Vous êtes sur la page 1sur 4

Communication

Cite This: J. Am. Chem. Soc. 2018, 140, 2024−2027 pubs.acs.org/JACS

Asymmetric Synthesis of Chiral Primary Amines by Ruthenium-


Catalyzed Direct Reductive Amination of Alkyl Aryl Ketones with
Ammonium Salts and Molecular H2
Xuefeng Tan,† Shuang Gao,† Weijun Zeng,† Shan Xin,† Qin Yin,*,†,‡ and Xumu Zhang*,†

Department of Chemistry, Southern University of Science and Technology, Shenzhen 518000, People’s Republic of China

Academy for Advanced Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen 518000, People’s
Republic of China
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via PERPUSTAKAAN UNIV INDONESIA on March 25, 2019 at 09:36:24 (UTC).

Scheme 1. Representative Strategies for Transition Metal-


ABSTRACT: A ruthenium/C3-TunePhos catalytic system Catalyzed ARA
has been identified for highly efficient direct reductive
amination of simple ketones. The strategy makes use of
ammonium acetate as the amine source and H2 as the
reductant and is a user-friendly and operatively simple
access to industrially relevant primary amines. Excellent
enantiocontrol (>90% ee for most cases) was achieved
with a wide range of alkyl aryl ketones. The practicability
of this methodology has been highlighted by scalable
synthesis of key intermediates of three drug molecules.
Moreover, an improved synthetic route to the optimal
diphosphine ligand C3-TunePhos is also presented.

C hiral amines are prevalent structural units in a large


number of pharmaceutical drug molecules, agrochemicals,
and commodity chemicals. Therefore, efficient synthetic routes
toward chiral amines have attracted tremendous attention.1
Transition metal-catalyzed reductive amination represents a
step-economic and efficient strategy by directly converting
carbonyl compounds into amines in the presence of amine
sources and reductants.2 However, compared to numerous also documented. In all cases, a functional group such as ester
studies on asymmetric imine reduction,3 only limited success or amide in the β-position is necessary to ensure high efficiency
has been achieved regarding asymmetric reductive amination (eq 2, Scheme 1). In comparison, Kadyrov’s group reported a
(ARA), partly due to the incompatibility of transition-metal ruthenium-catalyzed enantioselective reductive amination of
hydrides with ketones. Another factor that cannot be ignored simple aryl ketones via transfer hydrogenation. However, the
lies in coordination of the amine to the metal catalyst, often formation of undesired formylated amine could not be avoided,
resulting in its inhibition.4 To enhance the catalytic efficiency, thus requiring an additional hydrolysis step to achieve the
some special amines such as aryl amines,5 benzylamines6 or primary amine product (eq 3, Scheme 1).14 Despite the above-
others7 were used as amine sources, but these furnish secondary mentioned state-of-the-art, a highly efficient and enantioselec-
amines. An additional synthetic operation to remove the aryl or tive asymmetric reductive amination of simple aryl ketones with
benzyl substituents is thus required to liberate the synthetically ammonia or its equivalents using molecular H2 has remained
more versatile primary amine (eq 1, Scheme 1). Asymmetric rare15 and is a formidable challenge to be overcome (eq 4,
reductive amination for the direct synthesis of chiral primary Scheme 1). Herein, we report our recent efforts toward
amines from ketones and ammonia or its equivalents is highly asymmetric reductive amination of simple ketones with
desirable but still underdeveloped. In 2005, Takasago company ammonium salts utilizing molecular H2 as the reducing reagent.
patented an asymmetric reductive amination of β-keto esters We commenced our study by testing the performance of
with NH4OAc to provide chiral β-amino esters.8 After that, various combinations of ruthenium precatalysts and different
several works using a similar strategy have been reported, chiral diphosphine ligands in the reaction of model substrate
including the application of different ligands such as acetophenone (5a) and ammonium salts under 55 bar of H2.
ClMeOBIPHEP9 and segphos.10 The syntheses of top-selling
drug sitagliptin,11 (S)-3-amino-4-methoxy-butan-1-ol12 as well Received: December 6, 2017
as methyl (3R)-3-aminobutyrate13 via this methodology were Published: January 29, 2018

© 2018 American Chemical Society 2024 DOI: 10.1021/jacs.7b12898


J. Am. Chem. Soc. 2018, 140, 2024−2027
Journal of the American Chemical Society Communication

From an initial screening of diverse diphosphine ligands, C3- Table 1. Reaction Condition Optimizationa
TunePhos (4), a derivative of the Cn-TunePhos (n = 0−6)
series developed by our group in 2000,16 emerged as optimal
regarding enantiocontrol. Existing procedures17 to synthesize
these privileged ligands suffer from long synthetic routes and
limited possibilities for derivatization. An improved and
divergent preparation to five C3-TunePhos derivatives 4a−4e Entry Ru cat. Yield (%) ee (%)
is shown in Scheme 2 (for details, see the Supporting 1 RuCl2(4a)(DMF)n 90 78
Information). 2 RuCl2(4b)(DMF)n 88 74
3 RuCl2(4c)(DMF)n 87 76
Scheme 2. Improved Synthetic Route to C3-TunePhosa 4 RuCl2(4d)(DMF)n 93 89
5 RuCl2(4e)(DMF)n 94 94
6 Ru(OAc)2(4e) 92 97
7a Ru(OAc)2(4e) 96 96
8 Ru(OAc)2(4a) 93 75
9 Ru((R)-BINAP)(OAc)2 92 65
10 Ru((R)-SegPhos)(OAc)2 92 67
a
Reaction conditions: acetophenone 5a (0.2 mmol), NH4OAc (0.4
mmol), [Ru] (1 mol %), TFE (0.4 mL), H2 (55 bar), 80 °C, 24 h. The
pressure refers to the actual value at 80 °C. The free amine 6a was
obtained after neutralizing its corresponding acetate salt with a base.
The ee values were determined by chiral HPLC after sample acylation.
b
Reaction was carried out with 0.5 mol % catalyst loading at 100 °C.
a
Reaction conditions: (a) (2R,4R)-pentane-2,4-diol, DIAD, Ph3P, TFE = trifluoroethanol.
81%; (b) n-BuLi; then CuCN; then p-benzoquinone, one pot, 55%;
(c) KOH, then NaNO2 and KI, 40% over two steps; (d) n-BuLi,
ClPAr2, 37−61%. 5b−g) and electron-withdrawing substituents (F, Cl, Br, CF3 as
in 5h−5o) on the benzene ring were tolerated, and the
reactivity and selectivity were high throughout. The hetero-
With these C3-TunePhos ligands in hand, further condition aromatic ketone 5q reacted with decreased enantiomeric excess
optimization was conducted to improve the reaction efficiency probably due to a weak coordination effect of the oxygen atom.
and selectivity, including the screening of the ruthenium source, Besides methyl ketone, ethyl ketone 5r demonstrated
solvent, additive, ammonium salt, substrate concentration, comparable reactivity and selectivity, providing 6r in 78%
reaction temperature as well as hydrogen pressure with the yield and 97% ee. It is worth mentioning that catalyst
standard substrate acetophenone (5a) (Tables S1−S6). Finally, Ru(OAc)2(4e) showed very low reactivity toward 1-benzosu-
the combination of 1 mol % Ru(OAc)2(4e) as catalyst, berone (5s), whereas good conversion and enantiocontrol of 5s
trifluoroethanol (TFE) as solvent, two equivalents of could be achieved by using Ru(OAc)2 (4a) as catalyst.
ammonium acetate as amine source, 0.5 M substrate Unfortunately, the dialkyl ketone 5t was not a suitable substrate
concentration, 55 bar of H2 at 80 °C stood out, providing for this catalytic system, and poor yield and enantiocontrol
the corresponding primary amine 6a in 92% yield and 97% ee were obtained (20% ee, 6t).
(entry 6, Table 1). It is noteworthy that excellent conversion is To demonstrate the significance and practicality of this
obtained when TFE is used as solvent, and possible reason is methodology, scale-up syntheses of key intermediates of drug
that TFE facilitates the formation of imine intermediate.9 The molecules, including Tecalcet hydrochloride, Cinacalcet, and
formation of secondary alcohol could be inhibited through Rivastigmine were performed (Scheme 3). Tecalcet hydro-
choosing suitable ammonium salts in TFE (see Tables S2 and chloride and Cinacalcet are drugs that act as a calcimimetic and
S3 for details). Interestingly, we did not observe the formation used to treat secondary hyperparathyroidism, the former
of secondary amine arising from double reductive aminations. A currently in Phase II clinical trials18 and the latter sold by
comparison of C3-TunePhos 4a with commercially available Amgen under the trade name of Sensipar in North America and
diphosphine ligands binap and segphos is summarized in Table Australia, and of Mimpara in Europe.19 The critical synthon to
1 (entry 8 vs entries 9 and 10), This demonstrates the Tecalcet hydrochloride is (R)-1-(3-methoxyphenyl)ethyl amine
superiority of C3-TunePhos in this reaction. When the reaction (6g), which could be easily accessed under standard conditions
temperature was increased from 80 to 100 °C, similar results with high enantioselectivity (94% ee) and TON up to 500 (eq
were obtained in the presence of only 0.5 mol % catalyst 1, Scheme 3). Similarly, the key chiral motif (R)-1-(naphth-1-
loading (entry 7). Further substrate scope investigation was yl) ethyl amine in Cinacalcet could be efficiently synthesized
then carried out at 100 °C with 0.5 mol % of catalyst. The with high enantioselectivity (up to 98% ee) and promising
absolute configuration of the product 6a was established by efficacy of S/C up to 1000 (eq 2, Scheme 3). The gram-scale
comparison of its optical rotation with that reported in syntheses of these two key intermediates reveal the potential of
literature (see the Supporting Information for details). practical industrial application. Rivastigmine, sold under the
Under the optimal reaction conditions, a wide range of trade name of Exelon, is a parasympathomimetic or cholinergic
simple aryl ketones were subjected to this direct reductive agent for the treatment of mild to moderate dementia of the
amination procedure. In most cases, the ketones could be Alzheimer’s type and dementia due to Parkinson’s disease.20 As
transformed into the corresponding primary amines with shown in eq 3 of Scheme 3, the key intermediate to (R)-
excellent enantioselectivities and moderate to high yields Rivastigmine was directly synthesized via reductive amination
(Table 2). Both electron-donating (Me, tBu or MeO as in of 5v with high efficiency.21
2025 DOI: 10.1021/jacs.7b12898
J. Am. Chem. Soc. 2018, 140, 2024−2027
Journal of the American Chemical Society Communication

Table 2. Substrate Scope molecular H2. The use of ammonia acetate as an amine source
and H2 as reductant constitutes a facile and user-friendly
approach to versatile primary amines, which can be easily
derived into more valuable amine products. This reaction
features broad substrate scope, good functional group
compatibility and excellent enantiocontrol (up to 98% ee).
Key intermediates of three drugs can be easily accessed in gram
scale from commercially available ketones, which showcases the
potential of practical usage. In addition, an improved synthetic
route to the optimal diphosphine ligand C3-TunePhos was also
developed.


*
ASSOCIATED CONTENT
S Supporting Information
The Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/jacs.7b12898.
Detailed experimental procedures, spectral data, and
analytical data (PDF)

■ AUTHOR INFORMATION
Corresponding Authors
*yinq@sustc.edu.cn
*zhangxm@sustc.edu.cn
ORCID
Xuefeng Tan: 0000-0002-6121-1499
Qin Yin: 0000-0003-3534-3786
Xumu Zhang: 0000-0001-5700-0608
Notes
The authors declare no competing financial interest.

a
■ ACKNOWLEDGMENTS
We are grateful to a start-up fund from Southern University of
Reaction condtions: ketone 5 (0.2 mmol), NH4OAc (0.4 mmol), Science and Technology, Shenzhen Science and Technology
Ru(OAc)2(4e) (0.5 mol %), TFE (0.4 mL), H2 (57 bar), 100 °C, 24 h. Innovation Committee (No. KQTD20150717103157174) and
The pressure refers to the actual value at 100 °C. The free amines 6a-t National Natural Science Foundation of China (No. 21432007)
were obtained after neutralizing their corresponding acetate salts with
a base. The ee values were determined by chiral HPLC after sample
for financial support. We greatly acknowledge Professor Martin
acylation. bRu(OAc)2(4a) was used instead of Ru(OAc)2(4e). cThe ee Oestreich (Technische Universität Berlin) for proofreading the
paper.


was determined by chiral HPLC after sample tosylation.

Scheme 3. Scale-up Syntheses of Key Intermediates of Drug REFERENCES


Molecules (1) (a) Blaser, H.-U.; Malan, C.; Pugin, B.; Spindler, F.; Steiner, H.;
Studer, M. Adv. Synth. Catal. 2003, 345, 103−151. (b) Breuer, M.;
Ditrich, K.; Habicher, T.; Hauer, B.; Keßeler, M.; Stürmer, R.; Zelinski,
T. Angew. Chem., Int. Ed. 2004, 43, 788−824. (c) Ager, D. In Science of
Synthesis: Stereoselective Synthesis; de Vries, J. G., Ed.; Thieme:
Stuttgart, 2010. (d) Stereoselective Formation of Amines; Li, W.,
Zhang, X., Eds; Springer: Berlin, 2014.
(2) Wang, C.; Xiao, J. Top. Curr. Chem. 2013, 343, 261−282.
(3) For selected recent reviews on TM-catalyzed asymmetric imine
hydrogenation, see: (a) Xie, J. H.; Zhu, S. F.; Zhou, Q. L. Chem. Rev.
2011, 111, 1713−1760. (b) Xiao, J.; Tang, W. Synthesis 2014, 46,
1297−1302.
(4) Nugent, T. C.; El-Shazly, M. Adv. Synth. Catal. 2010, 352, 753−
819.
(5) (a) Blaser, H.-U.; Buser, H.-P.; Jalett, H.-P.; Pugin, B.; Spindler,
F. Synlett 1999, 1999, 867−868. (b) Chi, Y.; Zhou, Y.-G.; Zhang, X. J.
Org. Chem. 2003, 68, 4120−4122. (c) Li, C.; Villa-Marcos, B.; Xiao, J.
J. Am. Chem. Soc. 2009, 131, 6967−6969. (d) Rubio-Pérez, L.; Pérez-
Flores, F. J.; Sharma, P.; Velasco, L.; Cabrera, A. Org. Lett. 2009, 11,
265−268. (e) Villa-Marcos, B.; Li, C.; Mulholland, K. R.; Hogan, P. J.;
In conclusion, we have established a highly chemo- and Xiao, J. Molecules 2010, 15, 2453. (f) Wang, C.; Pettman, A.; Basca, J.;
enantioselective catalytic system for the direct reductive Xiao, J. Angew. Chem., Int. Ed. 2010, 49, 7548−7552. (g) Zhou, S.;
amination of simple ketones with ammonium salts and Fleischer, S.; Jiao, H.; Junge, K.; Beller, M. Adv. Synth. Catal. 2014,

2026 DOI: 10.1021/jacs.7b12898


J. Am. Chem. Soc. 2018, 140, 2024−2027
Journal of the American Chemical Society Communication

356, 3451−3455. (h) Yang, P.; Lim, L. H.; Chuanprasit, P.; Hirao, H.;
Zhou, J. Angew. Chem., Int. Ed. 2016, 55, 12083−12087. (i) Huang, H.;
Wu, Z.; Gao, G.; Zhou, L.; Chang, M. Org. Chem. Front. 2017, 4,
1976−1980. (j) Li, B.; Zheng, J.; Zeng, W.; Li, Y.; Chen, L. Synthesis
2017, 49, 1349−1355.
(6) (a) Tararov, V. I.; Kadyrov, R.; Riermeier, T. H.; Borner, A.
Chem. Commun. 2000, 1867−1868. (b) Kadyrov, R.; Riermeier, T. H.;
Dingerdissen, U.; Tararov, V.; Börner, A. J. Org. Chem. 2003, 68,
4067−4070.
(7) (a) Zhou, P.; Zhang, Z.; Jiang, L.; Yu, C.; Lv, K.; Sun, J.; Wang, S.
Appl. Catal., B 2017, 210, 522−532. (b) Huang, H.; Liu, X.; Zhou, L.;
Chang, M.; Zhang, X. Angew. Chem., Int. Ed. 2016, 55, 5309−5312.
(c) Zhou, H.; Liu, Y.; Yang, S.; Zhou, L.; Chang, M. Angew. Chem., Int.
Ed. 2017, 56, 2725−2729. (d) Williams, G. D.; Pike, R. A.; Wade, C.
E.; Wills, M. Org. Lett. 2003, 5, 4227−4230. (e) Strotman, N. A.;
Baxter, C. A.; Brands, K. M. J.; Cleator, E.; Krska, S. W.; Reamer, R. A.;
Wallace, D. J.; Wright, T. J. J. Am. Chem. Soc. 2011, 133, 8362−8371.
(f) Chang, M.; Liu, S.; Huang, K.; Zhang, X. Org. Lett. 2013, 15,
4354−4357.
(8) Matsumura, K.; Saito, T. PCT Patent Appl. WO 2005/028419
A3, 2005.
(9) Bunlaksananusorn, T.; Rampf, F. Synlett 2005, 2682−2684.
(10) Shimizu, H.; Nagasaki, I.; Matsumura, K.; Sayo, N.; Saito, T. Acc.
Chem. Res. 2007, 40, 1385−1393.
(11) Steinhuebel, D.; Sun, Y.; Matsumura, K.; Sayo, N.; Saito, T. J.
Am. Chem. Soc. 2009, 131, 11316−11317.
(12) Mattei, P.; Moine, G.; Püntener, K.; Schmid, R. Org. Process Res.
Dev. 2011, 15, 353−359.
(13) Matsumura, K.; Zhang, X.; Hori, K.; Murayama, T.; Ohmiya, T.;
Shimizu, H.; Saito, T.; Sayo, N. Org. Process Res. Dev. 2011, 15, 1130−
1137.
(14) (a) Kadyrov, R.; Riermeier, T. H. Angew. Chem., Int. Ed. 2003,
42, 5472−5474. (b) Talwar, D.; Salguero, N. P.; Robertson, C. M.;
Xiao, J. Chem.−Eur. J. 2014, 20, 245−252.
(15) (a) Riermeier, T.; Haack, K.-J.; Dingerdissen, U.; Boerner, A.;
Tararov, V.; Kadyrov, R. U.S. Patent 6,884,887, 2005. (b) Gallardo-
Donaire, J.; Ernst, M.; Trapp, O.; Schaub, T. Adv. Synth. Catal. 2016,
358, 358−363. (c) During the preparation of this paper, Schaub and
co-workers disclosed an elegant Ru-catalyzed ARA of aryl ketones with
ammonia and H2, see: Gallardo-Donaire, J.; Hermsen, M.; Wysocki, J.;
Ernst, M.; Rominger, F.; Trapp, O.; Hashmi, A. S. K.; Schäfer, A.;
Comba, P.; Schaub, T. J. Am. Chem. Soc. 2018, 140, 355−361.
(16) Zhang, Z.; Qian, H.; Longmire, J.; Zhang, X. J. Org. Chem. 2000,
65, 6223−6226.
(17) (a) Qiu, L.; Kwong, F. Y.; Wu, J.; Lam, W. H.; Chan, S.; Yu, W.-
Y.; Li, Y.-M.; Guo, R.; Zhou, Z.; Chan, A. S. C. J. Am. Chem. Soc. 2006,
128, 5955−5965. (b) Sun, X.; Zhou, L.; Li, W.; Zhang, X. J. Org. Chem.
2008, 73, 1143−1146. (c) Sun, X.; Li, W.; Hou, G.; Zhou, L.; Zhang,
X. Adv. Synth. Catal. 2009, 351, 2553−2557.
(18) Yamazaki, N.; Atobe, M.; Kibayashi, C. Tetrahedron Lett. 2001,
42, 5029−5032.
(19) Nemeth, E. F.; Steffey, M. E.; Hammerland, L. G.; Hung, B. C.
P.; Van Wagenen, B. C.; DelMar, E. G.; Balandrin, M. F. Proc. Natl.
Acad. Sci. U. S. A. 1998, 95, 4040−4045.
(20) (a) Pan, Y.; Xu, X.; Wang, X. Br. J. Pharmacol. 2003, 140, 907−
912. (b) Farlow, M. R.; Cummings, J. L. Am. J. Med. 2007, 120, 388−
397.
(21) Fuchs, M.; Koszelewski, D.; Tauber, K.; Sattler, J.; Banko, W.;
Holzer, A. K.; Pickl, M.; Kroutil, W.; Faber, K. Tetrahedron 2012, 68,
7691−7694.

2027 DOI: 10.1021/jacs.7b12898


J. Am. Chem. Soc. 2018, 140, 2024−2027

Vous aimerez peut-être aussi