Free Essay

Copper-Catalysed Selective Hydroamination Reactions of Alkynes

In:

Submitted By beibeiao
Words 6397
Pages 26
ARTICLES
PUBLISHED ONLINE: 15 DECEMBER 2014 | DOI: 10.1038/NCHEM.2131

Copper-catalysed selective hydroamination reactions of alkynes
Shi-Liang Shi and Stephen L. Buchwald*
The development of selective reactions that utilize easily available and abundant precursors for the efficient synthesis of amines is a long-standing goal of chemical research. Despite the centrality of amines in a number of important research areas, including medicinal chemistry, total synthesis and materials science, a general, selective and step-efficient synthesis of amines is still needed. Here, we describe a set of mild catalytic conditions utilizing a single copper-based catalyst that enables the direct preparation of three distinct and important amine classes (enamines, α-chiral branched alkylamines and linear alkylamines) from readily available alkyne starting materials with high levels of chemo-, regio- and stereoselectivity. This methodology was applied to the asymmetric synthesis of rivastigmine and the formal synthesis of several other pharmaceutical agents, including duloxetine, atomoxetine, fluoxetine and tolterodine.

C

omplex organic molecules play a crucial role in the study and treatment of disease. The extent to which they can be utilized in these endeavours depends on the efficient and selective chemical methods for their construction1. Amines are widely represented in biologically active natural products and medicines2 (a small selection of which are shown in Fig. 1a). Consequently, the selective assembly of amines from readily available precursors is a prominent objective in chemical research3. There are a number of powerful methods that address this challenge, including metal-catalysed cross-coupling4,5, nucleophilic addition to imines6, C–H nitrogen insertion7 and enzymatic methods8,9. However, the direct production of amines from simple olefins or alkynes represents a highly attractive alternative, given the abundance and accessibility of these starting materials. For this reason, the addition of nitrogen and hydrogen across carbon–carbon multiple bonds (hydroamination) has long been pursued as a means to access amines10–12. Although much progress has been made, a generally effective strategy to achieve chemo-, regio- and enantioselective hydroamination of simple alkenes or alkynes remains elusive. We became interested in developing hydroamination reactions of alkynes as a convenient and powerful means of accessing aminated products (Fig. 1b). Reactions that employ alkynes as starting materials are synthetically versatile, because alkynes can be prepared by a variety of strategies, including Sonogashira coupling13, nucleophilic addition of metal acetylides14 and homologation of carbonyl groups15. In addition, one or both π-bonds of alkynes may be utilized, further increasing their flexibility as starting materials. For these reasons, the hydrofunctionalization of alkynes has recently become an active area of research16–22. We recently detailed23 (as have Hirano and Miura24) catalyst systems for the asymmetric hydroamination of styrenes that operate through the addition of a catalytic copper hydride species25–32 to a carbon–carbon double bond followed by carbon–nitrogen bond formation using an electrophilic nitrogen source33–36. We surmised that we could apply this approach to the selective functionalization of alkynes wherein alkyne hydrocupration would give a stereodefined vinylcopper intermediate. We anticipated that, in analogy to our previous work, direct interception of this intermediate would potentially enable the stereoselective formation of enamines (Fig. 1b, A). Enamines are versatile intermediates in organic synthesis, and although catalytic methods have been

developed for their synthesis by alkyne hydroamination, control of the regio- and stereochemistry of enamine formation is non-trivial16. In addition to enamine synthesis, we speculated on the possibility that conditions could be developed to convert alkynes to enantioenriched α-branched alkylamines and/or linear alkylamines in one synthetic operation (Fig. 1b, B). Such cascade processes are highly desirable in organic synthesis, because potentially difficult work-up and isolation steps can be avoided and the generation of chemical waste is minimized37. In particular, we envisioned a scenario in which the starting alkyne is initially reduced to a transiently formed alkene, which would then undergo hydroamination to afford the alkylamine product. If successful, this approach would be particularly attractive due to the ease and low cost of the Sonogashira process for the preparation of alkyne starting materials relative to the cross-coupling of stereodefined vinylmetal reagents or other routes used to access geometrically pure alkenes for hydroamination. We were aware that, mechanistically, the vinyl- and alkylcopper intermediates in the proposed process are required to react in a highly chemoselective manner (Fig. 1c). Specifically, the vinylcopper species formed upon hydrocupration of the alkyne would need to be selectively intercepted by the proton source in the presence of the aminating reagent to furnish the intermediate alkene, while the alkylcopper species formed upon hydrocupration of this alkene would need to selectively engage the electrophilic nitrogen source in the presence of a proton donor to ultimately furnish the desired alkylamine product. Although both steps (that is, alkyne semireduction38–40 and alkene hydroamination23) are well precedented, the ability to achieve the desired selectivity in one pot via a cascade sequence has never been demonstrated41,42. Here, we report mild and scalable conditions for the highly chemo-, regio- and stereoselective synthesis of enamines (‘direct hydroamination’) and alkylamines (‘reductive hydroamination’) products from alkynes, using a single copper catalyst system.

Results and discussion
Direct hydroamination: development and scope. To assess the feasibility of the outlined alkyne hydroamination (Fig. 1b, A), we treated 1,2-diphenylacetylene (1a) with N,N-dibenzyl-Obenzoylhydroxylamine (2a, 1.2 equiv.) and an excess of diethoxymethylsilane (3) in the presence of 2 mol% copper

Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA. * e-mail: sbuchwal@mit.edu
38
NATURE CHEMISTRY | VOL 7 | JANUARY 2015 | www.nature.com/naturechemistry

© 2014 Macmillan Publishers Limited. All rights reserved.

NATURE CHEMISTRY a N

DOI: 10.1038/NCHEM.2131

ARTICLES
H N

HO S O O MeO OH N Quinine (malaria) Me Me OH O Me Et N H Atomoxetine (depression) Me Me N Me Rivastigmine (dementia) Me Me N O O Me H HO Morphine (pain) Duloxetine (depression) H N Me

c Required selectivity for reductive hydroamination cascade
Me R R1 R2 R R2N–OBz R1 Undesired amination R2 Vinylcopper intermediate Desired protonation

N

L*Cu ROH R1 R2

N

Me · Regioselective Me · Chemoselective · Stereoselective

L*Cu (Cat.) hydrosilane R1 ROH R2N–OBz R2

Tolterodine (urinary disorders)

b This work: divergent access to enamines and alkylamines
R
1

N

R A R1 R2 R2 Alkyne B

R
1

N

R or R2

R

N

R

R R1

N

R R2N–OBz

L*Cu ROH R1 R2 Alkylcopper intermediate Undesired protonation R1 R2

R

R

R1

R2

Desired amination

Direct hydroamination

Reductive hydroamination

Figure 1 | Bioactive amines, the synthesis of amines from alkynes, and the reductive hydroamination cascade strategy. a, Representative alkaloids and drugs demonstrating the ubiquitous nature of amines in bioactive organic molecules. b, Catalytic hydroamination of alkynes to generate three product classes. c, Contrasting reactivities of vinyl- and alkylcopper intermediates required to achieve a reductive hydroamination cascade.

acetate and a range of phosphine ligands. A number of ligands could be used to perform the direct hydroamination reaction, and the resulting enamine 4a was efficiently produced as a single geometric isomer, as determined by 1H NMR analysis (Table 1, entries 1–4). Although copper catalysts based on 2,2′-bis(diphenylphosphino)-1,1′-binaphthalene (BINAP, L1), 4,5-bis(diphenylphosphino)-9,9-dimethylxanthene (Xantphos, L2) or 4,4′-bi-1,3-benzodioxole-5,5′-diylbis(diphenylphosphane) (SEGPHOS, L3) were effective in this context, the catalyst based on 5,5′-bis[di(3,5-di-tert-butyl-4-methoxyphenyl)phosphino]-4,4′-bi-1,3benzodioxole (DTBM-SEGPHOS, L4) was found to be the most efficient and generally applicable. We then evaluated the substrate scope of this reaction and, as shown in entries 5–9, a diverse range of aryl-substituted internal alkyne substrates could be converted to the corresponding (E)-enamines 4 with complete stereoselectivity (4b–4e; 80–99%). Notably, sterically hindered amines, which were problematic substrates for previously reported hydroamination reactions of alkynes43, could be successfully transformed using the current conditions (4b and 4d). More importantly, direct hydroamination of unsymmetrical internal alkynes occurred with excellent regioselectivity (4c–4e; >19:1). In addition, we found that a 1,2-dialkylacetylene was left intact under these conditions (4e) and pharmaceutically important heterocycles, including morpholine (4c), thiophene (4d), piperidine (4e) and pyrimidine (4e), were well-tolerated. Although the direct hydroamination of terminal alkynes to construct monosubstituted enamines was not successful, the current method represents a rare example of a highly regio- and stereoselective hydroamination of internal alkynes for the construction of dialkyl enamines43. Reductive hydroamination: development and scope. As previously described, we were hopeful that the addition of a protic additive could divert this reaction from direct alkyne hydroamination to the outlined reductive hydroamination by
NATURE CHEMISTRY | VOL 7 | JANUARY 2015 | www.nature.com/naturechemistry

selective protonation of the formed vinylcopper intermediate (Fig. 1c). Indeed, inclusion of methanol as an additive under the reaction conditions in Table 1 resulted in the formation of the desired reductive hydroamination product 5a in moderate yield, along with a significant amount of enamine 4a (18%) and stilbene (17%) as side products (Table 2, entry 1). Fortunately, an evaluation of alcohol additives revealed that ethanol was a suitable proton source, which minimized the formation of these side products to afford benzylamine 5a in excellent yield and high enantioselectivity (entry 2, 92% yield, 89% e.e.). Interestingly, in contrast to the direct alkyne hydroamination protocol for enamine formation, L4 was uniquely able to perform the reductive hydroamination cascade reaction. Reaction utilizing copper catalysts based on L1, L2 or L3 provided only enamine 4a in high yields, even in the presence of ethanol (entries 4–6). We attribute the success of the catalyst system based on L4 to the ability of the CuH species to hydrocuprate alkynes and alkenes more rapidly. In contrast, the hydrocupration of alkynes occurred less efficiently when L1–L3 were used, resulting in consumption of the alcohol additive by the CuH before alkyne hydrocupration could take place. Therefore, only the enamine product was obtained in these cases. In addition, we found that arylacetylenes could also undergo reductive hydroamination, although in the case of these substrates, isopropanol was a superior protic additive (entry 8). Under the optimized set of reaction conditions, a range of chiral benzylamine derivatives could be prepared in moderate to high yield (61–85%) with very high levels of enantioselectivity (≥97% e.e., Table 3). These mild catalytic conditions tolerated a range of common functional groups including ethers (5c, 5h), alcohols (5i), aryl halides (5e, 5f ), pyridines (5d), indoles (5g), acetals (5j) and ketals (5m, 5n). Moreover, a reaction conducted on a 10 mmol scale proceeded efficiently in the presence of 1 mol% catalyst, furnishing the product in undiminished yield and enantioselectivity (5j). The applicability of new synthetic methods to the late-stage
39

© 2014 Macmillan Publishers Limited. All rights reserved.

ARTICLES
Cu(OAc)2 (2.0 mol%) R1 Ar Aryl alkyne 1 + R2 N R3 ligand (2.2 mol%)

NATURE CHEMISTRY

DOI: 10.1038/NCHEM.2131

Table 1 | Optimization and scope of copper-catalysed direct hydroamination of aryl alkynes.
R2 N R3

OBz Electrophilic amine 2

Ar HSiMe(OEt)2 (3, 3.0 equiv.) THF, 45 °C, 18 h R1 (E)-enamine 4

Entry 1 2 3 4 5*
Ph

1

2

Ligand rac-L1

4

Yield 90%

Ph

Bn

N

Bn

L2
Ph

Bn N Bn

95% 87% 99% 99% (88%)
L1 (BINAP) PPh2 PPh2

OBz

rac-L3 rac-L4 rac-L4
Me Me Ph

1a

2a

4a

Ph

Me Me N OBz

Me Me

6*

Ph

rac-L4

Ph

N Me Me

99% (93%)†
Me Me O

PPh2

1a

Ph

2b
O Ph Ph N OBz

4b
O N

PPh2

L2 (Xantphos)

7*

nBu

rac-L4 nBu 97%‡ 4c

(94%)
O O PPh2 PPh2

1b

2c
S S iPr iPr

O

8*

N

iPr

N

NC

OBz

rac-L4

iPr

NC

80%‡ (70%)

O L3 (SEGPHOS)

1c

2d

4d
N N N N N O O PAr2 PAr2

Ph

9* nBu N BzO N

rac-L4 nBu Ph

N

90%‡ (85%)

O O

1d

2e

4e

L4 (DTBM-SEGPHOS) Ar = 3,5-(tBu)2-4-MeO-C6H2

Conditions: 1a (0.2 mmol), 2a (0.24 mmol), 3 (0.6 mmol), Cu(OAc)2 (2.0 mol%), ligand (2.2 mol%), THF (1 M), 45 °C, 18 h. Yields and stereoselectivities were determined by 1H NMR analysis using 1,1,2,2-tetrachloroethane as an internal standard. In all cases only (E)-enamine products were observed. *Conditions: 1 (1.0 mmol), 2 (1.2 mmol), 3 (3.0 mmol), Cu(OAc)2 (2.0 mol%), rac-L4 (2.2 mol%), THF (1 M), 45 °C, 18 h. Isolated yields of products after reduction are given in parentheses (average of two runs). †Isolated yield of enamine after purification by flash column chromatography. ‡Major regioisomers are shown; in all cases, <5% of the minor regioisomer was observed as determined by 1H NMR analysis of the crude enamine product.

modification of complex natural products is a highly desirable feature, as analogues of bioactive molecules can be prepared without the need for de novo synthesis. Accordingly, readily available alkynes derived from the natural products δ-tocopherol and oestrone were subjected to asymmetric reductive hydroamination conditions to afford aminated products with good yields and excellent catalyst-controlled diastereoselectivities (d.r.: 99:1, 5k−5n). It is noteworthy that in all reductive hydroamination reactions employing aryl-substituted alkynes, the amination products were delivered with exclusive Markovnikov regioselectivity, with C–N bond formation occurring adjacent to the aryl group. In addition to aryl-substituted alkynes, we found that terminal aliphatic alkynes readily participate in catalytic reductive hydroamination to deliver alkylamines (Table 4). In contrast to aryl-substituted alkynes, anti-Markovnikov regioselectivity was
40

observed when simple alkylacetylene substrates were used, giving rise to linear tertiary amines in high yields (71–88% yield). We note that it was crucial to use a slight excess of isopropanol compared to the alkylacetylene substrate in the case of terminal alkyne substrates, probably due to deactivation of the catalyst through formation of a copper acetylide species when the amount of isopropanol was insufficient40. It is noteworthy that this methodology could be applied to a substrate bearing an unprotected secondary amine to provide 1,3-diamine 6a in high yield. Furthermore, alkynol silyl ethers were suitable substrates for the current method. Upon reductive hydroamination and silyl deprotection, 1,3-amino alcohol products were prepared in good yields (6f, 6g). An enantioenriched 1,3-amino alcohol could be generated from the optically active alkynol silyl ether (98% e.e.) without erosion of enantiomeric excess (6g, 98% e.e.). The use of
NATURE CHEMISTRY | VOL 7 | JANUARY 2015 | www.nature.com/naturechemistry

© 2014 Macmillan Publishers Limited. All rights reserved.

NATURE CHEMISTRY

DOI: 10.1038/NCHEM.2131

ARTICLES
Bn Bn

Table 2 | Optimization of reductive hydroamination.
Cu(OAc)2 (2.0 mol%) ligand (2.2 mol%) Ph Ph 1a 2a (1.2 equiv.) Ph HSiMe(OEt)2 (3, 4.0 equiv.) alcohol (1.5 equiv.) THF, 40 °C, 18 h Bn Conditions above nBu nBu

Bn

N

Bn

N

the current reductive hydroamination methodology, in conjunction with well-developed methods for the asymmetric synthesis of alkynols14, represents an attractive approach for the synthesis of these biologically relevant compounds. Application to drug synthesis. The functionalized amines obtained by the cascade hydroamination developed in this study have broad utility. To illustrate this point, 1,3-amino alcohol 6f can be readily transformed to the antidepressant drug duloxetine following the literature procedure44 (Fig. 2a). Moreover, 1,3-amino alcohol 6g is a key intermediate for the synthesis of the attention deficit hyperactivity disorder (ADHD) drug atomoxetine45, as well as the antidepressant drug fluoxetine46 (Fig. 2b). Furthermore, diisopropylamine 8 could be prepared by reductive hydroamination of readily synthesized alkyne 7 (Fig. 2c). Tolterodine, a drug used for symptomatic treatment of urinary incontinence, could be prepared by demethylation of 847. Our method was also applied to a new two-step synthesis of rivastigmine, a drug used to treat dementia (Fig. 2d). Condensation of commercially available reagents gave carbamate 9, which, upon asymmetric reductive hydroamination, furnished the target in an enantiopure form and good chemical yield. These syntheses exemplify the potential utility of the current hydroamination method for the rapid and efficient construction of medicinally relevant molecules. Mechanistic discussion. The proposed mechanisms for the formation of enamines and alkylamines (Fig. 3a) both commence with syn-selective Cu–H addition to the alkyne substrate to give vinylcopper species 11. In the absence of a proton source

+ Ph Ph 4a 5a Bn Ph

N

Me

1e

5b

Entry 1 2 3 4 5 6 7 8†

Substrate 1a 1a 1a 1a 1a 1a 1e 1e

Ligand (R)-L4 (R)-L4 (R)-L4 (R)-L1 L2 (R)-L3 (R)-L4 (R)-L4

Alcohol MeOH EtOH i PrOH EtOH EtOH EtOH EtOH i PrOH

Yield 4a 18% 2% 2% 83% 95% 80% –* –*

Yield 5a or 5b (e.e.) 60% (89% e.e.) 92% (89% e.e.) 83% (89% e.e.) 0 0 0 78% (99% e.e.) 83% (99% e.e.)

Conditions: 1a or 1e (0.2 mmol), 2a (0.24 mmol), 3 (0.8 mmol), alcohol (0.3 mmol), Cu(OAc)2 (2.0 mol%), ligand (2.2 mol%), THF (1 M), 40 °C, 18 h. Yields were determined by gas chromatography using dodecane as an internal standard. Enantiomeric excesses (e.e.) were determined by HPLC analysis using chiral stationary phases. *The corresponding enamines were not observed. †Conditions: 1e (1.1 mmol), 2a (1.0 mmol), 3 (4.0 mmol), i-PrOH (1.2 mmol), Cu(OAc)2 (2.0 mol%), (R)-L4 (2.2 mol%), THF (1 M), 40 °C, 18 h.

Table 3 | Scope of copper-catalysed reductive hydroamination of aryl alkynes.
2.0 mol % Cu(OAc)2 R2 + R1 Bn N Bn + Me EtO Si OEt H Hydrosilane 3 Bn Bn Bn Bn Cl Alcohol THF, 40 °C, 18 h Aryl alkyne 1 Bn Bn Ph n 2.2 mol % (R)-L4

R3 R1

N

R4

OBz Electrophilic amine 2a Bn Bn

R2 Chiral amine 5 Bn Bn

N

N

N

N

N

Me Bu MeO

Me N

Me

Me

5a, 85% yield, 89% e.e.*

5b, 75% yield, 99% e.e.

5c, 81% yield, >99% e.e.

5d, 63% yield, 98% e.e.

5e, 71% yield, 97% e.e.

Bn Bn N Bn N

Bn Bn Me N Bn Bn N Bn Bn N Bn OEt OEt

Me F 5f, 75% yield, 97% e.e.

OMe

OH

N Ts 5g, 80% yield, 98% e.e. 5h, 83% yield, 98% e.e.* 5i, 70% yield, 98% e.e.* 5j, 85% yield, 99% e.e.*,†

Bn

N

Bn

Bn

N

Bn H

Me

O

O H

Me

O

O

Me O Me

Me 3

Me

Me 3

Me 3 Me

Me O Me

Me 3

Me

Me 3

Me 3 Me Bn

Bn N Me

H

H Bn

Bn N Me

H

H

with (R)-L4: 5k, 75% yield, >99:1 d.r.

with (S)-L4: 5l, 76% yield, <1:99 d.r.

with (R)-L4: 5m, 62% yield, 99:1 d.r.

with (S)-L4: 5n, 61% yield, <1:99 d.r.

Conditions: 1 (1.1 mmol), 2 (1.0 mmol), 3 (4.0 mmol), i-PrOH (1.2 mmol), Cu(OAc)2 (2.0 mol%), (R)-L4 (2.2 mol %), THF (1 M), 40 °C, 18 h. Isolated yields are reported (average of two runs). Enantiomeric excesses (e.e.) and diastereomeric ratios (d.r.) were determined by HPLC analysis. *Using EtOH instead of i-PrOH. †Reaction performed on 10 mmol scale using 1 mol% of catalyst. See Supplementary Section 3 for details.

NATURE CHEMISTRY | VOL 7 | JANUARY 2015 | www.nature.com/naturechemistry

41

© 2014 Macmillan Publishers Limited. All rights reserved.

ARTICLES
Table 4 | Reductive hydroamination of alkylacetylenes.
2.0 mol% Cu(OAc)2 2.2 mol% rac-L4 H R1 Alkyne 1 2 (1.0 equiv.) iPrOH NATURE CHEMISTRY

DOI: 10.1038/NCHEM.2131

R3 R1 N R2

(1.2 equiv.) Alkyl amine 6

HSiMe(OEt)2 (3, 4.0 equiv.) THF, 40 °C, 18 h

N(H)Bn

Bn N Bn Bn Ph 5 N Bn tBu Bn N Bn

OMe

Bn N Bn

6a, 86% yield

6b, 76% yield

6c, 71% yield

6d, 88% yield

Ph

Me

Bn N Bn S OH 6f, 72% yield

Me N Bn OH 6g, 72% yield

Bn N Me

OTBS 6e, 73% yield

(from silyl ether)

(from silyl ether)

Conditions: 1 (1.1 mmol), 2 (1.0 mmol), 3 (4.0 mmol), i-PrOH (1.2 mmol), Cu(OAc)2 (2.0 mol%), rac-L4 (2.2 mol%), THF (1 M), 40 °C, 18 h. Isolated yields are reported (average of two runs). See Supplementary Section 3 for details.

(alcohol), direct interception by electrophilic amine 2 (likely via oxidative addition/reductive elimination) would produce the (E)-enamine product 4 and copper benzoate complex 13 that, upon transmetallation with hydrosilane, would regenerate the active CuH species 10. On the other hand, in the presence of alcohol, direct protonation of the vinylcopper intermediate 11 could afford the cis disubstituted alkene 14. A hydroamination similar to that we have previously reported via alkylcopper species a H N Me N S OH 6f (±)-Duloxetine (antidepressant) Me Ref. 44 Bn S O O Me

15 and 16 would deliver the alkylamine 5. As previously detailed, these amination protocols both provide products in excellent regioselectivity; aryl-substituted internal alkynes (and alkenes) proceed through a Markovnikov selective hydrocupration event and terminal aliphatic alkynes (and alkenes) undergo antiMarkovinikov hydrometallation. This regioselectivity is in accord with results previously reported for copper-catalysed alkyne semireduction39,40 and hydroamination23. We rationalize the Markovnikov regioselectivity observed for aryl alkynes as arising from the electronic stabilization of the vinyl- or alkylcopper species by an adjacent aryl group. In contrast, we attribute the anti-Markovnikov selectivity observed in the case of terminal aliphatic alkynes to steric effects (that is, formation of the less substituted vinyl- or alkylcopper species). We conducted mechanistic experiments to gain a better understanding of the factors that result in the high selectivities observed in our system. Subjecting enamine 4a to the conditions developed for either alkyne hydroamination or reductive hydroamination resulted in no observed reaction (Fig. 3b). The inertness of the enamine under these conditions accounts for the exclusive formation of monoamination product in the case of alkyne hydroamination. In addition, these experiments suggest that alkyne hydroamination followed by enamine reduction does not occur in the case of reductive hydroamination. Furthermore, we subjected cis-stilbene (18) to the hydroamination conditions in the presence of 1.5 equiv. ethanol (Fig. 3c). Although a small amount of 1,2-diphenylethane (19, 3% yield) was formed, presumably as a result of protonation of the alkylcopper intermediate48, hydroamination adduct 5a was generated as the predominant product (97% yield). This result suggests that amination of the alkylcopper species 15 occurs selectively in the presence of a proton source. Combined, the results of these experiments are in agreement with our original hypothesis that vinylcopper species 11 and alkylcopper species 15 undergo selective protonation and amination, respectively, thereby allowing the desired cascade reaction to proceed as designed.

b
H N Me Ref. 45 Bn N OH 6g Atomoxetine (ADHD drug) Fluoxetine (antidepressant) CF3 Ref. 46 Me O H N Me

c
OMe

Cu(OAc)2 (4.0 mol%) rac-L4 (4.4 mol%) (iPr)2N–OBz (1.0 equiv.) 3 (4.0 equiv.) Ph 7, 60% yield iPrOH d
OMe i-Pr N Ph 8, 74% yield i-Pr Et Me N O 9, 93% yield O

Cu(OAc)2 (4.0 mol%) (R)-L4 (4.4 mol%) Me2N–OBz (2.0 equiv.) 3 (4.0 equiv.) iPrOH Me

Me

(1.2 equiv.)

(1.1 equiv.)

THF, 40 °C, 18 h

THF, rt, 18 h Me O O Rivastigmine (antidementia drug) 69% yield, >99% e.e. Me

Ref. 47 FeCl3 (10 mol%) 4-methylanisole (2.0 equiv.) CH3CN, 60 °C, 16 h Me HO Ph Ph (±)-Tolterodine (urinary incontinence drug) OH iPr K2CO3 (1.4 equiv.) Carbamic chloride (1.1 equiv.) EtOAc, 80 °C, 8 h Et

Me N

N

Me

N

iPr

HO

Figure 2 | Concise routes to drugs through cascade reductive hydroamination of alkynes. a, Hydroamination product 6f as a key synthetic precursor for duloxetine synthesis. b, Hydroamination product 6g as a known synthetic precursor to atomoxetine and fluoxetine. c, Rapid synthesis of 8, a known precursor to tolterodine. d, Two-step synthesis of rivastigmine.
42
NATURE CHEMISTRY | VOL 7 | JANUARY 2015 | www.nature.com/naturechemistry

© 2014 Macmillan Publishers Limited. All rights reserved.

NATURE CHEMISTRY a BzO LCu N R4 R1 R2 12 R3 R1 R2 (E)-enamine 4 LCu 13 OBz R4 R3

DOI: 10.1038/NCHEM.2131

ARTICLES
LCu R1 R2 11 ROH Protonation 14 15 R3 R4 R1 R2 LCu H LCu R1 R2

R3

N

R4

OBz 2 Amination

N

N

Direct hydroamination

R1

R2

Reductive hydroamination

OBz 2 BzO R3 R4 R1

LCu HSiR3 R3SiOBz 10

H R3SiOBz HSiR3

LCu

OBz 13 R3 R4

LCu N

R2 16

Aryl alkynes LCu Ar H R

Aliphatic alkynes H R CuL H R1

N

R2 Amine 5

Markovnikov selectivity

Anti-Markovnikov selectivity

b
Bn N Bn

c
Cu(OAc)2 (2.0 mo%) (R)-L4 (2.2 mol%) 3 (4.0 equiv.) Bn N Bn Bn N Bn Bn N Ph 17 0% 5a 0% Bn + Bn N Bn Cu(OAc)2 (2.0 mol%) (R)-L4 (2.2 mol%) Ph Ph 18 Ph 3 (4.0 equiv.) 2a (1.2 equiv.) EtOH (1.5 equiv.) THF, 40 °C, 18 h 5a 97% Ph Ph 19 3% Bn N Bn

Ph Ph 4a With or without alcohol additive 2a (1.2 equiv.) THF, 45 °C, 18 h

Ph Ph 4a 99% recovered

+

Ph

Ph

+

Ph Ph

Figure 3 | Proposed catalytic cycles and mechanistic experiments. a, Both direct hydroamination and reductive hydroamination are proposed to proceed through vinylcopper intermediate 11. In the absence of alcohol additive, electrophilic amination occurs directly to give the enamine product (left cycle). In the presence of alcohol additive, protonation of 11 affords alkene 14, which undergoes further hydrocupration and electrophilic amination to furnish the alkylamine product (right cycle). Internal aryl alkynes and terminal aliphatic alkynes undergo hydrocupration with Markovnikov and anti-Markovnikov regioselectivity, respectively. b, Enamine 4a is inert to direct hydroamination conditions (no alcohol additive), thus accounting for the lack of over-reduction or diamination products observed under these conditions. Enamine 4a is also inert to reductive hydroamination conditions (alcohol additive present). Enamine 4a is thus unlikely to be an intermediate in the reductive hydroamination reaction. c, In the presence of ethanol as a proton source and electrophilic amination reagent 2a, cis-stilbene selectively undergoes hydroamination rather than reduction. Hence, the presence of alcohol does not interfere with the hydroamination of proposed alkene intermediate 14, allowing the reductive hydroamination cascade to proceed efficiently to afford the alkylamine product.

Conclusion
We have developed catalytic conditions that allow for the controlled construction of enamines or alkylamines from alkynes and electrophilic amine sources. The products from these complementary systems were obtained with uniformly high levels of regio- and stereocontrol. Both catalytic processes operate through the formation of a vinylcopper intermediate, the product being determined by the presence or absence of an alcohol additive. The development of a protocol for the direct conversion of alkynes to alkylamines is especially notable, given the ease of access to requisite substrates and the demonstrable applicability of this method to the rapid synthesis of a number of pharmaceutical agents. Beyond the broad utility of this new protocol, we anticipate that this cascade strategy will motivate the design of other cascade processes for the more efficient synthesis of valuable targets.

ester 2a (381 mg, 1.2 mmol, 1.2 equiv.). The reaction tube was sealed with a screwcap septum, and then evacuated and backfilled with argon (this process was repeated a total of three times). Anhydrous THF (0.5 ml) and EtOH (88 µl, 1.5 mmol, 1.5 equiv.) were added, followed by dropwise addition of the catalyst solution from the first vial to the stirred reaction mixture at RT. The reaction mixture was then heated at 40 °C for 18 h. After cooling to RT, the reaction was quenched by addition of EtOAc and a saturated aqueous solution of Na2CO3. The phases were separated, the organic phase was concentrated, and product 5a was purified by flash column chromatography. The e.e. of each product was determined by HPLC analysis using chiral stationary phases. All new compounds were fully characterized (see Supplementary Information).

Received 18 September 2014; accepted 6 November 2014; published online 15 December 2014; corrected after print 18 December 2014

References
1. Trost, B. M. Selectivity: the key to synthetic efficiency. Science 219, 245–250 (1983). 2. Dewick, P. M. Medicinal Natural Products: A Biosynthetic Approach 3rd edn (Wiley, 2008). 3. Nugent, T. C. Chiral Amine Synthesis: Methods, Developments and Applications (Wiley, 2010). 4. Surry, D. S. & Buchwald, S. L. Biaryl phosphane ligands in palladium-catalyzed amination. Angew. Chem. Int. Ed. 47, 6338–6361 (2008). 5. Surry, D. S. & Buchwald, S. L. Dialkylbiaryl phosphines in Pd-catalyzed amination: a user’s guide. Chem. Sci. 2, 27−50 (2011). 6. Robak, M. T., Herbage, M. A. & Ellman, J. A. Synthesis and applications of tert-butanesulfinamide. Chem. Rev. 110, 3600–3740 (2010). 7. Roizen, J. L., Harvey, M. E. & Du Bois, J. Metal-catalyzed nitrogen-atom transfer methods for the oxidation of aliphatic C–H bonds. Acc. Chem. Res. 45, 911–922 (2012).
43

Methods
A typical procedure for the copper-catalysed reductive hydroamination of alkynes 1 is as follows (all reactions were set up on the benchtop using a standard Schlenk technique). An oven-dried screw-top reaction tube equipped with a magnetic stir bar was charged with Cu(OAc)2 (3.6 mg, 0.02 mmol, 2 mol%) and (R)-L4 (26 mg, 0.022 mmol, 2.2 mol%). The reaction tube was sealed with a screw-cap septum, then evacuated and backfilled with argon (this process was repeated a total of three times). Anhydrous THF (0.5 ml) and hydrosilane 3 (0.64 ml, 4.0 mmol, 4.0 equiv.) were added sequentially via syringe. The resulting mixture was stirred at room temperature (RT) for 15 min and the colour of the mixture changed from blue to orange. A second oven-dried screw-top reaction tube equipped with a stir bar was charged with alkyne substrate 1a (178 mg, 1.0 mmol, 1.0 equiv.) and hydroxylamine
NATURE CHEMISTRY | VOL 7 | JANUARY 2015 | www.nature.com/naturechemistry

© 2014 Macmillan Publishers Limited. All rights reserved.

ARTICLES
8. Simon, R. C. et al. Regio- and stereoselective monoamination of diketones without protecting groups. Angew. Chem. Int. Ed. 51, 6713–6716 (2012). 9. Höhne, M., Kühl, S., Robins, K. & Bornscheuer, U. T. Efficient asymmetric synthesis of chiral amines by combining transaminase and pyruvate decarboxylase. ChemBioChem 9, 363–365 (2008). 10. Hultzsch, K. C. Catalytic asymmetric hydroamination of non-activated olefins. Org. Biomol. Chem. 3, 1819−1824 (2005). 11. Hannedouche, J. & Schulz, E. Asymmetric hydroamination: a survey of the most recent developments. Chem. Eur. J. 19, 4972–4985 (2013). 12. Müller, T. E., Hultzsch, K. C., Yus, M., Foubelo, F. & Tada, M. Hydroamination: direct addition of amines to alkenes and alkynes. Chem. Rev. 108, 3795–3892 (2008). 13. Chinchilla, R. & Nájera, C. Recent advances in Sonogashira reactions. Chem. Soc. Rev. 40, 5084−5121 (2011). 14. Trost, B. M. & Weiss, A. H. The enantioselective addition of alkyne nucleophiles to carbonyl groups. Adv. Synth. Catal. 351, 963–983 (2009). 15. Habrant, D., Rauhala, V. & Koskinen, A. M. P. Conversion of carbonyl compounds to alkynes: general overview and recent developments. Chem. Soc. Rev. 39, 2007–2017 (2010). 16. Severin, R. & Doye, S. The catalytic hydroamination of alkynes. Chem. Soc. Rev. 36, 1407–1420 (2007). 17. Alonso, F., Beletskaya, I. P. & Yus, M. Transition-metal-catalyzed addition of heteroatom–hydrogen bonds to alkynes. Chem. Rev. 104, 3079–3160 (2004). 18. Zeng, X. Recent advances in catalytic sequential reactions involving hydroelement addition to carbon–carbon multiple bonds. Chem. Rev. 113, 6864–6900 (2013). 19. Li, L. & Herzon, S. B. Regioselective reductive hydration of alkynes to form branched or linear alcohols. J. Am. Chem. Soc. 134, 17376−17379 (2012). 20. Li, L. & Herzon, S. B. Temporal separation of catalytic activities allows antiMarkovnikov reductive functionalization of terminal alkynes. Nature Chem. 6, 22–27 (2014). 21. Zeng, M., Li, L. & Herzon, S. B. A highly active and air-stable ruthenium complex for the ambient temperature anti-Markovnikov reductive hydration of terminal alkynes. J. Am. Chem. Soc. 136, 7058–7067 (2014). 22. Uehling, M. R., Rucker, R. P. & Lalic, G. Catalytic anti-Markovnikov hydrobromination of alkynes. J. Am. Chem. Soc. 136, 8799–8803 (2014). 23. Zhu, S., Niljianskul, N. & Buchwald, S. L. Enantio- and regioselective CuH-catalyzed hydroamination of alkenes. J. Am. Chem. Soc. 135, 15746−15749 (2013). 24. Miki, Y., Hirano, K., Satoh, T. & Miura, M. Copper-catalyzed intermolecular regioselective hydroamination of styrenes with polymethylhydrosiloxane and hydroxylamines. Angew. Chem. Int. Ed. 52, 10830–10834 (2013). 25. Deutsch, C., Krause, N. & Lipshutz, B. H. CuH-catalyzed reactions. Chem. Rev. 108, 2916–2927 (2008). 26. Mahoney, W. S. & Stryker, J. M. Hydride-mediated homogeneous catalysis. Catalytic reduction of α,β-unsaturated ketones using [(Ph3P)CuH]6 and H2. J. Am. Chem. Soc. 111, 8818–8823 (1989). 27. Lipshutz, B. H., Keith, J., Papa, P. & Vivian, R. A convenient, efficient method for conjugate reductions using catalytic quantities of Cu(I). Tetrahedron Lett. 39, 4627–4630 (1998). 28. Appella, D. H., Moritani, Y., Shintani, R., Ferreira, E. M. & Buchwald, S. L. Asymmetric conjugate reduction of α,β-unsaturated esters using a chiral phosphine–copper catalyst. J. Am. Chem. Soc. 121, 9473–9474 (1999). 29. Moritani, Y., Appella, D. H., Jurkauskas, V. & Buchwald, S. L. Synthesis of β-alkyl cyclopentanones in high enantiomeric excess via copper-catalyzed asymmetric conjugate reduction. J. Am. Chem. Soc. 122, 6797–6798 (2000). 30. Yun, J. & Buchwald, S. L. One-pot synthesis of enantiomerically enriched 2,3disubstituted cyclopentanones via copper-catalyzed 1,4-reduction and alkylation. Org. Lett. 3, 1129–1131 (2001). 31. Jurkauskas, V. & Buchwald, S. L. Dynamic kinetic resolution via asymmetric conjugate reduction: enantio- and diastereoselective synthesis of 2,4-dialkyl cyclopentanones. J. Am. Chem. Soc. 124, 2892–2893 (2002). 32. Rainka, M. P., Aye, Y. & Buchwald, S. L. Copper-catalyzed asymmetric conjugate reduction as a route to novel β-azaheterocyclic acid derivatives. Proc. Natl Acad. Sci. USA 101, 5821−5823 (2004).

NATURE CHEMISTRY

DOI: 10.1038/NCHEM.2131

33. Erdik, E. & Ay, M. Electrophilic amination of carbanions. Chem. Rev. 89, 1947–1980 (1989). 34. Barker, T. J. & Jarvo, E. R. Developments in transition-metal-catalyzed reactions using electrophilic nitrogen sources. Synthesis 24, 3954–3964 (2011). 35. Berman, A. M. & Johnson, J. S. Copper-catalyzed electrophilic amination of diorganozinc reagents. J. Am. Chem. Soc. 126, 5680–5681 (2004). 36. Campbell, M. J. & Johnson, J. S. Mechanistic studies of the copper-catalyzed electrophilic amination of diorganozinc reagents and development of a zinc-free protocol. Org. Lett. 9, 1521–1524 (2007). 37. Xu, P-F. & Wang, W. Catalytic Cascade Reactions (Wiley, 2014). 38. Daeuble, J. F., McGettigan, C. & Stryker, J. M. Selective reduction of alkynes to cis-alkenes by hydrometallation using [(Ph3P)CuH]6. Tetrahedron Lett. 31, 2397–2400 (1990). 39. Semba, K., Fujihara, T., Xu, T., Terao, J. & Tsuji, Y. Copper-catalyzed highly selective semihydrogenation of non-polar carbon–carbon multiple bonds using a silane and an alcohol. Adv. Synth. Catal. 354, 1542–1550 (2012). 40. Whittaker, A. M. & Lalic, G. Monophasic catalytic system for the selective semireduction of alkynes. Org. Lett. 15, 1112−1115 (2013). 41. Field, L. D., Messerle, B. A. & Wren, S. L. One-pot tandem hydroamination/hydrosilation catalyzed by cationic iridium(I) complexes. Organometallics 22, 4393–4395 (2003). 42. Heutling, A., Pohlki, F., Bytschkov, I. & Doye, S. Hydroamination/hydrosilylation sequence catalyzed by titanium complexes. Angew. Chem. Int. Ed. 44, 2951–2954 (2005). 43. Hesp, K. D. & Stradiotto, M. Stereo- and regioselective gold-catalyzed hydroamination of internal alkynes with dialkylamines. J. Am. Chem. Soc. 132, 18026–18029 (2010). 44. Jozsef, B. et al. Process for preparation of duloxetine and intermediates. International patent WO 2008078124 (2008). 45. Wu, F., Chen, G. & Yang, X. Process for preparation of propylamine derivatives and application in manufacturing tomoxetine. Chinese patent CN1948277 (2007). 46. Bhandari, K., Srivastava, S., Shanker, G. & Nath, C. Substituted propanolamines and alkylamines derived from fluoxetine as potent appetite suppressants. Bioorg. Med. Chem. 13, 1739–1747 (2005). 47. Paras, N. A., Simmons, B. & MacMillan, D. W. C. A process for the rapid removal of dialkylamino-substituents from aromatic rings. Application to the expedient synthesis of (R)-tolterodine. Tetrahedron 65, 3232−3238 (2009). 48. Rupnicki, L., Saxena, A. & Lam, H. W. Aromatic heterocycles as activating groups for asymmetric conjugate addition reactions: enantioselective coppercatalyzed reduction of 2-alkenylheteroarenes. J. Am. Chem. Soc. 131, 10386–10387 (2009).

Acknowledgements
The authors acknowledge the National Institutes of Health for financial support (GM58160). The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. The authors thank S. Zhu (MIT), P.J. Milner (MIT) and M. Pirnot (MIT) for discussions. The authors thank Y. Wang (MIT) and N.T. Jui (Emory University) for help with the preparation of this manuscript.

Author contributions
S-L.S. and S.L.B. designed the project, analysed the data and wrote the manuscript. S-L.S. performed the experiments.

Additional information
Supplementary information and chemical compound information are available in the online version of the paper. Reprints and permissions information is available online at www.nature.com/reprints. Correspondence and requests for materials should be addressed to S.L.B.

Competing financial interests

The authors declare no competing financial interests.

44

NATURE CHEMISTRY | VOL 7 | JANUARY 2015 | www.nature.com/naturechemistry

© 2014 Macmillan Publishers Limited. All rights reserved.

ERRATUM

Copper-catalysed selective hydroamination reactions of alkynes
Shi-Liang Shi and Stephen L. Buchwald
Nature Chemistry 7, 38–44 (2015); published online 15 December 2014; corrected after print 18 December 2014. In the version of this Article originally published, the red arrow at the top right of Fig. 1c should have been solid not dashed. This has now been corrected in the online versions of the Article.

© 2014 Macmillan Publishers Limited. All rights reserved.

Similar Documents