding the reactivity of 2-nitroanilines in the presence of Cu-PMO catalysts in scMeOH, we performed test reactions using 2-nitrotoluene in order to gain insight into the outcome of nitro group reduction. The reaction proceeded with full conversion to N,2-dimethylaniline (57%) and N,N,2-trimethylaniline (43%) after 6 h with Cu-PMO. Under the same experimental conditions, conversion was only 38% with 2-nitroaniline as substrate (Table 6.4, Entry 1). While the reduction of aromatic nitro compounds is already known to proceed in presence of copper catalysts,38–40 to the best of our knowledge no data are available on the copper-mediated cyclization of 2-nitroanilines to benzimidazoles directly from alcohols. Methods employing semiconductor photocatalytic systems are known to perform the mentioned one-pot process.18,20,21

Scheme 6.5 Cyclization of N-methyl-o-phenylendiamine (top) and direct methylation of

benzimidazole (bottom) in scMeOH and in presence of Cu-PMO.

Scheme 6.6 Proposed pathways network for the formation of 1a and 1b from 2-nitroaniline.

Based on thus far available experimental data and previous reports, our preliminary proposed mechanism comprises several steps, summarized in Scheme 6.6. The high selectivity to benzimidazole 1a hints a direct cyclization step at the hydroxylamine stage. The

(17)

132

formed product might undergo subsequent dehydration (Scheme 6.6). In contrast, reduction to 1,2-diaminobenzene, would instead lead to a mixture of 1a and 1b (as shown in Table 6.1) as previously established. The competing hydrogenation of the N=CH2 bond does not take

place to a large extent and the small amount of 1b in these runs seems to coincide in most cases (see Table 6.5) with slow methylation of 1a (see Schemes 6.5 and 6.6). Another interesting point relates to the presence of dopants in the PMO compositions. In this regard, more extended studies on different compositions differing in the Zn loading should shed light on the promoting effect of this dopant towards the nitro group reduction.

6.3 Conclusions

A variety of benzimidazole derivatives have been successfully synthesized in supercritical methanol by means of inexpensive copper catalysts that are derived from earth abundant materials, whose structure and catalytic properties are easily tunable. The solvent serves as a source of in situ formed formaldehyde, thus useful carbon for the construction of the benzimidazole core through the acceptorless dehydrogenative condensation strategy. 1,2-Diaminobenzenes could be converted with moderate to excellent combined yields of benzimidazoles and N-methylbenzimidazoles. Moreover, more readily available 2-nitroanilines could also be used to yield benzimidazoles in even higher selectivity. The described methodology displays a number of advantages: (a) neat methanol is both solvent and reactant and can be replaced by other n-alcohols in order to access targeted benzimidazoles or 2-alkylbenzimidazoles, (b) no additives other than the solvent (oxidants, bases or acids) are needed, (c) catalysts consist of readily available, inexpensive metals and (d) only water and hydrogen are generated as by-products. Future efforts will focus on the synthesis of valuable N- and O-heterocyclic compounds and developing more active dehydrogenation catalysts that can operate at milder reaction conditions.

6.4 Experimental section

6.4.1 Materials and general methods

Chemicals were purchased from Sigma Aldrich, Alfa Aesar or TCI. All chemicals were used as received without further purification.

1

H and 13C NMR spectra were recorded on a Varian AMX400 spectrometer. Powder X-ray analysis was performed on a Bruker XRD diffractometer using Cu Kα radiation and the spectra were recorded in the 2θ angle range of 5°-70°. Elemental analyses were performed on a Perkin Elmer instrument (Optima 7000DV). Mass spectra were recorded on LTQ Orbitrap XL (ESI+). Kinetic modeling of the 2-nitroaniline conversion to the main products benzimidazole and N-methylbenzimidazole was performed with DynaFit 4.

(18)

133

6.4.2 Catalyst preparation

The HTC (hydrotalcite) catalyst precursors were prepared by a coprecipitation method. In a typical procedure, a solution containing AlCl3·6H2O (12.07 g, 0.05 mol), Cu(NO3)2·2.5H2O

(6.98 g, 0.03 mol) and MgCl2·6H2O (24.40 g, 0.12 mol) in deionized water (0.2 L) was added

to a solution containing Na2CO3 (5.30 g, 0.05 mol) in water (0.3 L) at 60oC under vigorous

stirring. The pH was kept between 9 and 10 by addition of small portions of a 1M solution of NaOH. The mixture was vigorously stirred at 60oC for 72 h. After cooling to room temperature, the light blue solid was filtered and resuspended in a 2M solution of Na2CO3

(0.3 L) and stirred overnight at 40oC. The catalyst precursor was filtered and washed with deionized water. After drying the solid for 6 h at 100oC, an hydrotalcite precursor (HTC) was obtained. The corresponding copper-doped porous metal oxide (Cu-PMO) was obtained after calcining this material at 460oC for 24 h in air. Other doped catalysts were prepared according to the same procedure, replacing a defined amount of Mg2+ and/or Al3+ with Zn2+ , Ni2+ and Ru3+.

6.4.3 Catalytic reactions

In a typical experiment, the substrate (1 mmol) was placed in a Swagelok stainless steel microreactor (10 mL) and dissolved in methanol (3 mL). The appropriate amount of catalyst (50 mg) was added, the reactor was sealed and placed in a pre-heated aluminum block at the desired temperature. After the indicated reaction time, the microreactor was cooled down in an ice-water bath and the liquid sample was separated by filtration. Samples were analysed by GC-MS-FID (Hewlett Packard 5890) equipped with a Restek RTX-1701 capillary column (40 oC hold for 5 minutes, then increased to 280 oC with a rate of 10 oC/min and hold for 5 minutes). Isolated products were purified by column chromatography (silica gel, eluent dichloromethane/hexane/methanol).

6.4.4 Recycling tests

After a typical catalytic run (1 mmol 1,2-diaminobenzene, 100 mg Cu-PMO, 3 mL methanol, 250 °C, 6h), the catalyst was separated from the reaction solution by centrifugation and subsequent decantation, additionally washed with methanol (2×10 mL), then with acetone (1×10 mL), and dried overnight at room temperature prior to the next run.

6.4.5 Spectral data of isolated compounds

Benzoimidazole (1a) 1H NMR (400 MHz, DMSO-d 6) δ 12.42 (s, 1H), 8.18 (d, J = 1.6 Hz, 1H), 7.56 (s, 2H), 7.17 (dt, J = 6.5, 3.4 Hz, 2H). 13 C NMR (101 MHz, Chloroform-d) δ 144.96, 141.33, 126.29, 118.69. 1-methyl-1H-benzoimidazole (1b) 1 H NMR (400 MHz, Chloroform-d) δ 7.90 (s, 1H), 7.81 (dd, J = 7.5, 1.7 Hz, 1H), 7.42 – 7.36 (m, 1H), 7.35 – 7.24 (m, 2H), 3.84 (d, J = 1.0 Hz, 3H). 13C NMR (101 MHz, Chloroform-d) δ 143.86, 143.33, 133.50, 123.29, 123.10,

(19)

134 118.69, 110.32, 34.5. 5-methoxy-1H-benzoimidazole (2a) 1 H NMR (400 MHz, Chloroform-d) δ 11.76 (s, 1H), 8.10 (s, 1H), 7.56 (dd, J = 8.9, 0.5 Hz, 1H), 7.11 (dd, J = 2.4, 0.5 Hz, 1H), 6.94 (dd, J = 8.8, 2.4 Hz, 1H), 3.81 (s, 3H). 13 C NMR (101 MHz, Chloroform-d) δ 156.62, 140.58, 137.70, 133.06, 116.49, 112.65, 97.54, 55.81.

Mixture of 5-methoxy-1-methyl-1H-benzo[d]imidazole and 6-methoxy-1-methyl-1H-benzo[d]imidazole (2b) 1 H NMR (400 MHz, Chloroform-d) δ 7.59 – 6.12 (m, 4H), 3.47 – 3.39 (m, 3H), 3.33 (dt, J = 8.6, 1.6 Hz, 3H). 13C NMR (101 MHz, Chloroform-d) δ 156.79, 156.11, 144.40, 143.62, 142.70, 138.08, 129.18, 120.63, 113.12, 111.46, 109.71, 102.13, 92.70, 55.81, 55.76, 31.09, 30.91. 5-methyl-1H-benzo[d]imidazole (3a) 1 H NMR (400 MHz, Chloroform-d) δ 9.64 (s, 1H), 8.08 (s, 1H), 7.55 (d, J = 8.1 Hz, 1H), 7.43 (s, 1H), 7.10 (d, J = 7.9 Hz, 1H), 2.45 (s, 3H). 13 C NMR (101 MHz, Chloroform-d) δ 140.84, 137.65, 136.45 , 132.69, 124.37, 115.42, 114.88, 21.68, 21.67.

Mixture of 1,6-dimethyl-1H-benzo[d]imidazole and 1,5-dimethyl-1H-benzo[d]imidazole (3b) 1 H NMR (400 MHz, Chloroform-d) δ 7.69 – 7.57 (m, 1H), 7.52 (dq, J = 1.6, 0.8 Hz, 1H), 7.14 (d, J = 8.2 Hz, 1H), 7.08 – 6.98 (m, 1H), 3.63 (d, J = 3.6 Hz, 3H), 2.42 (d, J = 5.2 Hz, 3H). 13 C NMR (101 MHz, Chloroform-d) δ 143.99, 143.43, 143.06, 141.74, 134.71, 132.79, 132.61, 131.59, 124.34, 123.60, 119.84, 119.59, 109.23, 108.85, 30.89, 30.78, 21.76, 21.50. Mixture of 5-(tert-butyl)-1-methyl-1H-benzoimidazole and

6-(tert-butyl)-1-methyl-1H-benzoimidazole (4b) 1H NMR (400 MHz, Chloroform-d) δ 7.81 (dd, J = 1.8, 0.7 Hz, 1H), 7.78 (d, J = 3.5 Hz, 1H), 7.42 – 7.27 (m, 2H), 3.76 (s, 3H), 1.40 (s, 9H). 13 C NMR (101 MHz, Chloroform-d) δ 146.53, 145.43, 143.78, 143.52, 143.36, 132.46, 121.00, 120.31, 119.50, 116.43, 108.63, 105.41, 77.30, 34.75, 31.86, 30.96, 30.87.

(20)

135 Mixture of 1,3'-dimethyl-1H,3'H-5,5'-bibenzoimidazole and

3,3'-dimethyl-1H,3'H-5,5'-bibenzoimidazole (5a)

1H-NMR (400 MHz, DMSO-d6): δ 8.21, 8.18 (d, Ar-H), δ 8.00 (s, Ar-H), δ 7.91, 7.89 (d, Ar-H), δ 7.73-7.55 (m, Ar-H), δ 3.91, 3.87 (d, CH3) ppm;

13

C NMR (101 MHz, DMSO-d6) δ 145.69, 145.59, 145.43,

144.53, 136.22, 135.76, 135.61, 135.41, 122.47, 121.67, 121.57, 119.84, 117.97, 117.86, 110.86, 108.95, 31.22, 31.20 ppm. HMRS (ESI) calculated for C16H15N4+ ([M+H]+): 263.12912, found 263.12935.

Mixture of 3-methyl-3H,3'H-5,5'-bibenzoimidazole and 1-methyl-3H,3'H-5,5'-bibenzoimidazole (5b)

1

H-NMR (400 MHz, DMSO-d6): δ 8.26 (s, Ar-H), δ 8.21,

8.20 (d, Ar-H), δ 7.88, 7.85 (d Ar-H), δ 7.70-7.53 (m, Ar-H), δ 3.90, 3.87 (d, CH3) ppm.

13

C NMR (101 MHz, DMSO-d6) δ 145.61, 145.45, 144.50,

143.11, 143.00, 142.95, 136.39, 135.95, 135.77, 135.75, 135.73, 134.28, 122.53, 122.14, 122.10, 121.64, 119.88, 117.88, 116.18, 113.75, 110.86, 108.97, 31.18, 31.15 ppm. HMRS (ESI) calculated for C15H13N4+ ([M+H]+): 249.11347, found 263.11373. 3H,3'H-5,5'-bibenzoimidazole (5c) 1 H NMR (400 MHz, DMSO-d6) δ 8.59 (s, 2H), 8.18 (d, J = 1.7 Hz, 2H), 7.97 (d, J = 8.4 Hz, 2H), 7.82 (dd, J = 8.4, 1.7 Hz, 2H). 13 C NMR (101 MHz, DMSO-d6) δ 143.02, 138.87, 138.01, 136.02, 122.09, 116.24, 113.69. 1H-imidazo[4,5-b]pyridine (6a) 1H NMR (400 MHz, Methanol-d 4) δ 9.08 (d, J = 1.0 Hz, 1H), 8.55 (s, 1H), 8.42 (d, J = 6.0 Hz, 1H), 7.87 (dd, J = 6.1, 1.0 Hz, 1H). 13C NMR (101 MHz, Methanol-d4) δ 147.95, 144.28, 139.28, 138.22, 112.79, 111.57. 5,6-dimethyl-1H-benzoimidazole (10a) 1H NMR (400 MHz, Chloroform-d) δ 9.20 (s, 1H), 8.03 (s, 1H), 7.44 (s, 2H), 2.36 (s, 6H). 13 C NMR (101 MHz, Chloroform-d) δ 139.90, 136.25 , 131.83 , 115.51 , 20.34.

(21)

136 7-methyl-1H-benzoimidazole (12a) 1 H NMR (400 MHz, Chloroform-d) δ 13.16 (s, 1H), 8.29 (s, 1H), 7.61 (d, J = 8.0 Hz, 1H), 7.29 (t, J = 7.7 Hz, 1H), 7.23 – 7.15 (m, 1H), 2.71 (s, 3H). 13 C NMR (101 MHz, Chloroform-d) δ 140.81, 137.72, 137.58, 125.83, 123.31, 122.94, 112.93, 17.54. 2-methyl-1H-benzoimidazole (14a) 1 H NMR (400 MHz, Chloroform-d) δ 8.61 (s, 1H), 7.30 (dd, J = 6.0, 3.2 Hz, 2H), 6.96 (dd, J = 6.0, 3.2 Hz, 2H), 2.39 (s, 3H). 13C NMR (101 MHz, Chloroform-d) δ 148.70, 136.04, 119.59, 111.90, 12.38. 1-ethyl-2-methyl-benzoimidazole (14b) 1 H NMR (400 MHz, Chloroform-d) δ 8.18 – 8.06 (m, 1H), 7.75 – 7.68 (m, 1H), 7.67 – 7.60 (m, 2H), 4.54 (q, J = 7.5 Hz, 2H), 3.00 (s, 3H), 1.80 (t, J = 7.3 Hz, 3H). 13 C NMR (101 MHz, Chloroform-d) δ 151.44, 143.01, 135.04, 122.29, 122.08, 119.37, 109.32, 38.83, 15.29, 14.08. 2-ethyl-1H-benzoimidazole (15a) 1H NMR (400 MHz, Methanol-d 4) δ 7.47 (dd, J = 6.0, 3.2 Hz, 2H), 7.16 (dd, J = 6.0, 3.2 Hz, 2H), 2.90 (q, J = 7.7 Hz, 2H), 1.39 (t, J = 7.6 Hz, 3H). 13 C NMR (101 MHz, Methanol-d4) δ 121.68, 113.83, 21.68, 11.29. 2-ethyl-1-propyl-benzoimidazole (15b) 1 H NMR (400 MHz, Chloroform-d) δ 7.72 (dd, J = 6.0, 2.4 Hz, 1H), 7.33 – 7.26 (m, 1H), 7.26 – 7.15 (m, 2H), 4.09 – 4.02 (m, 2H), 2.89 (q, J = 7.5 Hz, 2H), 1.83 (p, J = 7.4 Hz, 2H), 1.47 (t, J = 7.5 Hz, 3H), 0.97 (t, J = 7.4 Hz, 3H). 13 C NMR (101 MHz, Chloroform-d) δ 155.96, 142.55, 121.88, 121.63, 119.11, 109.22, 45.09, 23.11, 20.77, 11.89, 11.41. 2-propyl-1H-benzoimidazole (16a) 1 H NMR (400 MHz, Methanol-d4) δ 7.46 (dd, J = 6.0, 3.2 Hz, 2H), 7.13 (dd, J = 6.1, 3.2 Hz, 2H), 2.80 (t, J = 7.5 Hz, 2H), 1.81 (h, J = 7.4 Hz, 2H), 0.94 (t, J = 7.4 Hz, 3H). 13 C NMR (101 MHz, Methanol-d4) δ 155.34, 138.11, 121.73, 113.90, 30.31, 21.30, 12.70. 1-butyl-2-propyl-benzoimidazole (16b) 1H NMR (400 MHz, Chloroform-d) δ 7.76 – 7.68 (m, 1H), 7.34 – 7.26 (m, 1H), 7.26 – 7.17 (m, 2H), 4.12 – 4.04 (m, 2H), 2.88 – 2.77 (m, 2H), 1.93 (h, J = 7.4 Hz, 2H), 1.82 – 1.71 (m, 2H), 1.39 (dq, J = 14.8, 7.4 Hz, 2H), 1.06 (t, J = 7.4 Hz, 3H), 0.96 (t, J = 7.3 Hz, 3H). 13 C NMR (101 MHz, Chloroform-d) δ 154.86, 142.64, 134.99, 121.79,

(22)

137 121.60, 119.11, 109.23, 43.40, 31.98, 29.38, 21.20, 20.21, 14.10, 13.76. 2-butyl-1H-benzoimidazole (17a) 1 H NMR (400 MHz, Methanol-d4) δ 7.47 (dd, J = 6.0, 3.2 Hz, 2H), 7.17 (dd, J = 6.0, 3.2 Hz, 2H), 2.88 (t, J = 7.7 Hz, 2H), 1.81 (p, J = 7.6 Hz, 2H), 1.41 (q, J = 7.5 Hz, 2H), 0.97 (t, J = 7.4 Hz, 3H). 13 C NMR (101 MHz, Methanol-d4) δ 156.85, 123.04, 115.20, 31.40, 29.42, 23.32, 14.01. 2-butyl-1-pentyl-benzoimidazole (17b) 1 H NMR (400 MHz, Chloroform-d) δ 7.74 – 7.69 (m, 1H), 7.30 – 7.25 (m, 1H), 7.21 (dt, J = 5.6, 2.1 Hz, 2H), 4.10 – 4.03 (m, 2H), 2.88 – 2.81 (m, 2H), 1.88 (p, J = 7.6 Hz, 2H), 1.78 (t, J = 7.5 Hz, 2H), 1.53 – 1.42 (m, 2H), 1.38 – 1.31 (m, 4H), 0.98 (t, J = 7.3 Hz, 3H), 0.92 – 0.87 (m, 3H). 13 C NMR (101 MHz, Chloroform-d) δ 155.04, 121.78, 121.58, 119.11, 109.20, 43.61, 29.91, 29.58, 29.05, 27.20, 22.68, 22.34, 13.89, 13.86.

6.5 References

1 M. T. H. Khan, Bioactive Heterocycles IV, Springer, 2007, vol. 10. 2 L. B. Townsend and G. R. Revankar, Chem. Rev., 1970, 70, 389–438.

3 A. R. Porcari, R. V Devivar, L. S. Kucera, J. C. Drach and L. B. Townsend, J. Med. Chem., 1998, 41, 1252–1262.

4 M. T. Migawa, J. Girardet, J. a W. Ii, G. W. Koszalka, S. D. Chamberlain, J. C. Drach and L. B. Townsend, J. Med. Chem., 1998, 41, 1242–1251.

5 V. Yerragunta, S. Srujana, R. Devi, A. Divya, A. Pradesh, A. Pradesh and M. Drugs, PharmaTutor, 2014, 2, 109–113.

6 L. C. R. Carvalho, E. Fernandes and M. M. B. Marques, Chem. Eur. J, 2011, 17, 12544– 12555.

7 D. A. Horton, G. T. Bourne and M. L. Smythe, Chem. Rev., 2003, 103, 893–930. 8 E. C. Wagner and W. H. Millett, Org. Synth., 1939, 19, 12.

9 C. Zhu and Y. Wei, ChemSusChem, 2011, 4, 1082–1086. 10 S. M. Landge and B. Török, Catal. Lett., 2008, 122, 338–343.

11 Z. Mao, Z. Wang, J. Li, X. Song and Y. Luo, Synth. Commun., 2010, 40, 1963–1977. 12 G. Navarrete‐Vázquez, H. Moreno‐Diaz, S. Estrada‐Soto, M. Torres‐Piedra, I. León‐

Rivera, H. Tlahuext, O. Muñoz‐Muñiz and H. Torres‐Gómez, Synth. Commun., 2007, 37, 2815–2825.

13 G. V. Reddy, V. V. V. N. S. Rama Rao, B. Narsaiah and P. Shanthan Rao, Synth. Commun., 2002, 32, 2467–2476.

14 H. H. Khoo, Renew. Sustain. Energy Rev., 2015, 46, 100–119. 15 K. Barta and P. C. Ford, Acc. Chem. Res., 2014, 47, 1503–1512.

16 V. R. Ruiz, A. Corma and M. J. Sabater, Tetrahedron, 2010, 66, 730–735.

17 M. Bala, P. K. Verma, U. Sharma, N. Kumar and B. Singh, Green Chem., 2013, 15, 1687– 1693.

(23)

138

18 H. Wang, R. E. Partch and Y. Li, J. Org. Chem., 1997, 62, 5222–5225.

19 Y. Shiraishi, Y. Sugano, S. Tanaka and T. Hirai, Angew. Chem. Int. Ed., 2010, 49, 1656– 1660.

20 K. Selvam and M. Swaminathan, Tetrahedron Lett., 2011, 52, 3386–3392.

21 M. Annadhasan, K. Selvam and M. Swaminathan, Synth. Commun., 2012, 42, 1500– 1508.

22 C. Gunanathan and D. Milstein, Science, 2013, 341, 1229712: 1-11. 23 G. E. Dobereiner and R. H. Crabtree, Chem. Rev., 2010, 110, 681–703. 24 S. Michlik and R. Kempe, Nat. Chem., 2013, 5, 140–144.

25 D. Forberg, J. Obenauf, M. Friedrich, S.-M. Hühne, W. Mader, G. Motz and R. Kempe, Catal. Sci. Technol., 2014, 4, 4188–4192.

26 D. Pingen and D. Vogt, Catal. Sci. Technol., 2014, 4, 47-52.

27 T. Hille, T. Irrgang and R. Kempe, Chem. - A Eur. J., 2014, 20, 5569–5572.

28 A. J. Blacker, M. M. Farah, M. I. Hall, S. P. Marsden, O. Saidi and J. M. J. Williams, Org. Lett., 2009, 11, 2039–2042.

29 K. Barta, T. D. Matson, M. L. Fettig, S. L. Scott, A. V. Iretskii and P. C. Ford, Green Chem., 2010, 12, 1640–1647.

30 T. D. Matson, K. Barta, A. V Iretskii and P. C. Ford, J. Am. Chem. Soc., 2011, 133, 14090–14097.

31 G. S. Macala, T. D. Matson, C. L. Johnson, R. S. Lewis, A. V. Iretskii and P. C. Ford, ChemSusChem, 2009, 2, 215–217.

32 A. V. F. Cavani, F. Trifirb, Catal. Today, 1991, 11, 173–301. 33 R. A. Smiley, Ullmann’s Encycl. Ind. Chem., 2000.

34 D. S. VanVliet, P. Gillespie and J. J. Scicinski, Tetrahedron Lett., 2005, 46, 6741–6743. 35 E. J. Hanan, B. K. Chan, A. a. Estrada, D. G. Shore and J. P. Lyssikatos, Synlett, 2010, 18,

2759–2764.

36 T. B. Nguyen, J. Le Bescont, L. Ermolenko and A. Al-mourabit, Org. Lett., 2013, 15, 6218–6221.

37 N. a. Weires, J. Boster and J. Magolan, European J. Org. Chem., 2012, 6508–6512. 38 A. Saha and B. Ranu, J. Org. Chem., 2008, 73, 6867–6870.

39 R. Kaur, C. Giordano, M. Gradzielski and S. K. Mehta, Chem. Asian J., 2014, 9, 189–198. 40 H. Feng, Y. Li, S. Lin, E. V Van der Eycken and G. Song, Sustain. Chem. Process., 2014, 2,

14.

41 J. C. Lewis, R. G. Bergman and J. a. Ellman, Acc. Chem. Res., 2008, 41, 1013–1025. 42 V. Ritleng, C. Sirlin and M. Pfeffer, Chem. Rev., 2002, 102, 1731–1769.

43 K. L. Tan, S. Park, J. a. Ellman and R. G. Bergman, J. Org. Chem., 2004, 69, 7329–7335. 44 K. L. Tan, R. G. Bergman and J. a. Ellman, J. Am. Chem. Soc., 2002, 124, 13964–13965. 45 S. Pivsa-Art, T. Satoh, Y. Kawamura, M. Miura and M. Nomura, Bull. Chem. Soc. Jpn.,

1998, 71, 467–473.

46 E. J. Moore, W. R. Pretzer, T. J. Oconnell, J. Harris, L. Labounty, L. Chou and S. S. Grimmer, J. Am. Chem. Soc., 1992, 114, 5888–5890.

47 K. L. Tan, R. G. Bergman and J. A. Ellman, J. Am. Chem. Soc., 2001, 123, 2685–2686. 48 L. Zhang, Y. Zhang, Y. Deng and F. Shi, Catal. Sci. Technol., 2015, 5, 3226–3234. 49 J. Bain, P. Cho and A. Voutchkova-Kostal, Green Chem., 2015, 17, 2271–2280. 50 P. Ghosh and A. Mandal, Catal. Commun., 2011, 12, 744–747.

51 M. Perspective and W. M. Dehydrogenation, ACS Nano, 2013, 7, 6181–6187.

Referenties

GERELATEERDE DOCUMENTEN

However, by using capillary microreactors made of polytetrafluoroethylene (PTFE), a dewetted slug flow was generated under certain operating conditions (i.e., characterized

The oxidation of benzyl alcohol using such catalyst is highly selective and functions as a model reaction to develop mass transfer characteristics of similar

Sustainable Pathways to Chemicals and Fuels from Lignocellulose via Catalytic Cleavage and Coupling Reactions Zhuohua Sun PhD Thesis University of Groningen ISBN:

As cellulose and hemicellulose have relatively simpler structure compared to lignin, various methods regarding valorization of (hemi)cellulose have been developed for a long time and

Indeed, while the 2D NMR spectrum of a sample obtained on standard catalytic treatment (Cu20-PMO and H 2 ) was assigned to the main product 1G (Figure 2.2a) without any

Selective functionalization strategies should target atom economic and waste free pathways for the direct conversion of aliphatic and aromatic alcohols to amines or the

Cyclopentanone has a cyclic structure and can be obtained by catalytic conversion of furfural 20,21 (Scheme 4.1) - which is exclusively produced from

Figure 5.12 Product formation profile for the Guerbet reaction of ethanol to 1-butanol in a continuous flow fixed-bed reactor at different temperature... 109 Figure 5.13