Enantioselective Copper Catalyzed Alkyne-Azide Cycloaddition by

and chemo-orthogonal.3–5 ... chemistry.5 Triazoles, formed by CuAAC, are now common ..... Catalyzed 1,3-Dipolar Cycloadditions of Terminal Alkynes t...
0 downloads 0 Views 286KB Size
Subscriber access provided by Drexel University Libraries

Communication

Enantioselective Copper Catalyzed Alkyne-Azide Cycloaddition by Dynamic Kinetic Resolution En-Chih Liu, and Joseph J. Topczewski J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.9b01091 • Publication Date (Web): 19 Mar 2019 Downloaded from http://pubs.acs.org on March 19, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Enantioselective Copper Catalyzed Alkyne-Azide Cycloaddition by Dynamic Kinetic Resolution En-Chih Liu, Joseph J. Topczewski* Department of Chemistry, University of Minnesota Twin Cities, Minneapolis MN 55455

Supporting Information Placeholder ABSTRACT: The copper(I) catalyzed alkyne-azide cycloaddition (CuAAC), a click reaction, is one of the most powerful catalytic reactions developed during the last two decades. Conducting CuAAC enantioselectively would add a third dimension to this reaction and would enable the direct synthesis of α-chiral triazoles. Doing so is demanding because the two precursors have linear geometries, and the triazole product is a flat heterocycle. Designing a chiral catalyst is further complicated by the complex mechanism of CuAAC. We report an enantioselective CuAAC (E-CuAAC), enabled by dynamic kinetic resolution (DKR). The E-CuAAC is high yielding and affords up to 99:1 er. The E-CuAAC can directly generate α-chiral triazoles in a complex molecular environment.

able to sense remote stereochemical information. Furthermore, even in the ideal sense, kinetic resolution proceeds with a maximum theoretical yield of 50%.16 Others have attempted to use bis-alkynes and/or bis-azides for E-CuAAC by desymmetrization (Figure 1c).15,17–24 Reactions based on this approach are likewise limited in scope and occur with modest chemoselectivity due to the competitive formation of bis-triazoles. Figure 1. CuAAC Reaction and Approaches to E-CuAAC a CuAAC Reaction +

R'

R

[Cu]

N

N

R N

N N

N

R'

b Kinetic Resolution

The copper(I) catalyzed alkyne-azide cycloaddition (CuAAC) has transformed many aspects of modern chemical synthesis since it was first reported contemporaneously by Meldal, Sharpless, and coworkers.1,2 The CuAAC reaction is robust, mild, high yielding, and chemo-orthogonal.3–5 Applications for CuAAC have permeated and transformed numerous fields including chemical biology, material science, polymer chemistry, and medicinal chemistry.5 Triazoles, formed by CuAAC, are now common peptidomimetics and pharmaceutical building blocks.6 With the tremendous utility of CuAAC, a versatile catalyst that could impart enantioselectivity to the process would likely find numerous applications, especially as examples of α-chiral triazoles are emerging in active biological agents.7–11 Facilitating an E-CuAAC reaction presents several fundamental challenges. First, CuAAC uses an alkyne and an azide and forms a triazole (Figure 1a). Alkynes and azides have a linear geometry and the resulting triazole is a sp2 hybridized heterocycle. No new stereogenic centers are formed in most Cu-AAC reactions. Therefore, E-CuAAC requires the transmission of stereochemical information beyond the forming triazole. Second, an E-CuAAC reaction must outcompete the facile background CuAAC reaction.12 This is non-trivial because the CuAAC reaction is an extremely efficient process that proceeds by a complex and dynamic reaction mechanism.13,14 Herein, we report that E-CuAAC can be enabled by dynamic kinetic resolution (>95% yield and up to 99:1 er). Fokin and Finn originally reported attempts at an E-CuAAC through a kinetic resolution (Figure 1b).15 The results were significantly limited in scope and proceeded with only a modest selectivity (selectivity factor s up to 6). These early results implied a twofold problem with E-CuAAC. First, the background CuAAC, in the absence of a chiral ligand, is fast and must be outcompeted or suppressed. Second, the catalyst’s ligand environment must be

N3

N3 Ph

CH3

+

Ph

N N

Ph [Cu], Ligand

Ph

Ph

CH3

N3

N +

CH3

c Desymmetrization BnN3 [Cu], Ligand

R

Ph

CH3

50% maximum theoretical yield 32% conversion, s = 6

Limited scope

R

R'

R'

N N N Bn

R

+

R' N Bn N Bn N N N N competitive bis-triazole formation

d Dynamic Kinetic Resolution R'

R' [3,3]

N3 R

R

N3 R

R

R'

R" [Cu], Ligand R

N N N

R"

*R

95% yield, 99:1 er

Our group has an interest in using dynamic kinetic resolution (DKR)25–29 to enable enantioselective synthetic methods. A DKR couples a pathway for racemization to an enantioselective functionalization. Methods based on DKR use racemic starting material and can result in both high yield and high enantioselectivity. Our lab envisioned using allylic azides in DKR because allylic azides spontaneously rearrange.30 The rearrangement complicates using these intermediates but, a few inspirational reports described successfully trapping allylic azides.31–35 We hypothesized that allylic azides could be used to enable an E-CuAAC reaction (Figure 1d). Herein, we report a DKR enabled E-CuAAC that proceeds in yields exceeding 95%, with er up to 99:1, which is compatible with a complex molecular environment.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

This study began with allylic azide 1a and tert-butyl propiolate (2a, Table 1). We observed minimal enantioselectivity with copper iodide (entry 1). Changing to a cationic copper(I) precatalyst had a notable impact on both the rate and enantioselectivity of the reaction (entry 2). A collection of ligands were screened that included bidentate and tridentate phosphorous and nitrogen ligands (entries 2 – 9 and Supporting Information). Based on these initial results, aryl-PYBOX ligands appeared to be particularly effective (entry 4 and 5). We hypothesized that increasing the ligand loading would slow the background click reaction by saturating the copper center. An increase in ligand loading enhanced the observed er to 88:12 (entry 10). Increasing the temperature had a notable positive effect (entry 11), which resulted in a quantitative yield (>98%) and high enantioselectivity (>99:1 er). The increased temperature likely increases the relative rate of racemization via a sigmatropic pathway.36 Other copper precatalysts were not as effective (entries 12 and 13) and the conditions outlined in entry 11 were selected as being optimal.

CO2tBu PMP N3

2a (1.2 equiv) 2.5 mol% [Cu] X mol% Ligand

PMP N3

[3,3]

Ligand

Temp (˚C)

Yield (%)

er

1

CuI

L1 (2.5%)

rt

80

57:43

2

(CuOTf)2PhMe

L1 (2.5%)

rt

>98

60:40

3

(CuOTf)2PhMe

L2 (2.5%)

rt

95

53:47

4

(CuOTf)2PhMe

L3 (2.5%)

rt

93

86:14

5

(CuOTf)2PhMe

L4 (2.5%)

rt

80

76:24

6

(CuOTf)2PhMe

L5 (2.5%)

rt

58

51:49

7

(CuOTf)2PhMe

L6 (2.5%)

rt

87

52:48

8

(CuOTf)2PhMe

L7 (2.5%)

rt

65

52:48

9

(CuOTf)2PhMe

L8 (2.5%)

rt

>98

52:48

10

(CuOTf)2PhMe

L4 (5.0%)

rt

83

88:12

11

(CuOTf)2PhMe

L4 (5.0%)

40

>98

99:1

12

Cu(MeCN)4PF6

L4 (5.0%)

40

73

90:10

13

Cu(MeCN)4BF4

L4 (5.0%)

40

82

91:9

Reactions conducted with allylic azide 1a (0.1 mmol), alkyne 2a (0.12 mmol), in dimethoxyethane (0.2 M), with 2.5 mol% [Cu] and either 2.5 mol% or 5 mol% ligand. Yields based on 1H NMR analysis using 1,3,5-trimethoxybenzene as an internal standard. Chiral HPLC was used to determine er. All yield and er values reflect the average of duplicate trials. See Supporting Information for full details. O

O

R

O N

N R

Bn

Bn

N

O O

N N

Bn

L5

i

L1 R = Pr L2 R = tBu L3 R = Ph L4 R = 4-Cl-Ph

+

3a 95%, >99:1 er

Bu

PMP N N N

PMP N N N

PMP

PMP N N N

Ph

PMP N N N

CH3

3c 91%, 93:7 er

3e 85%, 96:4 er

Cl

R

*

3

3b 89%, 93:7 er

t

PMP N N N

DME, 20-48 h 40 °C

PMP N N N

CO2tBu

PMP N N N

R

2 (1.2 equiv)

3g 91%, 92:8 er

[Cu] source

N

N3

(±)-1a (1 equiv)

PMP N N N

1.25 mol% (CuOTf)2PhMe 5 mol% L4

PMP [3,3]

F

3f 90%, 93:7 er

Br

PMP N N N

CF3

3h 86%, 95:5 er

3i 92%, 90:10 er

3a

Entry

N

PMP N3

CO2tBu

*

(±)-1a (1.0 equiv)

O

Table 2. Scope of DKR E-CuAAC with Respect to Alkyne 2

PMP N N N

PMP N N N

DME, 24 h, temp

scaffold may promote dimerization, which is consistent with crystallographic data on Cu(I)-PYBOX complexes.38,39

3d 96%, 92: 8 er

Table 1. Optimization of E-CuAAC by DKRa

Page 2 of 6

L6 Bn O

F3C O

N N L7

t

Bu

O

PPh2

O

PPh2

O

L8

We conducted a non-linear experiment by determining the effect of varying the ligand’s enantiopurity on the enantiomeric excess of the product.37 It has been reported that E-CuAAC by desymmetrization can proceed with a posative19,24 or negative17 non-linear effect. A negative non-linear effect was observed for this E-CuAAC reaction (see Supporting Information). The PYBOX

PMP N N N

PMP N N N Cl

3j 80%, 95:5 er PMP N N N

OTBS

3m 95%, 94:6 er

CH3 CH3 CH3

3k 71%, 92:8 er PMP N N N

O CH3

3n 96%, 92:8 er

PMP N N N

NBoc

3l 96%, 93:7 er PMP N N N

3o 80%, 97:3 er

Yields are reported for isolated and purified products. Enantiomeric ratio was determined by chiral HPLC. Yield and er values reflect the average of duplicate trials. See Supporting Information for details.

The scope of our E-CuAAC reaction was investigated with respect to the alkyne (Table 2). Using the azide as the limiting reagent, the model substrate 3a was isolated in 95% yield and >99:1 er. The scope of the alkyne was quite broad. Electron rich and electron deficient aryl alkynes could participate in the E-CuAAC (3b – 3i). A crystal of product 3h was suitable for diffraction analysis, which unambiguously assigned the absolute configuration of the product generated from the (S,S)-PYBOX/Cu catalyst as having the (R)-configuration (see Supporting Information). The configuration of the other products were assigned based on analogy to product 3h. An ortho-substituted arene provided an acceptable yield in high er (3j). An alkyne containing an aliphatic (3j), heterocycle (3k), protected alcohol (3l), ketone (3m), and cyclopropyl (3o) group could all be used in E-CuAAC. The azide component was varied (Table 3). The cyclohexyl-ring could be contracted (3p), expanded (3q), or modified (3r – 3x). In all of these cases, both the yield and enantioselectivity remained acceptably high. The 2-aryl group is not required (3y and 3z) and an acyclic substrate was tolerated (3aa). To further demonstrate the features of this E-CuAAC reaction, we conducted the reaction with (R)-1-phenyl-2-propyn-1-ol (4, Scheme 1). Both enantiomers of the ligand were used to test for matched/mismatched behavior related to double diastereoselectivity. Changing the ligand’s stereochemistry reversed the diastereoselectivity (products 5a and 5b), indicating a robust catalyst.

ACS Paragon Plus Environment

Page 3 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Experimental procedures and data. Table 4. DKR E-CuAAC in a Complex Molecular Setting

Table 3. Scope of DKR E-CuAAC with Respect to Azide 1 R N3

N3

[3,3] R'

1.25 mol% (CuOTf)2PhMe 5 mol% L4

R R'

R"

+

DME, 24 h, 40 °C

R'

R'

(±)-1 (1 equiv)

N N N

R R'

2 (1.2 equiv)

*R'

PMP

R"

3

N3

+

PMP N N

CO2tBu

N

CO2tBu

Ph

N N N

(±)-1a (1 equiv)

3q 99%, 90:10 er

CH3

CH3

O

CH3

O

O O

N N

PMP H

N N N

PMP N N N

CO2tBu

CO2tBu

Ph

N N N

H3C CH3 3t 85%, 96:4 er

3s 96%, 94:6 er

PMP

Ph

*

H

*

H

O

N

6b 99%, 90:10 dr

OAc

N N N

AcO AcO

6c 99%, 86:14 dr

3u 85%, 96:4 er

OAc

H 3C H

6a 92%, 97:3 dr

CF3

OH

AcO

(H3C)2HC

CH3

3r 96%, 95:5 er

complex fragment

6

O H 3C

3p 79%, 94:6 er

*

2 (1.2 equiv)

N N N

*

CO2tBu

PMP N N N

DME, 24 h, 40 °C

PMP

PMP N N N

1.25 mol% (CuOTf)2PhMe 5 mol% L4

complex fragment

CF3

N N N

O OAc

O

*

PMP 6d 99%, 93:7 dr

N N N

Ph Ph

3v 82%, 93:7 er H 3C OH H 3C N N N

N N N

Ph CO2Et

3w 85%, 96:4 er

N N N

O

* N N N

CO2tBu

O

CO2tBu

N N N

CO2tBu

Ph

N N N

O OCH3

CO2tBu

PMP

OH +

*1

Ph

4 (1.2 equiv)

DME, 24 h, 40 °C

PMP N N N

*

OH

* Ph

5a 95%, 95:5 dr 1.25 mol% (CuOTf)2PhMe 5 mol% ent-L4 DME, 24 h, 40 °C

PMP N N N

*

OH

* Ph

5b 92%, 6:94 dr

Several additional examples demonstrate that the E-CuAAC is viable in a complex molecular setting (Table 4). Derivatives of vitamin E (6a), gibberellic acid (6b), esterone (6c), glucose (6d), mycophenolate mofetil (6e), and moexipril (6f) could all be successfully “clicked” by E-CuAAC in near perfect yield and excellent selectivity. This illustrates that E-CuAAC is capable of generating α-chiral triazoles in the presence of densely functionalized molecules with robust stereochemical fidelity. We report an effective system for the enantioselective copper(I) catalyzed alkyne-azide cycloaddition (E-CuAAC) “click” reaction that is enabled by the dynamic kinetic resolution of allylic azides. A negative non-linear effect was observed in this system. The reaction proceeds in high yield and high selectivity. The scope of this process is broad and the reaction can proceed in a complex molecular environment.

ASSOCIATED CONTENT

O

*

O

OBn

N

OCH3 OCH3 6f 99%, 97:3 dr

Yields are reported for isolated and purified products. Diastereomeric ratio or enantiomeric ratio was determined by HPLC or 1H NMR. Yield, er, and dr values reflect the average of duplicate trials. Substrate 6f was prepared using (R,R)-4-Cl-PhPYBOX ligand (ent-L4). See Supporting Information for full details.

AUTHOR INFORMATION

Scheme 1. Test for Matched/Mismatched Behavior

N3

O

N N

6e 98%, 93:7 er

3aa 92%, 86:14 er 93:7 E:Z

Yields are reported for isolated and purified products. Enantiomeric ratio was determined by chiral HPLC. Yield and er values reflect the average of duplicate trials. aReaction used alkyne as limiting reagent. See Supporting Information for details.

1.25 mol% (CuOTf)2PhMe 5 mol% L4

PMP

N

CH3

CH3 CH3 3za 89%, 87:13 er

N

CH3

O

I

3y 96%, 90:10 er

(±)-1a (1.0 equiv)

O

O

3xa 90%, 85:15 er

PMP

Corresponding Author *[email protected]

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT We acknowledge the University of Minnesota and The American Chemical Society’s Petroleum Research Fund (PRF #56505DNI1) for financial support. This research was supported by the National Institute of General Medical Sciences of the National Institutes of Health under Award Number R35GM124718.

REFERENCES (1)

Tornøe, C. W.; Christensen, C.; Meldal, M. Peptidotriazoles on Solid Phase: [1,2,3]-Triazoles by Regiospecific Copper(I)Catalyzed 1,3-Dipolar Cycloadditions of Terminal Alkynes to Azides. J. Org. Chem. 2002, 67, 3057–3064. https://doi.org/10.1021/jo011148j.

(2)

Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. A Stepwise Huisgen Cycloaddition Process: Copper(I)-Catalyzed Regioselective “Ligation” of Azides and Terminal Alkynes. Angew. Chemie Int. Ed. 2002, 41, 2596–2599. https://doi.org/10.1002/1521-3773(20020715)41:143.0.CO;2-4.

(3)

Hein, J. E.; Fokin, V. V. Copper-Catalyzed Azide-Alkyne Cycloaddition (CuAAC) and beyond: New Reactivity of

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(21)

Osako, T.; Uozumi, Y. Enantioposition-Selective CopperCatalyzed Azide–Alkyne Cycloaddition for Construction of Chiral Biaryl Derivatives. Org. Lett. 2014, 16, 5866. https://doi.org/10.1021/ol502778j.

(22)

Brittain, W. D. G.; Buckley, B. R.; Fossey, J. S. Kinetic Resolution of Alkyne-Substituted Quaternary Oxindoles via Copper Catalysed Azide-Alkyne Cycloadditions. Chem. Commun. 2015, 51, 17217–17220. https://doi.org/10.1039/C5CC04886A.

(23)

Song, T.; Li, L.; Zhou, W.; Zheng, Z.-J.; Deng, Y.; Xu, Z.; Xu, L.-W. Enantioselective Copper-Catalyzed Azide–Alkyne Click Cycloaddition to Desymmetrization of Maleimide-Based Bis(Alkynes). Chem. – A Eur. J. 2015, 21, 554–558. https://doi.org/10.1002/chem.201405420.

(24)

Chen, M.-Y.; Xu, Z.; Chen, L.; Song, T.; Zheng, Z.-J.; Cao, J.; Cui, Y.-M.; Xu, L.-W. Catalytic Asymmetric Huisgen Alkyne– Azide Cycloaddition of Bisalkynes by Copper(I) Nanoparticles. ChemCatChem 2018, 10, 280–286. https://doi.org/10.1002/cctc.201701336.

(25)

Verho, O.; Bäckvall, J. E. Chemoenzymatic Dynamic Kinetic Resolution: A Powerful Tool for the Preparation of Enantiomerically Pure Alcohols and Amines. J. Am. Chem. Soc. 2015, 137, 3996–4009. https://doi.org/10.1021/jacs.5b01031.

(26)

Bhat, V.; Welin, E. R.; Guo, X.; Stoltz, B. M. Advances in Stereoconvergent Catalysis from 2005 to 2015: Transition-MetalMediated Stereoablative Reactions, Dynamic Kinetic Resolutions, and Dynamic Kinetic Asymmetric Transformations. Chem. Rev. 2017, 117, 4528–4561. https://doi.org/10.1021/acs.chemrev.6b00731.

(27)

Pellissier, H. Recent Developments in Dynamic Kinetic Resolution. Tetrahedron 2008, 64, 3769–3802. https://doi.org/http://dx.doi.org/10.1016/j.tet.2007.10.080.

(28)

Huerta, F.; Minidis, A. Racemisation in Asymmetric Synthesis. Dynamic Kinetic Resolution and Related Processes in Enzyme and Metal Catalysis. Chem. Soc. Rev. 2001, 30, 321–331. https://doi.org/10.1039/b105464n.

(29)

Breit, B.; Hilpert, L.; Sieger, S.; Haydl, A. Palladium‐ and Rhodium‐Catalyzed Dynamic Kinetic Resolution of Racemic Internal Allenes Towards Chiral Pyrazoles. Angew. Chemie Int. Ed. 2018, 58, 3378–3381. https://doi.org/10.1002/anie.201812984.

(30)

Gagneux, A.; Winstein, S.; Young, W. G. Rearrangement of Allyl Azides. J. Am. Chem. Soc. 1960, 82, 5956–5957.

(31)

Meng, J. C.; Fokin, V. V.; Finn, M. G. Kinetic Resolution by Copper-Catalyzed Azide-Alkyne Cycloaddition. Tetrahedron Lett. 2005, 46, 4543–4546. https://doi.org/10.1016/j.tetlet.2005.05.019.

Feldman, A. K.; Colasson, B. B.; Sharpless, K. B.; Fokin, V. V. The Allylic Azide Rearrangement: Achieving Selectivity. J. Am. Chem. Soc. 2005, 127, 13444–13445. https://doi.org/10.1021/ja050622q.

(32)

Keith, J. M.; Larrow, J. F.; Jacobsen, E. N. Practical Considerations in Kinetic Resolution Reactions. Adv. Synth. Catal. 2001, 343, 5–26. https://doi.org/10.1002/16154169(20010129)343:13.0.CO;2-I.

Liu, R.; Gutierrez, O.; Tantillo, D. J.; Aubé, J. Stereocontrol in a Combined Allylic Azide Rearrangement and Intramolecular Schmidt Reaction. J. Am. Chem. Soc. 2012, 134, 6528–6531. https://doi.org/10.1021/ja300369c.

(33)

Zhou, F.; Tan, C.; Tang, J.; Zhang, Y. Y.; Gao, W. M.; Wu, H. H.; Yu, Y. H.; Zhou, J. Asymmetric Copper(I)-Catalyzed AzideAlkyne Cycloaddition to Quaternary Oxindoles. J. Am. Chem. Soc. 2013, 135, 10994–10997. https://doi.org/10.1021/ja4066656.

Ott, A. A.; Goshey, C. S.; Topczewski, J. J. Dynamic Kinetic Resolution of Allylic Azides via Asymmetric Dihydroxylation. J. Am. Chem. Soc. 2017, 139, 7737–7740. https://doi.org/10.1021/jacs.7b04203.

(34)

Brittain, W. D. G.; Buckley, B. R.; Fossey, J. S. Asymmetric Copper-Catalyzed Azide-Alkyne Cycloadditions. ACS Catal. 2016, 6, 3629–3636. https://doi.org/10.1021/acscatal.6b00996.

Porter, M. R.; Shaker, R. M.; Calcanas, C.; Topczewski, J. J. Stereoselective Dynamic Cyclization of Allylic Azides: Synthesis of Tetralins, Chromanes, and Tetrahydroquinolines. J. Am. Chem. Soc. 2018, 140, 1211–1214. https://doi.org/10.1021/jacs.7b11299.

(35)

Osako, T.; Uozumi, Y. Mechanistic Insights into CopperCatalyzed Azide–Alkyne Cycloaddition (CuAAC): Observation of Asymmetric Amplification. Synlett 2015, 26, 1475–1479. https://doi.org/10.1055/s-0034-1380534.

Vekariya, R. H.; Liu, R.; Aubé, J. A Concomitant Allylic Azide Rearrangement/Intramolecular Azide-Alkyne Cycloaddition Sequence. Org. Lett. 2014, 16, 1844–1847. https://doi.org/10.1021/ol500011f.

(36)

Ott, A. A.; Packard, M. H.; Ortuño, M. A.; Johnson, A.; Suding, V. P.; Cramer, C. J.; Topczewski, J. J. Evidence for a Sigmatropic and an Ionic Pathway in the Winstein Rearrangement. J. Org. Chem. 2018, 83, 8214. https://doi.org/10.1021/acs.joc.8b00961.

(37)

Girard, C.; Kagan, H. B. Nonlinear Effects in Asymmetric Synthesis and Stereoselective Reactions: Ten Years of

Copper(I) Acetylides. Chem. Soc. Rev. 2010, 39, 1302–1315. https://doi.org/10.1039/B904091A. (4)

Meldal, M.; Tornøe, C. W. Cu-Catalyzed Azide−Alkyne Cycloaddition. Chem. Rev. 2008, 108, 2952–3015. https://doi.org/10.1021/cr0783479.

(5)

Finn, M. G.; Fokin, V. V. Click Chemistry: Function Follows Form. Chem. Soc. Rev. 2010, 39, 1231–1232. https://doi.org/10.1039/C003740K.

(6)

Pedersen, D. S.; Abell, A. 1,2,3-Triazoles in Peptidomimetic Chemistry. Eur. J. Org. Chem. 2011, 2011, 2399–2411. https://doi.org/10.1002/ejoc.201100157.

(7)

Wood, W. J. L.; Patterson, A. W.; Tsuruoka, H.; Jain, R. K.; Ellman, J. A. Substrate Activity Screening: A Fragment-Based Method for the Rapid Identification of Nonpeptidic Protease Inhibitors. J. Am. Chem. Soc. 2005, 127, 15521–15527. https://doi.org/10.1021/ja0547230.

(8)

(9)

(10)

Lee, T.; Cho, M.; Ko, S. Y.; Youn, H. J.; Dong, J. B.; Cho, W. J.; Kang, C. Y.; Kim, S. Synthesis and Evaluation of 1,2,3-Triazole Containing Analogues of the Immunostimulant α-GalCer. J. Med. Chem. 2007, 50, 585–589. https://doi.org/10.1021/jm061243q. Divakaran, A.; Talluri, S. K.; Ayoub, A. M.; Mishra, N.; Cui, H.; Widen, J. C.; Berndt, N.; Zhu, J.-Y.; Carlson, A. S.; Topczewski, J. J.; et al. Molecular Basis for the N-Terminal Bromodomain and Extra Terminal (BET) Family Selectivity of a Dual KinaseBromodomain Inhibitor. J. Med. Chem. 2018, 61, 9316. https://doi.org/10.1021/acs.jmedchem.8b01248. Rezaei, Z.; Khabnadideh, S.; Pakshir, K.; Hossaini, Z.; Amiri, F.; Assadpour, E. Design, Synthesis, and Antifungal Activity of Triazole and Benzotriazole Derivatives. Eur. J. Med. Chem. 2009, 44, 3064–3067. https://doi.org/https://doi.org/10.1016/j.ejmech.2008.07.012.

(11)

Berthold, D.; Breit, B. Chemo-, Regio-, and Enantioselective Rhodium-Catalyzed Allylation of Triazoles with Internal Alkynes and Terminal Allenes. Org. Lett. 2018, 20, 598–601. https://doi.org/10.1021/acs.orglett.7b03708.

(12)

Berrisford, D. J.; Bolm, C.; Sharpless, K. B. Ligand‐Accelerated Catalysis. Angew. Chemie Int. Ed. 1995, 34, 1059–1070. https://doi.org/10.1002/anie.199510591.

(13)

Worrell, B. T.; Malik, J. A.; Fokin, V. V. Direct Evidence of a Dinuclear Copper Intermediate in Cu(I)-Catalyzed Azide-Alkyne Cycloadditions. Science 2013, 340, 457–460.

(14)

Ziegler, M. S.; Lakshmi, K. V; Tilley, T. D. Dicopper Cu(I)Cu(I) and Cu(I)Cu(II) Complexes in Copper-Catalyzed Azide–Alkyne Cycloaddition. J. Am. Chem. Soc. 2017, 139, 5378–5386. https://doi.org/10.1021/jacs.6b13261.

(15)

(16)

(17)

(18)

(19)

(20)

Mu-Yi, C.; Zheng, X.; Li, C.; Tao, S.; Zhan-Jiang, Z.; Jian, C.; Yu-Ming, C.; Li-Wen, X. Catalytic Asymmetric Huisgen Alkyne–Azide Cycloaddition of Bisalkynes by Copper(I) Nanoparticles. ChemCatChem 2017, 10, 280–286. https://doi.org/10.1002/cctc.201701336.

ACS Paragon Plus Environment

Page 4 of 6

Page 5 of 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society Investigation. Angew. Chemie Int. Ed. 1998, 37, 2922–2959. https://doi.org/10.1002/(SICI)15213773(19981116)37:213.0.CO;2-1. (38)

Díez, J.; Gamasa, M. P.; Panera, M. Tetra-, Di-, and Mononuclear Copper(I) Complexes Containing (S,S)-IPr-Pybox and (R,R)-PhPybox Ligands. Inorg. Chem. 2006, 45, 10043–10045.

https://doi.org/10.1021/ic061453t. (39)

Nakajima, K.; Shibata, M.; Nishibayashi, Y. Copper-Catalyzed Enantioselective Propargylic Etherification of Propargylic Esters with Alcohols. J. Am. Chem. Soc. 2015, 137, 2472–2475. https://doi.org/10.1021/jacs.5b00004.

ACS Paragon Plus Environment

Journal of the American Chemical Society 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

R' N3 R

R'

R" [Cu], Ligand R

R

*

N N N

Page 6 of 6

R"

R

>95% yield, 99:1 er racemic azide limiting 35 examples, low Cu loading, complex molecules

ACS Paragon Plus Environment

6