Sustainable and Selective Monomethylation of ... - ACS Publications

Oct 30, 2017 - J. Am. Chem. Soc. 2013, 135, 11704−11707. (65) Torres, P. U. Cinacalcet HCl: a novel treatment for secondary hyperparathyroidism caus...
1 downloads 0 Views 2MB Size
Research Article pubs.acs.org/journal/ascecg

Sustainable and Selective Monomethylation of Anilines by Methanol with Solid Molecular NHC-Ir Catalysts Jiangbo Chen,† Jiajie Wu,† and Tao Tu*,† †

Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Department of Chemistry, Fudan University, 220 Handan Road, Shanghai, 200433, China S Supporting Information *

ABSTRACT: Using feedstock methanol as a green methylation reagent, the selective N-monomethylation of anilines is realized under mild reaction conditions by using N-heterocyclic carbene iridium (NHC-Ir) coordination assemblies as highly efficient solid molecular catalysts. Along with a broad substrate scope and good functional group tolerance, up to quantitative yield and 2.0 × 104 turnover numbers (TONs) are obtained even at low catalyst loadings. Notably, the solid NHC-Ir molecular catalyst can be easily recovered and recycled more than 20 times without obvious loss of reactivity and selectivity. Furthermore, this selective practical protocol can be successfully extended to direct methylation of highly functionalized bioactive compounds including 3aminoestrone, cinacalcet, and their analogues in excellent yields and selectivities, highlighting their potential application in pharmaceuticals. KEYWORDS: Amine synthesis, Coordination assembly, Iridium, Methanol activation, Catalytic methylation



using Pd/TiO223 or Ag/TiO224 as nanophotocatalysts. Besides fully methylated products are produced in the presence of primary amine substrates, however, the recyclability of the heterogeneous catalyst is not addressed in their studies. By using a “hydrogen-borrowing” strategy,25−32 N-methylation of primary amines by MeOH in the presence of bifunctional homogeneous catalysts has been disclosed. In general, the bifunctional catalysts not only favor the oxidative dehydrogenation of methanol but also accelerate the hydrogenation of the imine intermediate to produce the desired amines.29−32 Among them, Li and co-workers realized the first Ir-catalyzed Nmonomethylation of aromatic primary amines,33 although ortho-substituted aromatic amines are not suitable substrates and a high reaction temperature is still required. By using air sensitive [RuCp*Cl2]2/dpePhos as a catalyst, Seayad and coworkers fulfilled N-monomethylation of less-bulky aromatic primary amines with MeOH.34 In light of their robustness and high catalytic activity in various transformations,35−49 Crabtree and co-workers first applied N-heterocyclic carbene (NHC) Ir complexes in this challenging reaction.50 However, high catalyst loading and limited substrate scope even under the microwave assisted conditions still hindered the protocol feasibility. Furthermore, besides these advancements achieved by the homogeneous catalysts, the difficulty in the separation and recyclability of these noble catalysts is still considered as a stumbling block in practical industrial applications. In addition,

INTRODUCTION Direct N-methylation of amines is one of the most important transformations in organic chemistry and is widely applied in the syntheses of fine chemicals, pharmaceuticals, and natural products.1−4 The traditional N-methylation protocols, usually involving methyl halides, dimethylsulfate, and diazomethane, have several serious limitations, especially in the selectivity of the N-monomethylation of primary amines.5−7 The excess amount of strong bases, overmethylated products, and toxic halogenated byproducts causes tedious separation and thus hampering their practical application in the industrial setting. Therefore, reductive amination of toxic formaldehyde still constitutes a chief approach in the industrial production of methyl amines.8−11 As more environmentally benign reagents, carbon dioxide,12 formic acid,13−15 and dimethyl carbonate16−18 have also demonstrated their applicability in the methylation of amines. However, high reaction temperature (>150 °C) and excess reducing agents are usually required to achieve satisfactory yields. An intrinsic problem is the often concurrent production of byproducts tertiary amines with these substances. The formed nucleophilic monomethylated amines prefer the tertiary amines formation. Therefore, the direct and selective N-monomethylation of amines remains elusive. As an important biomass and cost-effective feedstock fine chemical, methanol has also been utilized for this purpose. Various Lewis acids have demonstrated their catalytic efficiency in the N-methylation of amines with MeOH, although harsh reaction conditions are still required to achieve satisfactory results for fully methylated products.19−22 Recently, Saito and co-workers reported a room temperature N-methylation by © 2017 American Chemical Society

Received: September 14, 2017 Revised: October 27, 2017 Published: October 30, 2017 11744

DOI: 10.1021/acssuschemeng.7b03246 ACS Sustainable Chem. Eng. 2017, 5, 11744−11751

Research Article

ACS Sustainable Chemistry & Engineering methanol is also well applied in the hydrogen transfer reactions51−53 and intermolecular dehydrogenated coupling with amines,54−56 the study on the competition and selectivity in the N-methylation of primary amines containing sensitive groups including olefins, ketones, nitriles, hydroxyl groups, etc. is still less studied. Therefore, searching efficient, selective and robust recyclable catalysts for sustainable N-methylation is very concerned.



Scheme 1. NHC-Ir Coordination Assemblies 1a−f and Corresponding Molecular Complexes 2a,b

EXPERIMENTAL SECTION

General Synthetic Procedure of N-Heterocyclic Carbene Iridium (NHC-Ir) Coordination Assemblies. Benzobisimidazolium salts (0.5 mmol) and the corresponding iridium precursor (0.5 mmol) were dissolved in DMF under N2 at room temperature, and LiHMDS (1 mmol) solution in THF was added in dropwise. The resulting mixture was stirred at 80 °C for 12 h. The dark brown solids were isolated after filtration and washed with deionized water three times. The coordination polymer solids 1a−1f were then dried under vacuum. General Procedure of N-Monomethylation of Aniline by Methanol and the Solid Molecular Catalyst Recovery. A 15 mL sealed tube equipped with a magnetic stirring bar was charged with aniline (279 mg, 3 mmol), solid molecular catalyst 1a (8.1 mg, 0.5 mol %), KOtBu (336 mg, 3 mmol), and methanol (6 mL). After heating to the desired temperature for 12 h, the mixture was cooled to room temperature; the solid catalysts were readily recovered after centrifugation and decantation. The recovered solid was washed with methanol (3 × 5 mL) and reused directly in the next run without additional activation steps, just by simply recharging aniline, KOtBu, and MeOH in the sealed tube. General Procedure for the N-Methylation of Primary Amines by Methanol. Amine (1.0 mmol), solid catalyst 1a (2.6 mg, 0.5 mol %), and KOtBu (112 mg, 1.0 equiv) were charged in a 15 mL pressure tube with a magnetic bar. Then, 2 mL of methanol was added into the mixture. The reaction tube was closed with the Teflon stopper and heated to 130 °C. After reaction completion, the reaction mixture was allowed to cool to room temperature. The solvent was removed under vacuum, and the resulting mixture was purified by flash column chromatography. The isolated products were fully characterized. Synthetic Procedure of Secondary Amines 25a−d. The aliphatic primary amine and the same equivalent ketone or aldehyde was condensed with Ti(O-iPr)4 at room temperature. After amine consumption, methanol was added. The resulting mixture was cooled to 0 °C. Then NaBH4 was added in portion to give the desired secondary amines 25a−d in good to excellent yields.

Initially, the N-monomethylation of aniline in pure MeOH was selected as a model reaction to investigate the catalyst efficiency. After intensively screening various bases (see the Supporting Information, Table S1), delightedly, in the presence of 0.1 mol % solid catalyst 1a (Ir content), KOtBu was found as a suitable base for this heterogeneous catalytic reaction and to deliver the best GC-yield at 130 °C (80%, entry 1, Table 1). Table 1. Reaction Optimization of N-Monomethylation of Aniline by Methanola

entry

[Cat]

t (h)

yield (%)b

1 2 3 4 5 6 7 8 9 10c

1a 1b 1c 1d 1e 1f 2a 2b 1a 1a

12 12 12 12 12 12 12 12 18 12

80 trace 60 trace 6 8 85 55 >99 >99

a

Reaction was carried out with 4 mmol aniline, 0.1 mol % [Cat], and 1.0 equiv KOtBu in 8 mL of methanol. bYield was determined by GCMS using p-xylene as the internal standard. cWith 1 mmol aniline, 0.5 mol % [Cat], and 1.0 equiv KOtBu and 2 mL of methanol.



RESULTS AND DISCUSSION Recently, using a coordination assembly strategy,57−60 we have prepared a series of coordination assemblies based on bisNHC-M complexes (M = Ir and Ru),61−63 which function as solid bifunctional molecular catalysts not only in oxidative dehydrogenation of polyols61 but also in the reductive amination of levulinic acid and its analogues.62 Notably, these solid assemblies could be easily recovered and reused dozens of times without obvious observation of activity or selectivity loss, highlighting their potential in industrial applications. Among these transformations, we found that these solid assemblies exhibited better catalytic activity than their corresponding molecular NHC-M complexes. Therefore, in our continuing interests in developing solid molecular catalysts and their catalytic application in the selective biomass transformation, herein, we wish to report the results of the studies by uncovering NHC-Ir coordination assemblies (Scheme 1) functioning as efficient and recyclable solid bifunctional molecular catalysts in the selective N-monomethylation of diverse anilines by MeOH via reductive amination process.

The reaction temperature affected the methylation outcomes dramatically: the yield of product 3 was descended along with the reaction temperature decreasing. For example, a 98% GCyield of the desired product 3 was found at 130 °C, while only a 34% yield was detected at 100 °C with the same reaction duration (Table S1, entry 15 vs 3). Subsequently, other solid catalysts (1b−f) were probed. In contrast with anion effects, ancillary ligands show stronger impact on the methylation process. When iodide containing solid catalyst 1c was applied, product 3 was still formed in a 60% yield (entry 3, Table 1). However, when COD (cyclooctadiene) ligand containing catalysts 1b and 1d were utilized, almost no methylated products were detected (entries 2 and 4, Table 1). Slightly increased yields were observed with the solid catalysts containing longer alkyl groups (6% and 8% for 1e and 1f, respectively, entries 5 and 6, Table 1). When corresponding homogeneous NHC-Ir catalyst 2a was applied instead of solid 11745

DOI: 10.1021/acssuschemeng.7b03246 ACS Sustainable Chem. Eng. 2017, 5, 11744−11751

Research Article

ACS Sustainable Chemistry & Engineering

was carried out. Delightedly, the recovered solid 1a after 23 runs display similar images as the freshly prepared one (Figure 2a−c vs d−f) which indicate that assembly 1a preserves its molecular structural features during recycling. In order to exclude the possibility of the formation of Ir nanoparticles during the reaction, mercury poisoning experiments were performed. After heating, one drop of mercury was added into the reaction mixture after 0 and 4 h. To our delight, the reactions are not prohibited, and excellent yields are still observed (98% and 97%, respectively). These outcomes exclude the possibility of nanoparticle formation, which is also consistent with our previous study that NHC-Ir coordination assemblies 1a functioned as a solid molecular catalyst in these transformations.62,63 Furthermore, the filtrates after each run were analyzed by inductively coupled plasma-atomic emission spectroscopy (ICP-AES) for potential iridium leaching (see the Supporting Information, Table S3). Approximately 0.2 μg/mL iridium leaching was observed for each in the initial three runs, which might be the trace amount of the inactive iridium species or precursors entrapped in the assemblies’ matrix cavities during the metal coordination process. In the later, the real leaching amount of every run was negligible, which further indicated the robustness of NHC-Ir assemblies even in the polar solvent like methanol. The yield decreasing after 20 runs might be caused by the catalyst loss during the dozens of centrifugation, decantation, and washing processes. Encouraged by the performance of solid molecular catalyst 1a observed so far, a scale-up reaction with 100 mmol of aniline in 75 mL of methanol was then conducted (Scheme S5). In the presence of 5 × 10−3 mmol of 1a, aniline was completely converted to Nmethyl aniline 3 at 150 °C with an extended reaction time, and the highest TONs, 2 × 104, was achieved for this challenging transformation. With this practical and efficient protocol in hand, the scope on other amines was then investigated (Table 2). The substituent position on the aromatic shows limited impact on the transformation: meta- and para-methyl substituted anilines give excellent isolated yields (96% and 91% for Nmonomethylated products 4b and 4c, respectively), and

1a, a slightly better yield was obtained (85%, entry 7, Table 1). However, benzimidazolium analogue 2b only resulted in a 55% yield (entry 8, Table 1). To our delight, slightly extending the reaction time to 18 h or increasing the catalyst loading to 0.5 mol %, quantitative yields were finally achieved (entries 8 and 9, Table 1). Due to its insolubility in water, substrates, and the most organic solvents, the solid catalyst 1a is readily recovered quantitatively by convenient centrifugation. To our delight, after simple filtration and washing with additional methanol, the recovered solids could be directly applied in the next run by simply adding base, aniline, and methanol. Almost quantitative yields (>99%) were obtained in the first 13 runs (blue column, Figure 1) under the optimized reaction conditions. Slightly

Figure 1. Recycling and reuse of the NHC-Ir coordination assemblies 1a in the N-monomethylation of aniline: 3 mmol scale for each run (reaction time: blue 12 h and brown 24 h).

decreased yields were observed in the 14th (96%) and 15th (95%) runs. After extending the reaction time to 24 h (brown column, Figure 1), the yield could keep quantitatively until the 20th run. By using a scanning electron microscope (SEM), a transmission electron microscope (TEM), and energy dispersive X-ray spectroscopy (EDS) mapping, a comparison between the freshly prepared and recovered solid catalyst 1a

Figure 2. (a−c) TEM, SEM, and EDS images of freshly prepared NHC-Ir coordination assemblies 1a; (d−f) images of recovered solid 1a after the 23th run. 11746

DOI: 10.1021/acssuschemeng.7b03246 ACS Sustainable Chem. Eng. 2017, 5, 11744−11751

Research Article

ACS Sustainable Chemistry & Engineering Table 2. N-Methylation of Aromatic Primary Aminea

a

1 mmol scale in 2 mL of methanol with 0.5 mol % solid catalyst 1a; isolated yield. bAfter 48 h. cWith 10 mol % solid catalyst 1a for 3.5 d.

sterically hindered ortho-substituted aniline still results in a moderate yield (60%, 4a). Gratifyingly, the yield of 4a could readily be improved to 92% by simply prolonging the reaction time. Remarkably, amines containing both electron-rich (4-7) and electron-deficient (8-11) substitutes were fully compatible; the corresponding N-monomethylated amines were produced in excellent yields (89−99%). It is worth noting that nitrile group, easily reduced in the hydrogen transfer reactions,64 was not affected under the reaction conditions. The halide groups were also fully tolerated, and the corresponding N-monomethylated products (11−13) were isolated in 85−99% yields; the iodo and chloro groups are kept and ready for further transformations. To our delight, bulky 1-naphthalamine and heterocyclic aryl amines were all considered as suitable substrates, leading to excellent yields (82%−99%) for the desired products (14−19). Moreover, sulfonamide was also found as a suitable partner and selectively N-monomethylated to produce compound 20 in an excellent isolated yield, although an extended reaction time was required. When the substrate contained both amino and amide groups, the N-monomethylation selectively took place at the amine side resulting in the desired product 21 in an 85%

isolated yield. No amide-methylated product was observed. When aniline with an unprotected hydroxymethyl group was employed, a 91% yield of product 22 was obtained. No intermolecular dehydrogenated coupling products were detected,49−51 further highlighting the reaction selectivity and compatibility. Pleasingly, the protocol was successfully extended to diamine and triamine substrates and the corresponding dimethyl and trimethyl products (23 and 24) were produced in 99% and 72% isolated yields, respectively. Notably, this protocol is readily extended to the methylation of crucial secondary alkyl-amines. For instance, the methylated cinacalcet 25a, which is used to treat secondary hyperparathyroidism,65 was readily prepared in an 88% isolated yield. Its analogues (25b−d) were efficiently synthesized in good to excellent yields (84%−94%). It is well-known that olefins can be readily reduced via hydrogen transfer process in alcohol.51−53 When 4-aminostyrene was selected to test the selectivity of two competitive reactions. As we expected, poor selectivity was observed under the optimized conditions, and the fully reductive product 26 and N-monomethylated product 27 with olefin group were produced almost in equal amounts (Scheme 2). The 11747

DOI: 10.1021/acssuschemeng.7b03246 ACS Sustainable Chem. Eng. 2017, 5, 11744−11751

Research Article

ACS Sustainable Chemistry & Engineering Scheme 2. Competition between Olefin Hydrogen Transfer and N-Monomethylation of 4-Aminostyrene

Scheme 3. N-Monomethylation of 3-Aminoestrone

be first oxidized to formaldehyde, which was subsequently condensed with an amine to produce the aromatic imine intermediate with the liberation of 1 equiv. of water. This resulting imine was then hydrogenated to give the corresponding N-monomethylation product (Scheme 4). To gain

optimization studies of reaction conditions (see the Supporting Information, Table S1) revealed that bases dramatically influenced the methylation process. Therefore, we conceived the alternation of different bases may tune the selectivity between the hydrogenation of olefin and methylation of amine. It was found that (see the Supporting Information, Table S2), when NaOtBu was employed instead of KOtBu, delightedly, up to 81% isolated yield of N-monomethylated product 27 was obtained (Scheme 2). Alternatively, it was possible to selectively produce full reductive product 26 in a 95% isolated yield by using KOtBu as the base with extended reaction time (Scheme 2). To examine the potential feasibility of this selective, efficient, and environmental benign methodology, we would like to demonstrate its application in the N-monomethylation of bioactive complex aromatic amine, which could be exploited for drug discovery. 3-Aminoestrone is known as a versatile intermediate for the synthesis of a number of biologically active compounds for the treatment of hormono-sensitive diseases including prostate and breast cancers.66,67 Therefore, direct N-monomethylation of 3-aminoestrone was then carried out with the solid molecular catalyst 1a (Scheme 3). Due to the competitive C-methylation at the ortho site of the carbonyl group,68,69 product 28 was just obtained in a 31% isolated yield. In order to avoid this influence, the protected 3-aminoestrone 29 was then synthesized and applied under the identical reaction conditions, and a good yield (80%) was observed with N-monomethylated product 30. After simple hydrolysis, secondary amine 30 was readily converted to 28 almost in a quantitative yield. In light of the hydrogen-borrowing process that may be involved in this methylation reaction,29−32,70−72 methanol may

Scheme 4. Proposed Mechanism for Ir-Catalyzed NAlkylation of Amines Using Alcohol

mechanistic insights, the reaction trajectory of the Nmonomethylation of aniline by the mixture of methanol and deuterated methanol (1:1 ratio) was monitored by 1H NMR. In order to exclude the possible oxidation under ambient conditions, the reaction was carried out in a sealed NMR tube. The expected imine NCH2 signal was found at 8.57 ppm (see the Supporting Information, Figure S25), supporting the proposed reaction mechanism. With this plausible mechanism in hand, the protocol should be extended to other long-chain alcohols. When ethanol and 1propanol were applied as the solvent and alkylating reagents, the desired N-monoalkylated products were still detected in 59% and 29% yields, respectively (eqs 1 and 2), which further 11748

DOI: 10.1021/acssuschemeng.7b03246 ACS Sustainable Chem. Eng. 2017, 5, 11744−11751

Research Article

ACS Sustainable Chemistry & Engineering

(3) Amines: Synthesis, Properties, and Applications; Lawrence, S. A., Ed.; Cambridge University: Cambridge, U.K., 2006. (4) Amino Group Chemistry: From Synthesis to the Life Sciences; Ricci, A., Ed.; Wiley-VCH: Weinheim, Germany, 2008. (5) Selva, M.; Bomben, A.; Tundo, P. Selective mono-N-methylation of primary aromatic amines by dimethyl carbonate over faujasite Xand Y-type zeolites. J. Chem. Soc., Perkin Trans. 1 1997, 1, 1041−1046. (6) Selva, M.; Perosa, A.; Fabris, M. Sequential coupling of the transesterification of cyclic carbonates with the selective Nmethylation of anilines catalysed by faujasites. Green Chem. 2008, 10, 1068−1077. (7) Savourey, S.; Lefèvre, G.; Berthet, J.-C.; Cantat, T. Catalytic methylation of aromatic amines with formic acid as the unique carbon and hydrogen source. Chem. Commun. 2014, 50, 14033−14036. (8) Hollmann, D. Advances in asymmetric borrowing hydrogen catalysis. ChemSusChem 2014, 7, 2411−2413. (9) Usman, M.; Daud, W. M. A. W. Recent advances in the methanol synthesis via methane reforming processes. RSC Adv. 2015, 5, 21945− 21972. (10) Ali, K. A.; Abdullah, A. Z.; Mohamed, A. R. Recent development in catalytic technologies for methanol synthesis from renewable sources: A critical review. Renewable Sustainable Energy Rev. 2015, 44, 508−518. (11) Yan, G.; Borah, A. J.; Wang, L.; Yang, M. Recent Advances in Transition Metal-Catalyzed Methylation Reactions. Adv. Synth. Catal. 2015, 357, 1333−1350. (12) Li, Y.; Cui, X.; Dong, K.; Junge, K.; Beller, M. Utilization of CO2 as a C1 building block for catalytic methylation reactions. ACS Catal. 2017, 7, 1077−1086. (13) Savourey, S.; Lefevre, G.; Berthet, J.-C.; Cantat, T. Catalytic methylation of aromatic amines with formic acid as the unique carbon and hydrogen source. Chem. Commun. 2014, 50, 14033−14036. (14) Zhu, L.; Wang, L.-S.; Li, B.; Li, W.; Fu, B. Methylation of aromatic amines and imines using formic acid over a heterogeneous Pt/C catalyst. Catal. Sci. Technol. 2016, 6, 6172−6176. (15) Sorribes, I.; Junge, K.; Beller, M. General catalytic methylation of amines with formic acid under mild reaction conditions. Chem. Eur. J. 2014, 20, 7878−7883. (16) Zheng, J.; Darcel, C.; Sortais, J.-B. Methylation of secondary amines with dialkyl carbonates and hydrosilanes catalysed by iron complexes. Chem. Commun. 2014, 50, 14229−14232. (17) Li, Y.; Sorribes, I.; Vicent, C.; Junge, K.; Beller, M. Convenient reductive methylation of amines with carbonates at room temperature. Chem. - Eur. J. 2015, 21, 16759−16763. (18) Cabrero-Antonino, J. R.; Adam, R.; Junge, K.; Beller, M. A general protocol for the reductive N-methylation of amines using dimethyl carbonate and molecular hydrogen: mechanistic insights and kinetic studies. Catal. Sci. Technol. 2016, 6, 7956−7966. (19) Oku, T.; Ikariya, T. Enhanced product selectivity in continuous N-methylation of amino alcohols over solid acid-base catalysts with supercritical methanol. Angew. Chem., Int. Ed. 2002, 41, 3476−3479. (20) Luque, R.; Campelo, J. M.; Luna, D.; Marinas, J. M.; Romero, A. A. Catalytic performance of Al-MCM-41 materials in the N-alkylation of aniline. J. Mol. Catal. A: Chem. 2007, 269, 190−196. (21) Niphadkar, P. S.; Joshi, P. N.; Gurav, H. R.; Deshpande, S. S.; Bokade, V. V. Synthesis of N-methylaniline by aniline alkylation with methanol over Sn-MFI molecular sieve. Catal. Lett. 2009, 133, 175− 184. (22) Su, J.; Li, X.; Chen, Y.; Cui, Y.; Xu, J.; Qian, C.; Chen, X. NMethylation of amines with methanol in a hydrogen free system on a combined Al2O3−mordenite catalyst. RSC Adv. 2016, 6, 55643− 55649. (23) Zhang, L.; Zhang, Y.; Deng, Y.; Shi, F. Light-promoted N, Ndimethylation of amine and nitro compound with methanol catalyzed by Pd/TiO2 at room temperature. RSC Adv. 2015, 5, 14514−14521. (24) Tsarev, V. N.; Morioka, Y.; Caner, J.; Wang, Q.; Ushimaru, R.; Kudo, A.; Naka, H.; Saito, S. N-methylation of amines with methanol at room temperature. Org. Lett. 2015, 17, 2530−2533.

prove the reaction mechanism and protocol efficiency. The low yields of N-monoalkylated products may be caused by the selfaldol condensation reaction of the aldehydes under the basic reaction conditions.73−75 Moreover, the unsaturated aldehyde product from the self-aldol condensation would react with amine to give various alkylated byproducts, which also contributed to the diminished yield of the desired Nmonoalkylated products (Figure S120).



CONCLUSIONS In conclusion, by using the NHC-Ir solid molecular catalyst, we have developed the selective N-monomethylation of aromatic anilines with methanol under mild conditions. Besides good functional group tolerance and broad substrates scope even at low catalyst loading, the solid molecular catalyst was readily recovered by simple centrifugation and reused more than 20 times without obvious loss of activity. Up to 2 × 104 TONs was achieved and clearly demonstrated the efficiency of our developed method. In addition, due to the high selectivity observed in the N-monomethylation primary amines containing competitive groups including olefins, ketones, nitriles, and free hydroxyls, we successfully applied this method to the methylation of bioactive molecules with excellent performance, highlighting its great practical potential in the sustainable amine synthesis and pharmaceuticals.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.7b03246. Experimental details and characterization data. (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Tao Tu: 0000-0003-3420-7889 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support from the National Key R&D Program of China (2016YFA0202902), National Natural Science Foundation of China (Nos. 21572036 and 91127041), the Shanghai International Cooperation Program (14230710600), the External Cooperation Program of Jiangxi Province (20151BDH80045), and Department of Chemistry, Fudan University is gratefully acknowledged.



REFERENCES

(1) Modern Amination Methods; Ricci, A., Ed.; Wiley-VCH: Weinheim, Germany, 2000. (2) Carey, F. A.; Sundberg, R. J. Advanced Organic Chemistry, 4th ed.; Kluwer Academic: New York, 2001; Part B, Chapter 3.2.5. 11749

DOI: 10.1021/acssuschemeng.7b03246 ACS Sustainable Chem. Eng. 2017, 5, 11744−11751

Research Article

ACS Sustainable Chemistry & Engineering (25) Prades, A.; Corberán, R.; Poyatos, M.; Peris, E. [IrCl2Cp*(NHC)] complexes as highly versatile efficient catalysts for the crosscoupling of alcohols and amines. Chem. - Eur. J. 2008, 14, 11474− 11479. (26) Zanardi, A.; Mata, J. A.; Peris, E. One-Pot preparation of imines from nitroarenes by a tandem process with an Ir-Pd heterodimetallic catalyst. Chem. - Eur. J. 2010, 16, 10502−10506. (27) Crabtree, R. H. An organometallic future in green and energy chemistry? Organometallics 2011, 30, 17−19. (28) Elangovan, S.; Sortais, J.-B.; Beller, M.; Darcel, C. Iron-catalyzed α-alkylation of ketones with alcohols. Angew. Chem., Int. Ed. 2015, 54, 14483−14486. (29) Fristrup, P.; Tursky, M.; Madsen, R. Mechanistic investigation of the iridium-catalysed alkylation of amines with alcohols. Org. Biomol. Chem. 2012, 10, 2569−2577. (30) Jiménez, M. V.; Fernández-Tornos, J.; Modrego, F. J.; PérezTorrente, J. J.; Oro, L. A. Oxidation and β-alkylation of alcohols catalysed by iridium(I) complexes with functionalised N-heterocyclic carbene ligands. Chem. - Eur. J. 2015, 21, 17877−17889. (31) Bartoszewicz, A.; González Miera, G.; Marcos, R.; Norrby, P.O.; Martin-Matute, B. Mechanistic studies on the alkylation of amines with alcohols catalyzed by a bifunctional iridium complex. ACS Catal. 2015, 5, 3704−3716. (32) Zhang, D.-D.; Chen, X.-K.; Liu, H.-L.; Huang, X.-R. DFT study on the iridium-catalyzed multi-alkylation of alcohol with ammonia. RSC Adv. 2016, 6, 87362−87372. (33) Li, F.; Xie, J.; Shan, H.; Sun, C.; Chen, L. General and efficient method for direct N-monomethylation of aromatic primary amines with methanol. RSC Adv. 2012, 2, 8645−8652. (34) Dang, T. T.; Ramalingam, B.; Seayad, A. M. Efficient rutheniumcatalyzed N-methylation of amines using methanol. ACS Catal. 2015, 5, 4082−4088. (35) Chianese, A. R.; Li, X.; Janzen, M. C.; Faller, J. W.; Crabtree, R. H. Rhodium and iridium complexes of N-heterocyclic carbenes via transmetalation: structure and dynamics. Organometallics 2003, 22, 1663−1667. (36) Hahn, F. E. Heterocyclic carbenes. Angew. Chem., Int. Ed. 2006, 45, 1348−1352. (37) Hahn, F. E.; Jahnke, M. C. Heterocyclic carbenes: synthesis and coordination chemistry. Angew. Chem., Int. Ed. 2008, 47, 3122−3127. (38) Würtz, S.; Glorius, F. Surveying sterically demanding Nheterocyclic carbene ligands with restricted flexibility for palladiumcatalyzed cross-coupling reactions. Acc. Chem. Res. 2008, 41, 1523− 1533. (39) Song, G.; Su, Y.; Periana, R. A.; Crabtree, R. H.; Han, K.; Zhang, H.; Li, X. Anion-exchange-triggered 1, 3-shift of an NH proton to iridium in protic N-heterocyclic carbenes: hydrogen-bonding and ionpairing effects. Angew. Chem., Int. Ed. 2010, 49, 912−917. (40) Biju, A. T.; Kuhl, N.; Glorius, F. Extending NHC-catalysis: coupling aldehydes with unconventional reaction partners. Acc. Chem. Res. 2011, 44, 1182−1195. (41) Hashmi, A. S. K. Sub-nanosized gold catalysts. Science 2012, 338, 1434−1434. (42) Jaimes, M. C. B.; Böhling, C. R. N.; Serrano-Becerra, J. M.; Hashmi, A. S. K. Highly active mononuclear NAC−gold (I) catalysts. Angew. Angew. Chem., Int. Ed. 2013, 52, 7963−7966. (43) Zhang, L.; Hou, Z. N-Heterocyclic carbene (NHC)−coppercatalysed transformations of carbon dioxide. Chem. Sci. 2013, 4, 3395− 3403. (44) Hopkinson, M. N.; Richter, C.; Schedler, M.; Glorius, F. An overview of N-heterocyclic carbenes. Nature 2014, 510, 485−496. (45) Sharninghausen, L. S.; Campos, J.; Manas, M. G.; Crabtree, R. H. Efficient selective and atom economic catalytic conversion of glycerol to lactic acid. Nat. Commun. 2014, 5, 5084−5092. (46) Manas, M. G.; Campos, J.; Sharninghausen, L. S.; Lin, E.; Crabtree, R. H. Selective catalytic oxidation of sugar alcohols to lactic acid. Green Chem. 2015, 17, 594−600.

(47) Levin, E.; Ivry, E.; Diesendruck, C. E.; Lemcoff, N. G. Water in N-heterocyclic carbene-assisted catalysis. Chem. Rev. 2015, 115, 4607− 4692. (48) Nasr, A.; Winkler, A.; Tamm, M. Anionic N-heterocyclic carbenes: Synthesis, coordination chemistry and applications in homogeneous catalysis. Coord. Chem. Rev. 2016, 316, 68−124. (49) Zhao, D.; Candish, L.; Paul, D.; Glorius, F. N-heterocyclic carbenes in asymmetric hydrogenation. ACS Catal. 2016, 6, 5978− 5988. (50) Campos, J.; Sharninghausen, L. S.; Manas, M. G.; Crabtree, R. H. Methanol dehydrogenation by iridium N-heterocyclic carbene complexes. Inorg. Chem. 2015, 54, 5079−5084. (51) (a) Alonso, F.; Riente, P.; Yus, M. Transfer hydrogenation of olefins catalysed by nickel nanoparticles. Tetrahedron 2009, 65, 10637−10643. (52) Indra, A.; Maity, P.; Bhaduri, S.; Lahiri, G. K. Chemoselective hydrogenation and transfer hydrogenation of olefins and carbonyls with the cluster-derived ruthenium nanocatalyst in water. ChemCatChem 2013, 5, 322−330. (53) Mai, V. H.; Nikonov, G. I. Transfer hydrogenation of nitriles, olefins, and N-heterocycles catalyzed by an N-heterocyclic carbenesupported half-sandwich complex of ruthenium. Organometallics 2016, 35, 943−949. (54) Li, J.-Q.; Andersson, P. G. Room temperature and solvent-free iridium-catalyzed selective alkylation of anilines with alcohols. Chem. Commun. 2013, 49, 6131−6133. (55) Elangovan, S.; Neumann, J.; Sortais, J.-B.; Junge, K.; Darcel, C.; Beller, M. Efficient and selective N-alkylation of amines with alcohols catalysed by manganese pincer complexes. Nat. Commun. 2016, 7, 12641−12648. (56) Yan, L.; Liu, X.-X.; Fu, Y. N-Alkylation of amines with phenols over highly active heterogeneous palladium hydride catalysts. RSC Adv. 2016, 6, 109702−109705. (57) Dai, L.-X. Chiral metal-organic assemblies-a new approach to immobilizing homogeneous asymmetric catalysts. Angew. Chem., Int. Ed. 2004, 43, 5726−5729. (58) Ding, K.; Wang, Z.; Wang, X.; Liang, Y.; Wang, X. SelfSupported chiral catalysts for heterogeneous enantioselective reactions. Chem. - Eur. J. 2006, 12, 5188−5197. (59) Williams, K. A.; Boydston, A. J.; Bielawski, C. W. Main-chain organometallic polymers: synthetic strategies, applications, and perspectives. Chem. Soc. Rev. 2007, 36, 729−744. (60) Wang, Z.; Chen, G.; Ding, K. Self-Supported catalysts. Chem. Rev. 2009, 109, 322−359. (61) Sun, Z.; Liu, Y.; Chen, J.; Huang, C.; Tu, T. Robust iridium coordination polymers: highly selective, efficient, and recyclable catalysts for oxidative conversion of glycerol to potassium lactate with dihydrogen liberation. ACS Catal. 2015, 5, 6573−6578. (62) Sun, Z.; Chen, J.; Tu, T. NHC-based coordination polymers as solid molecular catalysts for reductive amination of biomass levulinic acid. Green Chem. 2017, 19, 789−794. (63) Liu, Y.; Sun, Z.; Huang, C.; Tu, T. Efficient hydrogenation of biomass oxoacids to lactones by using NHC−iridium coordination polymers as solid molecular catalysts. Chem. - Asian J. 2017, 12, 355− 360. (64) Kang, B.; Fu, Z.; Hong, S. H. Ruthenium-catalyzed redoxneutral and single-step amide synthesis from alcohol and nitrile with complete atom economy. J. Am. Chem. Soc. 2013, 135, 11704−11707. (65) Torres, P. U. Cinacalcet HCl: a novel treatment for secondary hyperparathyroidism caused by chronic kidney disease. J. Renal Nutr. 2006, 16, 253−258. (66) Radu, I.-I.; Poirier, D.; Provencher, L. New efficient pathway for the synthesis of 3-aminoestrone. Tetrahedron Lett. 2002, 43, 7617− 7619. (67) Schön, U.; Messinger, J.; Buchholz, M.; Reinecker, U.; Thole, H.; Prabhu, M. K. S.; Konda, A. An improved synthesis of 3aminoestrone. Tetrahedron Lett. 2005, 46, 7111−7115. (68) Ogawa, S.; Obora, Y. Iridium-catalyzed selective α-methylation of ketones with methanol. Chem. Commun. 2014, 50, 2491−2493. 11750

DOI: 10.1021/acssuschemeng.7b03246 ACS Sustainable Chem. Eng. 2017, 5, 11744−11751

Research Article

ACS Sustainable Chemistry & Engineering (69) Quan, X.; Kerdphon, S.; Andersson, P. G. C-C coupling of ketones with methanol catalyzed by a N-heterocyclic carbinephosphine iridium complex. Chem. - Eur. J. 2015, 21, 3576−3579. (70) Neumann, J.; Elangovan, S.; Spannenberg, A.; Junge, K.; Beller, M. Improved and general manganese-catalyzed N-methylation of aromatic amines using methanol. Chem. - Eur. J. 2017, 23, 5410−5413. (71) Bruneau-Voisine, A.; Wang, D.; Dorcet, V.; Roisnel, T.; Darcel, C.; Sortais, J.-B. Transfer hydrogenation of carbonyl derivatives catalyzed by an inexpensive phosphine-free manganese precatalyst. J. Catal. 2017, 347, 57−62. (72) Paul, B.; Shee, S.; Chakrabarti, K.; Kundu, S. Tandem transformation of nitro compounds into N-methylated amines: Greener strategy for the utilization of methanol as a methylating agent. ChemSusChem 2017, 10, 2370−2374. (73) Chakraborty, S.; Piszel, P. E.; Hayes, C. E.; Baker, R. T.; Jones, W. D. Highly selective formation of n-Butanol from ethanol through the Guerbet process: A tandem catalytic approach. J. Am. Chem. Soc. 2015, 137, 14264−14267. (74) Ho, C. R.; Shylesh, S.; Bell, A. T. Mechanism and kinetics of ethanol coupling to butanol over hydroxyapatite. ACS Catal. 2016, 6, 939−94. (75) Sun, Z.; Vasconcelos, A. C.; Bottari, G.; Stuart, M. C. A.; Bonura, G.; Cannilla, C.; Frusteri, F.; Barta, K. Efficient catalytic conversion of ethanol to 1-Butanol via the guerbet reaction over Copper- and Nickel-doped porous. ACS Sustainable Chem. Eng. 2017, 5, 1738−1746.

11751

DOI: 10.1021/acssuschemeng.7b03246 ACS Sustainable Chem. Eng. 2017, 5, 11744−11751