Silver-Catalyzed [3 + 3] Dipolar Cycloaddition of Trifluorodiazoethane

3 hours ago - Herein we disclose a silver-catalyzed [3 + 3] 1,3-dipolar cycloaddition reaction of trifluorodiazoethane with glycine imines. This react...
0 downloads 0 Views 416KB Size
Subscriber access provided by CAL STATE UNIV BAKERSFIELD

Article

Silver-Catalyzed [3 + 3] Dipolar Cycloaddition of Trifluorodiazoethane and Glycine Imines: Access to Highly Functionalized Trifluoromethyl-Substituted Triazines and Pyridines Zhen Chen, Nan Ren, Xiaoxiao Ma, Jing Nie, Fa-Guang Zhang, and Jun-An Ma ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.9b00846 • Publication Date (Web): 12 Apr 2019 Downloaded from http://pubs.acs.org on April 12, 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 8 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

ACS Catalysis

Silver-Catalyzed [3 + 3] Dipolar Cycloaddition of Trifluorodiazoethane and Glycine Imines: Access to Highly Functionalized Trifluoromethyl-Substituted Triazines and Pyridines Zhen Chen,§,† Nan Ren,§,† Xiaoxiao Ma,† Jing Nie,† Fa-Guang Zhang,†,* and Jun-An Ma†,‡,* † Department

of Chemistry, Tianjin Key Laboratory of Molecular Optoelectronic Sciences, and Tianjin Collaborative Innovation Center of Chemical Science & Engineering, Tianjin University, Tianjin 300072, P.R. China. ‡

State Key Laboratory of Elemento-Organic Chemistry, Nankai University, Tianjin 300071, P.R. China.

Supporting Information

ABSTRACT: Herein we disclose a silver-catalyzed [3 + 3] 1,3CF3 CF3 R1 CF3CHN2 R1 dipolar cycloaddition reaction of trifluorodiazoethane with glycine Ag [O] NH N [3 + 3] + imines. This reaction exhibits manifold remarkable features, such N N N NH 21 examples 34 examples 2 1 as easily available substrates, high regio- and diastereoselectivities, R N CO2R 55-99%, >99:1 dr 42-92% 2 CO2R CO2R2 mild reaction conditions, and good yields across a broad spectrum CF3 CF3 of substrates. Moreover, swift transformations of obtained 3 H 1 R4 R1 R3 R Click ligation R tetrahydrotriazines provide efficient access to a diverse set of or R3COR4 with s-TCO N N 12 examples highly functionalized trifluoromethyl-substituted triazines and OH k up to R4 50-99% 2 H pyridines. Particularly noteworthy is that such trifluoromethylated 2 CO2R2 99.24 M-1 s-1 CO2R 1,2,4-triazines demonstrate competent reactivity toward transcyclooctene (TCO) derivatives with the second order rate constants (k2) up to 99.24 M-1 s-1, which represents the highest value involving 1,2,4-triazines to date. KEYWORDS: dipolar cycloaddition, trifluorodiazoethane, silver catalysis, 1,2,4-triazines, trifluoromethylated hetereocycles 1. INTRODUCTION Scheme 1. Different Patterns of 1,3-Dipolar Cycloaddition Trifluoromethyl-substituted N-heterocycles are important core Reactions with CF3CHN2 structures found in numerous compounds of pharmaceutical, 1 agrochemical, and bioactive significance. Therefore, their a) Previous work: [3 + 2] DPC HN N construction has attracted considerable interest from organic R1 CF3 1 F3C community during the past few decades, and various useful R CH [3 + 2] C N R2 N N synthetic methods have been disclosed.2 Among them, the + N N CF3 1,3-dipolar cycloaddition intermolecular 1,3-dipolar cycloaddition (DPC) reactions R2 N N R3 N N N represent a powerful and atom-economical approach to N CF3 1,3-dipole 2 dipolarophile N versatile trifluoromethyl-substituted N-heterocycles from R3 3 simple and easily available starting materials. In this context, b)This work: [3 + 3] DPC 2,2,2-trifluorodiazoethane (CF3CHN2) has emerged as an Ar CO2R CO2R F3C attractive trifluoromethyl-containing 1,3-dipole in recent CH R5 4 [3 + 3] DPC N years, as demonstrated by a series of DPC reactions with NH N N + N silver catalysis NH alkenes,5 allenes,6 alkynes,7 benzynes,8 aldimines,9 Ar N R4 Ar 10 11 RO C CF3 2 isocynides, and arene-diazonium salts, thus allowing facile CF3 1,3-dipole dipolarophile preparation of trifluoromethyl-substituted pyrazolines, H pyrazoles, triazoles, and tetrazoles (Scheme 1a). However, all CF3 OH CO2R Ar such studies only focused on the [3 + 2] DPC transformations H N N OH to give 5-membered trifluoromethyl-substituted N-heterocycles. H N N fast click ligation with TCOs Ar In sharp contrast, higher-order DPC reactions of CF3CHN2 have H MeO2C CF3 received much less attention.12 Indeed, there is still no report on employing CF3CHN2 in the direct [3 + 3] variant of DPC As part of our continuing interest in the chemistry of reactions until now, albeit the realization of such protocol CF3CHN2,14 we report herein a silver-catalyzed [3 + 3] DPC would offer great promise for the efficient construction of reaction of CF3CHN2 with in situ generated azomethine ylides valuable six-membered trifluoromethyl-substituted N(Scheme 1b).15 This strategy not only diversifies the existing heterocycles, such as CF3-bearing pyridines that have been DPC reaction patterns of CF3CHN2, but also provides a direct widely identified as important scaffolds with diverse biological approach to trifluoromethyl-substituted tetrahydrotriazines in 13 activities.

ACS Paragon Plus Environment

ACS Catalysis 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

moderate to excellent yields with exclusive regio- and diastereoselectivities. Furthermore, the afforded cycloadducts can be readily converted into a diverse array of trifluoromethylsubstituted 1,2,4-triazines and pyridines. Of particular importance, these unique trifluoromethyl-substituted 1,2,4triazines, simultaneously containing a tunable aryl moiety (Ar) and two electron-withdrawing groups (CF3 and CO2Me), hold great promise for translation into bioorthogonal reagents,16 as exemplified by the fast trans-bicyclononene (s-TCO) ligation with the second-order rate constants (k2) impressively amounting to 99.24 M-1 s-1. 2. RESULTS AND DISCUSSION Table 1. Optimization of Reaction Conditions a H N

F3C

N

CO2Me

1a

Cl

Entr y

+ CF3CHN2

catalyst / base solvent, temp, time 4-ClC H 6 4

N 2a

NH CO2Me

Catalyst (mol%)

Base (x equiv)

Solvent / Temp (oC) / Time (h)

Yield (%)b

1

~~

Cs2CO3 (0.5)

THF / 25 / 24

20

2

CuI (10)

Cs2CO3 (0.5)

THF / 25 / 24

0c

3

Cu(OAc)2 (10)

Cs2CO3 (0.5)

THF / 25 / 24

5c

4

Zn(OAc)2 (10)

Cs2CO3 (0.5)

THF / 25 / 24

47

5

FeCl3 (10)

Cs2CO3 (0.5)

THF / 25 / 24

22

6

NiBr2 (10)

Cs2CO3 (0.5)

THF / 25 / 24

34

7

Ag2O (10)

Cs2CO3 (0.5)

THF / 25 / 24

60

8

Ag2CO3 (10)

Cs2CO3 (0.5)

THF / 25 / 24

58

9

AgF (10)

Cs2CO3 (0.5)

THF / 25 / 24

81

10

AgNO3 (10)

Cs2CO3 (0.5)

THF / 25 / 24

51

11

AgOAc (10)

Cs2CO3 (0.5)

THF / 25 / 24

82

12

AgOAc (10)

Na2CO3 (0.5)

THF / 25 / 24

0c

13

AgOAc (10)

K2CO3 (0.5)

THF / 25 / 24

0c

14

AgOAc (10)

Et3N (0.5)

THF / 25 / 24

0c

15

AgOAc (10)

DBU (0.5)

THF / 25 / 24

57

16

AgOAc (10)

Cs2CO3 (0.75)

THF / 25 / 24

88

17

AgOAc (10)

Cs2CO3 (1.0)

THF / 25 / 24

90

18

AgOAc (10)

Cs2CO3 (1.0)

THF / 0 / 24

97

19

AgOAc (5)

Cs2CO3 (1.0)

THF / 0 / 48

85

20

AgOAc (10)

Cs2CO3 (1.0)

toluene / 0 / 24

5c

21

AgOAc (10)

Cs2CO3 (1.0)

CH3CN / 0 / 24

40

22

AgOAc (10)

Cs2CO3 (1.0)

CH2Cl2 / 0 / 24

0c

23

AgOAc (10)

Cs2CO3 (1.0)

DMF / 0 / 24

12c

24d

AgOAc (10)

Cs2CO3 (1.25)

THF / 0 / 24

84

25d

AgF (10)

Cs2CO3 (1.25)

THF / 0 / 24

90

26

d

AgOTf (10)

Cs2CO3 (1.25)

THF / 0 / 24

52

27d

Ag2CO3 (10)

Cs2CO3 (1.25)

THF / 0 / 24

65

a

Reaction conditions: 1a (0.4 mmol), CF3CHN2 (0.8 mmol), solvent (4.0 mL) (Method A). b Isolated yield. dr values (> 99:1) were determined by 19F NMR analysis. c Determined by 19F NMR analysis. d CF3CHN2 was in situ generated from trifluoroethylamine (0.8 mmol) (Method B).

Reaction Optimization. As a starting point to this study, we set out to investigate the DPC reaction of CF3CHN2 with methyl (E)-2-((4-chlorobenzylidene)amino)acetate 1a in the presence of Cs2CO3 at room temperature (Table 1). Despite in low conversion, the desired cycloadduct 2a was isolated with

excellent regio- and diastereoselectivity (entry 1). Next, a series of metal catalysts, including copper, zinc, iron, nickel, and silver salts, were screened for the model reaction (entries 2–11). Among these, it was found that silver salts performed as the most powerful promoters (entries 7–11). Particularly, when AgOAc was employed as the catalyst, the DPC reaction proceeded efficiently to produce cycloadduct 2a in 82% yield (entry 11). On the basis of this observation, further optimization was carried out by screening different bases and lowering the reaction temperature (entries 12–18). Pleasingly, the best result for the formation of 2a was achieved (97% yield, >99:1 dr, entry 18), by performing this reaction with AgOAc (10 mol %) and Cs2CO3 (1.0 equiv) in THF at 0 °C for 24 hours. When the amount of silver was reduced to 5 mol %, the transformation also underwent with good performance, albeit in prolonged reaction time (entry 19). In addition, the solvent was found to have an important effect on the reactivity (entries 20–23), among which THF proves to the one of choice. Encouraged by the seminal studies of Carreira,17 Mykhailiuk,18 and our group19 on in situ preparation of diazo compounds, we also investigated the feasibility for in situ CF3CHN2 generation and its follow-up one-pot reaction with 1a (entries 24-27). In this scenario, a comparable yield was attained in an organic medium with AgF as catalyst (90% yield, >99:1 dr, entry 25). Substrate Scope of the DPC Reaction of Glycine Imines with CF3CHN2. With the optimal reaction conditions in hand (Table 1, entries 17 and 25), the scope of this silver-catalyzed DPC reaction of CF3CHN2 with a variety of glycine imines 1 for the synthesis of CF3-substituted tetrahydrotriazines was investigated. As shown in Scheme 2, the use of the stock solution of CF3CHN2 (Method A) or the in situ generation of CF3CHN2 (Method B) furnished the desired products in comparable or complementary yields. In the case of phenylaldehyde-derived glycine imines, the DPC reaction tolerates various substitution patterns and a range of substituents on the phenyl ring. Halo-, trifluoromethyl-, nitro-, methoxycarbonyl-, cyano-, amino-, alkoxy-, alkyl-, and arylsubstituted phenyl glycine aldimines all underwent the desired reaction to afford cycloadducts 2a–v in good to high yields. The single regio- and stereoisomer of product 2b was readily obtained by simple recrystallization using ethyl acetate/petroleum. X-ray structure analysis revealed the trans relationship between the CF3 group and 4-BrC6H4 substituent at the tetrahydrotriazine ring.20 Polycyclic and heterocyclic glycine aldimines were found to be competent substrates, thereby delivering products 2w–b’ in 73–96% yields. Besides methyl ester substituent, glycine imines with other ester substituents, such as ethyl and tert-butyl groups, were also compatible with this reaction to give cycloadducts 2c’ and 2d’ in decent yields. Remarkably, the tri(perfluorotertbutyl)pentaerythritol-modulated alkyl group (PFC) containing 27 fluorine atoms with identical chemical environment and without adjacent hydrogen atom was readily incorporated (2e’ and 2g’), which provides an unique opportunity for disease diagnosis using 19F magnetic resonance imaging (MRI).21 Moreover, glycine aldimines derived from the star AIEgen tetraphenylethene (TPE) in the

Scheme 2. Substrate Scope of Cycloaddition Reaction of CF3CHN2 with Various Glycine Imine Esters 1

ACS Paragon Plus Environment

Page 2 of 8

Page 3 of 8 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

ACS Catalysis CF3CHN2 + N 1

Ar

HN F3C R

F3C

AgOAc or AgF, Cs2CO3

CO2R

H N

H N

Ar

N

THF, 0 oC, 24 h

H N

N

H N

F3C

NH

H N

F3C

H N

H N

H N

F3C

NH

N R = H, 2a', 81% (73%) R = CF3, 2b', (81%)

NH O

Cl

O

O

F3C

Ar

N

Cs2CO3, THF, 0 oC, 24 h

D2-2a, 96%

H N

Ar

N

Ar

H

N

N

OR'

F3C N N

F3C

R

OR'

N N TS

2f', 78% (92%)

B

O AgLn

O

N

O

O 2g', 71% (76%)

OC(CF3)3 4

HN

OC(CF3)3 F3C OC(CF3)3

CO2 Bu N

CO2R' N 1 AgOAc Cs2CO3

O

Cs2CO3 AgOAc

CF3

R

CsHCO3

N

R N LnAg

2h', 72% (92%)

H N

N

F3C

Ac,

OR'

AgO

N

1

Cs F3C

CO2R'

2

R

CO2R'

N

C

H

3 CO

N

R

N

R

N

O A

Ln = AcO or solvent

field of aggregation-induced emission (AIE)22 also underwent the desired transformation efficiently to forge products 2f’ and 2g’ in good yields, which further highlights the mild nature and broad functional group compatibility of this method. It should be noted that in all cases the diastereoselectivites of these reactions were excellent (> 99:1). In addition to glycine aldimines, glycine ketimine also worked well under the same reaction conditions to give the desired product 2h’. Mechanistic Studies and Proposed Mechanism. To shed light on the mechanism for this [3 + 3] DPC reaction, several isotopic-labelling experiments were carried out (Scheme 3). When deuterium-labeled CF3CHN2 (Scheme 3a) and benzaldehyde-derived glycine imine D1-1a (Scheme 3b) were used as reactants under the standard reaction conditions, the ratios of deuterium to hydrogen at C6- and C5-positions of the tetrahydrotriazine ring were found to be identical to that of the starting materials, indicating that no cleavage of C−H bonds occured at these positions from reactants to products. By using D2-1a as the reactant, we did not observe any incorporation of deuterium at the 1- and/or 2-position of the tetrahydrotriazine ring (Scheme 3c). In the control experiments, however, we detected the deuterated tetrahydrotriazine when D2O was used as additive (Scheme 3d). Therefore, the intermolecular proton Scheme 3. Isotopic-Labeling Experiments (Ar = 4-ClC6H4) and Proposed Mechanism

CsHCO3

AgLn CF3

NH

CO2Me

N

D3-2a, 99%

e)

t

H/D (33/67) N 1 2NH/D (31/69)

2a

CO2Me

CO2Me

2a, 90%

3

CO2Me

NH

H N

3

N

F3C

D2O (5 equiv), THF, rt, 2 h

NH

R

OC(CF3)3 OC(CF3)3

H/D (100/0) N 1 2 NH/D (100/0)

F3C C N N F3C AgOAc (10 mol%) H + (90/10)D/H H/D (10/90) Cs2CO3, THF, 0 oC, 24 h Ar N CO2Me Ar D2-1a H N

CO2Me

N

Ar

CO2Me

NH

6 5 4

(100) D

D1-1a

F3C

H N

F3C

AgOAc (10 mol%)

c)

OC(CF3)3 4

2e', 65% (68%) H N

F3C C N N H (100) D +

R = Et, 2c', 81% (85%) R = tBu, 2d', 77% (90%)

N O

b)

OR

N

CO2Me

N R

2y, 81% (75%)

NH

N

CO2Me

N

CO2Me

F3C F3C

NH

a) F3C H N F3C C N N NH 6 (50/50)D/H AgOAc (10 mol%) H/D (50/50) 5 4 o + Ar CO2Me N Cs2CO3, THF, 0 C, 24 h N CO2Me Ar D1-2a, 95% 1a

d)

NH

S 2z, 85% (93%)

R = 3-Cl, 2l, 78% (86%) R = 3-NO2, 2m, 80% (74%) R = 3-OMe, 2n, 87% (85%) R = 3-NBn2, 2o, 86% (90%) R = 2-Cl, 2p, 66% (89%) R = 2-Br, 2q, 75% (86%) R = 2-Me, 2r, 70% (92%) R = 2-MeO, 2s, 55% (64%) R = 3,4-Cl2, 2t, 72% (73%) R = 3,5-(CF3)2, 2u, (71%) R = 3,4,5-F3, 2v, (75%)

2x, 96% (95%) F3C

CO2Me

yield in Method A CO2R (yield in Method B)

NH

N

CO2Me

2w, 82% (96%)

H N

NH

2

R = 4-Cl, 2a, 97% (90%) R = 4-Br, 2b, 99% (80%) CO2Me R = H, 2c, 75% (79%) R = 4-F, 2d, 96% (85%) R = 4-CF3, 2e, 82% (79%) N R = 4-NO2, 2f, 76% (70%) R = 4-CN, 2g, 81% (72%) R = 4-CO2Me, 2h, 93% (99%) R = 4-Me, 2i, 77% (82%) R = 4-MeO, 2j, 50% (43%) R = 4-Ph, 2k, 76% (86%)

N

F3C

F3C

H N

NH

N 2

CO2R'

exchange between cycloadduct-NH and water is possible. To further probe the progression of this cycloaddition, we monitored the reaction mixture of glycine imine 1a, CF3CHN2, AgOAc, and Cs2CO3 in THF by 19F NMR spectroscopy. 19F NMR analysis of the reaction filtrate revealed the appearance of two new signals at −73.7 (doublet) and −67.2 (doublet) ppm (Figure S5 in the SI). The doublet at −73.7 ppm was identified as the expected cycloadduct 2a. Interestingly, when the reaction was complete, another doublet signal at −67.2 ppm disappeared, probably pointing to the corresponding cycloaddition intermediate species.23 Based on these preliminary results, a tentative reaction mechanism was deduced.24 As depicted in Scheme 3e, in the presence of AgOAc and Cs2CO3, glycine imine was initially converted into silver enolate A. Follow-up [3 + 3] DPC reaction between silver enolate A and CF3CHN2 occurred to give Agcomplex B with excellent regio- and diasteroselective control via a chair-like six-membered transition state (TS), wherein the trifluoromethyl group and two substituents of glycine imines all orientated equatorially to minimize steric interactions.25 Subsequent 1,3-hydrogen shift and protonation Scheme 4. Synthetic Transformation of

ACS Paragon Plus Environment

ACS Catalysis 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

Tetrahydrotriazines 2 into 6-Trifluoromethyl-1,2,4triazines 3 F3C

H N

Ar

N

NH

DDQ (4 equiv)

F3C

N

Ar

N 3

THF, 50 oC, 20 h

CO2R

2 N

N

R = 4-Cl, 3a, 80% R = 4-Br, 3b, 72% R = H, 3c, 83% R = 4-F, 3d, 70% R = 4-CF3, 3e, 67% R = 4-NO2, 3f, 42% R = 4-CN, 3g, 50%

CO2Me N

F3C R

N

N

N

CO2Me N

N

F3C

N

F3C

N

CO2Me

N CO2R

R = 4-CO2Me, 3h, 63% R = 4-Me, 3i, 86% R = 4-MeO, 3j, 92% R = 3-MeO, 3k, 77% R = 3,5-(CF3)2, 3l, 55% R = 3,4,5-F3, 3m, 61% N

CO2Me N

F3C

N

N

CO2Me N

F3C

S

3n, 64%

3o, 66% N

F3C

F3C

N

F3C

N

N

3p, 54%

CO2Me

N

3q, 57%

N O

N

O

4

OC(CF3)3 OC(CF3)3

O 3r, 72% F3C

3t, 72% N N

N CO2Me

F3C

N

N O

N O

3s, 72%

OC(CF3)3

O

OC(CF3)3 4

OC(CF3)3 OC(CF3)3

3u, 90%

assisted by CsHCO3 proceeded rapidly to produce a relatively stable intermediates C. This hydrogen shift pathway was verified to some extent by the fact that no product was generated for alanine- or phenylalanine-derived imines, among which the hydrogen in intermediates B was replaced by methyl or phenyl group, respectively. Finally, another 1,3-hydrogen shift (isomerization) of C occurred to afford the desired product 2. Further Synthetic Transformations. Trifluoromethylated tetrahydrotriazines 2 are versatile synthetic intermediates, and direct oxidation of a series of tetrahydrotriazines using DDQ in THF was examined (Scheme 4). To our delight, tetrahydrotriazines 2 with a broad spectrum of substituents, including phenyl, naphthyl, anthracenyl, thiophenyl and pyridyl, were smoothly oxidized to the corresponding 6-trifluoromethyl1,2,4-triazines 3a–r in moderate to good yields (42–92%). Several tetrahydrotriazines with TPE and/or PFC functional groups also underwent this oxidation transformation successfully, and the expected products 3s–u were isolated in good to high yields. 3-Trifluoromethylpyridine is the core unit of many drugs, agrochemicals, and related candidates. Further synthetic transformation of 1,2,4-triazines 3 into 3trifluoromethylpyridines 4 was achieved by using various dienophile partners (Scheme 5). For example, performing the reaction of 1,2,4-triazines with 2,5-norbornadiene or alkynes in toluene at 110 oC readily delivered the corresponding pyridines 4a–f in 50–99% yields.26 Treatment of 1,2,4-triazines with aliphatic aldehydes by means of Boger’s protocol afforded the highly substituted pyridines 4g–j in good to high yields with

excellent regioselectivities.27 Compound 4g was subjected to Xray crystallographic analysis to further confirm the structure of these molecules.20 Both of cyclopentanone and cyclohexanone were also smoothly coupled with triazines to give products 4k and 4l in 99% and 83% yields, respectively. These promising results indicate that the present protocol provides a reliable and efficient approach for the synthesis of diverse CF3-containing N-heterocycles. Evaluation of the Reactivity of Trifluromethyl 1,2,4triazines 3 with TCOs. Recently, 1,2,4-triazines have been identified as an alternative class of bioorthogonal reagents, but they are still struggling for wide applications due to the limited reaction kinetics.28 We noted that the 6-trifluoromethyl-1,2,4triazines 3 obtained by our method contain a tunable aryl moiety and two electron-withdrawing substituents simultaneously, which quite possibly imparts them good reactivity towards trans-cyclooctene (TCO) derivatives, and thus making them suitable as a novel class of candidates in bioorthogonal reactions.14 To verify the feasibility of these 1,2,4-triazines as biorthogonal reagents, we first conducted a preliminary computational study to predict their reactivity in the inverseelectron-demand Diels–Alder (IEDDA) reactions with TCO. As shown in Scheme 6, the frontier molecular orbital analysis for three previously reported 1,2,4-triazines T1–3 and our newly developed 3f were performed. It turned out that trifluoromethyl-substituted 1,2,4-triazine 3f possesses the lowest energy in the lowest unoccupied molecular orbital (LUMO), indicating that 3f could exhibit exceedingly higher reaction kinetics in the IEDDA reactions compared with T1–3. Furthermore, the activation free energies (ΔG) for the IEDDA reaction between trifluoromethyl-substituted 1,2,4-triazine 3c and TCO was calculated to be only 20.4 kcal / mol. These results indicated that additional improvement on the TCO ligation reactivity might be feasible by fine-tuning the substituents on the triazine ring. Encouraged by these computational analyses, we then investigated the reactivity of CF3-1,2,4-triazines 3 towards TCO or s-TCO by monitoring the reactions with 19F NMR or UV detector. In all cases, the air-oxidized pyridine derivatives were observed, which is consistent with DFT calculations (Figure S8 in the SI). As shown in Table 2 and Table S1 (in the SI), triazines with strong electron-withdrawing substituent (3e–h and 3r) on the phenyl ring exhibited faster reaction rates than those with electron-neutral or donating groups (3c, 3i, and 3j), which is in consistent with the nature of IEDDA reactions. For triazines 3c–j, the ρ value was determined to be 0.68 via Hammett analysis of the IEDDA reaction rate constants (Figure 1), suggesting that only partial charge separation occurs during the reaction. Notably, the use of s-TCO as the reaction partner led to distinct improvements in reactivity (Table 2). For example, a second-order rate constant of 68.27 M-1 s-1 was recorded for the s-TCO ligation with 4-CF3-Ph-substituted CF3triazine 3e. Remarkably, the highest second-order rate constant (k2 = 99.24 M-1 s-1) was observed when 4-NO2-Ph-substituted CF3-triazine 3f was ligated with s-TCO, which represents the fastest reaction kinetics for the IEDDA transformations involving 1,2,4-triazines to date.

ACS Paragon Plus Environment

Page 4 of 8

Page 5 of 8 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

ACS Catalysis

Scheme 5. Further Transformation of 1,2,4-Triazines 3 into 3-Trifluoromethylpyridines 4. F3C

F3C N

F3C N

CO2Me

4a, 95%

N

CO2Me

4b, 82%

Cl

Cl

F3C

toluene, 110 oC F3C N

CO2Me

N

CO2Me

N

F3C

N

Ar

N

Cl CF3

CO2Me 4e, 50%

( )n

benzyne or phenylacetylene toluene, 110 oC

MeO2C Ph 4f, 83%

N

CO2Me Cl

4j 98%

CO2Me

CO2Me

N

N

O

N

Bn CO2 Me

CO2Me

CF3

N

F3C

Hex

N

3

Cl

CO2 Me 4h, 89%

Cl

4i, 92%

Cl

S 4d, 93%

4c, 99%

n

F3C

Pr

N

CO2Me

4g, 85%

RCH2CHO o pyrrolidine, 4A MS toluene, 110 oC

i

F3C

Me

(n = 1, 2) CF3

o

pyrrolidine, 4A MS CHCl3, 45 oC

CF3

Cl

4k, 99%

Cl

4l, 83%

Scheme 6. Preliminary Computational Studies Table 2. Second-Order Rate Constants (k2 in M-1 s-1) of the Click Ligations between CF3-1,2,4-Triazines 3 and Strained Alkenes

a) LUMO energies of various 1,2,4-triazinesa N N E (eV)

N N

N

N N

N

CO2Me N N

N

CF3 Ar

CF3

N

NO2

OH

N N

N

NO2

+

CO2Me

T1

T2

T3

Ar

-1.91

CF3 Ph

G N N

TS-1 20.4

N

F3C 3 Ph

G G

0 SM

G

N N N IM-1' -5.0

N N CF 3 N Ph

CF3 Ph

N N N

+ H s-TCO

OH

OH

k2

F3C

CH3CN Ar H 2O

N

CO2Me

Ar

[O] H

N CO2Me

Entry

1,2,4-triazine (Ar)

k2 with TCO

1 2 3 4 5 6

3c (Ph) 3e (4-CF3-Ph) 3f (4-NO2-Ph) 3g (4-CN-Ph) 3h (4-CO2Me-Ph) 3i (4-MePh)

0.30 ± 0.04 0.70 ± 0.09 1.00 ± 0.06 0.85 ± 0.15 0.39 ± 0.07 0.20 ± 0.04

OH

H F3C H

Ar N CO2Me

k2 with s-TCO 32.55 ± 4.95 68.27 ± 5.80 99.24 ± 4.44 81.19 ± 3.44 63.27 ± 3.02 26.64 ± 0.52

CF3 N Ph N N

TS-2 -1.6

G G

F3C

TS-2' -0.1

IM-1 -6.7

SM = Starting Material IM = Intermediate TS = Transition State

H

CO2Me

b) energetics of CF3-1,2,4-triazine 3c with TCOb

TS-1' 22.5

N N

N

-2.36

N N N

N

H

CF3

G

[O] CO2Me

Ar

-1.33

Gsol (kcal/mol) G = CO2Me

TCO

k2 CH3CN F C 3 D 2O

3f

-0.97

LUMO

HO

HO

N

CF3

CF3 N N N Ph

N2 IM-2 -46.9

Ph H CF3 Ph H G

N

aCalculated at M06-2X/6-31G(d) level. bCalculated at B3LYP/6-31G(d, p) // (SMD)M06-L/6-31++G(d, p) level.

Figure 1. Hammet plot for the reactions between TCO and a panel of CF3-1,2,4-triazines 3

ACS Paragon Plus Environment

ACS Catalysis 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

3. CONCLUSIONS In conclusion, we have successfully developed an efficient silver-catalyzed [3 + 3] DPC reaction of CF3CHN2 with glycine imines. Under simple and mild conditions, the reaction proceeded smoothly to afford a variety of trifluoromethylsubstituted tetrahydrotriazines with excellent regio- and diastereoselectivities. Furthermore, facile transformation of the obtained CF3-tetrahydrotriazines provides a novel process to formulate highly functionalized triazine and pyridine architectures. More importantly, the high reactivity of these triplet-activated triazines is demonstrated by the fast click ligation with TCOs (k2 up to 99.24 M-1 s-1), thus exhibiting exceptional promise for translation into bioorthogonal reagents. Future investigations including mechanistic elucidation, substrate scope expansion, as well as in vitro and in vivo applications in bioconjugate chemistry are ongoing in our laboratory, and these results will be reported in due course.



ASSOCIATED CONTENT



Supporting Information

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscatal.xxxx. Experimental procedures, compound characterization data, computation details, and 1H, 19F and 13C NMR spectra for new compounds (PDF) Crystallographic data of 2b (CIF) Crystallographic data of 4g (CIF)



AUTHOR INFORMATION

Corresponding Author *E-mail: [email protected] *E-mail: [email protected]

ORCID Jun-An Ma: 0000-0002-3902-6799 Fa-Guang Zhang: 0000-0002-0251-0456 Zhen Chen: 0000-0002-6289-4332

Author Contributions §Z.

C. and N. R. contributed equally to this work.

Notes The authors declare no competing financial interest.



ACKNOWLEDGMENT

This work was supported by the National Natural Science Foundation of China (No. 21472137, 21532008, and 21772142), the National Basic Research Program of China (973 Program, 2014CB745100), and Tianjin Municipal Science & Technology Commission (14JCZDJC33400). We thank Dr. Wei Meng of Institute of Chemistry (CAS) for help with computational studies.



REFERENCES

(1) For selected reviews, see: (a) Lamberth, C. Pyrazole Chemistry in Crop Protection. Heterocycles 2007, 71, 1467–1502. (b) Yet, L. Comprehensive Heterocyclic Chemistry III, 1st ed., Elsevier: Oxford, 2008, 4, 1. (c) Alvarez-Builla, J.; Vaquero, J. J.; Barluenga, J. Modern Heterocyclic Chemistry, 1st ed., Wiley-VCH: Weinheim, 2011. (d)

Page 6 of 8

Jeschke, P. Progress of Modern Agricultural Chemistry and Future Prospects. Pest Manag. Sci. 2015, 72, 433–455. (2) For selected reviews, see: (a) Petrov, V. A. Fluorinated Heterocyclic Compounds: Synthesis, Chemistry, and Applications; DuPont Central Research and Development, Wilmington, 2009. (b) Fustero, S.; Sánchez-Roselló, M.; Barrio, P.; Simón-Fuentes, A. From 2000 to Mid-2010: A Fruitful Decade for the Synthesis of Pyrazoles. Chem. Rev. 2011, 111, 6984–7034. (c) Nie, J.; Guo, H.-C.; Cahard, D.; Ma, J.-A. Asymmetric Construction of Stereogenic Carbon Centers Featuring a Trifluoromethyl Group from Prochiral Trifluoromethylated Substrates. Chem. Rev. 2011, 111, 455–529. (d) Tomashenko, O. A.; Grushin, V. V. Aromatic Trifluoromethylation with Metal Complexes. Chem. Rev. 2011, 111, 4475–4521. (e) Kawai, H.; Shibata, N. Asymmetric Synthesis of Agrochemically Attractive Trifluoromethylated Dihydroazoles and Related Compounds under Organocatalysis. Chem. Rec. 2014, 14, 1024–1040. (f) Egami, H.; Sodeoka, M. Trifluoromethylation of Alkenes with Concomitant Introduction of Additional Functional Groups. Angew. Chem., Int. Ed. 2014, 53, 8294–8308. (g) Besset, T.; Schneider, C.; Cahard, D. Tamed Arene and Heteroarene Trifluoromethylation. Angew. Chem., Int. Ed. 2014, 53, 8294–8308. (h) Huang, Y.-Y.; Yang, X.; Chen, Z.; Verpoort, F.; Shibata, N. Catalytic Asymmetric Synthesis of Enantioenriched Heterocycles Bearing a C-CF3 Stereogenic Center. Chem. Eur. J. 2015, 21, 8664–8684. (i) Li, S.; Ma, J.-A. Core-Structure-Inspired Asymmetric Addition Reactions: Enantioselective Synthesis of Dihydrobenzoxazinone- and Dihydroquinazolinone-Based Anti-HIV Agents. Chem. Soc. Rev. 2015, 44, 7439–7448. (3) For selected reviews, see: (a) Gothelf, K. V.; Jørgensen, K. A. Asymmetric 1,3-Dipolar Cycloaddition Reactions. Chem. Rev. 1998, 98, 863–909. (b) Gothelf, K. V. In Cycloaddition Reactions in Organic Synthesis. Kobayashi, S.; Jørgensen, K. A., Eds. Wiley-VCH: Weinheim, 2002; pp 211–247. (c) Padwa, A.; Pearson, W. H. Synthetic Applications of 1,3-Dipolar Cycloaddition Chemistry—Toward Heterocycles and Natural Products. Wiley: Hoboken, 2003. (d) Hashimoto, T.; Maruoka, K. In Handbook of Cyclization Reactions, Ma, S., Ed. Wiley-VCH: Weinheim, 2009; Vol. 3, pp 87–168. (e) Cheng, Q.-Q.; Deng, Y.; Lankelma, M.; Doyle, M. P. Cycloaddition reactions of enoldiazo compounds. Chem. Soc. Rev. 2017, 46, 5425–5443. (4) For recent reviews, see: (a) Grygorenko, O. O.; Artamonov, O. S.; Komarov, I. V.; Mykhailiuk, P. K. Trifluoromethyl-Substituted Cyclopropanes. Tetrahedron 2011, 67, 803–823. (b) Qiu, D.; Qiu, M.; Ma, R.; Zhang, Y.; Wang, J. Nitrogen Group Retaining Reaction in the Transformation of Diazo Compounds. Acta Chim. Sinica 2016, 74, 472–487. (c) Mertens, L.; Koenigs, R. M. Fluorinated diazoalkanes – a versatile class of reagents for the synthesis of fluorinated compound. Org. Biomol. Chem. 2016, 14, 10547–10556. For selected examples, see: (d) Mykhailiuk, P. K.; Afonin, S.; Palamarchuk, G. V.; Shishkin, O. V.; Ulrich, A. S.; Komarov, I. V. Synthesis of TrifluoromethylSubstituted Proline Analogues as 19F NMR Labels for Peptides in the Polyproline II Conformation. Angew. Chem., Int. Ed. 2008, 47, 5765– 5767. (e) Argintaru, O. A.; Ryu, D.; Aron, I.; Molander, G. A. Synthesis and Applications of α-Trifluoromethylated Alkylboron Compounds. Angew. Chem., Int. Ed. 2013, 52, 13656–13660. (f) Molander, G. A.; Ryu, D. Diastereoselective Synthesis of Vicinally Bis(trifluoromethylated) Alkylboron Compounds through Successive Insertions of 2,2,2-Trifluorodiazoethane. Angew Chem., Int. Ed. 2014, 53, 14181–14185. (g) Wu, G.; Deng, Y.; Wu, C.; Wang, X.; Zhang, Y.; Wang, J. Switchable 2,2,2-Trifluoroethylation and gemDifluorovinylation of Organoboronic Acids with 2,2,2Trifluorodiazoethane. Eur. J. Org. Chem. 2014, 4477–4481. (h) Luo, H.; Wu, G.; Zhang, Y.; Wang, J. Silver(I)-Catalyzed NTrifluoroethylation of Anilines and O-Trifluoroethylation of Amides with 2,2,2-Trifluorodiazoethane. Angew. Chem., Int. Ed. 2015, 54, 14503–14507. (i) Hyde, S.; Veliks, J.; Liégault, B.; Grassi, D.; Taillefer, M.; Gouverneur, V. Copper-Catalyzed Insertion into HeteroatomHydrogen Bonds with Trifluorodiazoalkanes. Angew. Chem., Int. Ed. 2016, 55, 3785–3789. (j) Peng, S.-Q.; Zhang, X.-W.; Zhang, L.; Hu, X.-G. Esterification of Carboxylic Acids with Difluoromethyl Diazomethane and Interrupted Esterification with Trifluoromethyl Diazomethane: A Fluorine Effect. Org. Lett. 2017, 19, 5689–5692. (k) Zhang, X.-W.; Hu, W.-L.; Chen, S.; Hu, X.-G. Cu-Catalyzed Synthesis

ACS Paragon Plus Environment

Page 7 of 8 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

ACS Catalysis of Fluoroalkylated Isoxazoles from Commercially Available Amines and Alkynes. Org. Lett. 2018, 20, 860–863. (l) Pieber, B.; Kappe, C. O. Generation and Synthetic Application of Trifluoromethyl Diazomethane Utilizing Continuous Flow Technologies. Org. Lett. 2016, 18, 1076–1079. (m) Arkhipov, A. V.; Arkhipov, V. V.; Cossy, J.; Kovtunenko, V. O.; Mykhailiuk, P. K. Unexpected Reactivity of Trifluoromethyl Diazomethane (CF3CHN2): Electrophilicity of the Terminal N‑Atom. Org. Lett. 2016, 18, 3406– 3409. (n) Kotozaki, M.; Chanthamath, S.; Fujii, T.; Shibatomi, K.; Iwasa, S. Chem. Commun. 2018, 54, 5110–5113. (5) (a) Artamonov, O. S.; Mykhailiuk, P. K.; Voievoda, N. M.; Volochnyuk, D. M.; Komarov, I. V. Simple and Efficient Procedure for a Multigram Synthesis of Both trans- and cis-1-Amino-2(trifluoromethyl)cyclopropane-1-carboxylic Acid. Synthesis 2010, 443–446. (b) Slobodyanyuk, E. V.; Artamonov, O. S.; Shishkin, O. V.; Mykhailiuk, P. K. One-Pot Synthesis of CF3-Substituted Pyrazolines/Pyrazoles from Electron-Deficient Alkenes/Alkynes and CF3CHN2 Generated in situ: Optimized Synthesis of Tris(trifluoromethyl)pyrazole. Eur. J. Org. Chem. 2014, 2487–2495. (c) Li, T.-R.; Duan,S.-W.; Ding, W.; Liu, Y.-Y.; Chen, J.-R.; Lu, L.-Q.; Xiao, W.-J. Synthesis of CF3-Containing 3,3′-Cyclopropyl Spirooxindoles by Sequential [3 + 2] Cycloaddition/Ring Contraction of Ylideneoxindoles with 2,2,2-Trifluorodiazoethane. J. Org. Chem. 2014, 79, 2296–2302. (d) Chen, Z.; Zheng, Y.; Ma, J.-A. Use of a Traceless Activating and Directing Group for the Construction of Trifluoromethylpyrazoles: One-Pot Transformation of Nitroolefins and Trifluorodiazoethane. Angew. Chem., Int. Ed. 2017, 56, 4569–4574. (e) Li, S.; Cao, W.-J.; Ma, J.-A. Stereoselective Cycloaddition of 2,2,2Trifluorodiazoethane with α-Methylene-β-lactams: Facile Synthesis of Trifluoromethyl-Substituted Spirocyclic β-Lactams. Synlett 2017, 28, 673–678. (f) Qin, S.; Zheng, Y.; Zhang, F.-G.; Ma, J.-A. One-Pot Cascade Transformations of Zinc Trifluorodiazoethylide and α,βUnsaturated Enones: Access to Trifluoromethylated Polycyclic Pyrazolines. Org. Lett. 2017, 19, 3406–3409. (6) Zhang, F.-G.; Wei, Y.; Yi, Y.-P.; Nie, J.; Ma, J.-A. Regioselective Cycloaddition of Trifluorodiazoethane with ElectronDeficient Allenic Esters and Ketones: Access to CF3-Substituted Pyrazolines and Pyrazoles. Org. Lett. 2014, 16, 3122–3125. (7) (a) Li, F.; Nie, J.; Sun, L.; Zheng, Y.; Ma, J.-A. Silver-Mediated Cycloaddition of Alkynes with CF3CHN2: Highly Regioselective Synthesis of 3-Trifluoromethylpyrazoles. Angew. Chem., Int. Ed. 2013, 52, 6255–6258. (b) Mykhailiuk, P. K. Three-Component Synthesis of Fluorinated Pyrazoles from Fluoroalkylamines, NaNO2 and ElectronDeficient Alkynes. Org. Biomol. Chem. 2015, 13, 3438–3445. (c) Mertens, L. K.; Hock, J.; Roenigs, R. M. Fluoroalkyl-Substituted Diazomethanes and Their Application in a General Synthesis of Pyrazoles and Pyrazolines. Chem. Eur. J. 2016, 22, 9542–9545. (d) Britton, J.; Jamison, T. F. A Unified Continuous Flow Assembly-Line Synthesis of Highly Substituted Pyrazoles and Pyrazolines. Angew. Chem., Int. Ed. 2017, 56, 8823–8827. (8) Sun, L.; Nie, J.; Zheng, Y.; Ma, J.-A. [3 + 2] Cycloaddition of Arynes with CF3CHN2: Access to 3-Trifluoromethyl-1H-indazoles. J. Fluorine Chem. 2015, 174, 88–94. (9) (a) Künzi, S. A.; Morandi, B.; Carreira, E. M. Preparation of TrifluoromethylSubstituted Aziridines with in Situ Generated CF3CHN2. Org. Lett. 2012, 14, 1900–1901. (b) Chai, Z.; Bouillon, J. P.; Cahard, D. Chiral Brønsted Acid-Catalyzed Diastereo- and Enantioselective Synthesis of CF3-Substituted Aziridines. Chem. Commun. 2012, 48, 9471–9473. (10) Wang, S.; Yang, L.-J.; Zeng, J.-L.; Zheng, Y.; Ma, J.-A. SilverCatalyzed [3 + 2] Cycloaddition of Isocyanides with Diazo Compounds: New Regioselective Access to 1,4-Disubstituted-1,2,3-triazoles. Org. Chem. Front. 2015, 2, 1468–1474. (11) Chen, Z.; Fan, S.-Q.; Zheng, Y.; Ma, J.-A. Silver-Catalyzed Regioselective [3 + 2] Cycloaddition of Arene-diazonium Salts with 2,2,2-Trifluorodiazoethane (CF3CHN2): A Facile Access to 2-Aryl-5trifluoromethyltetrazoles. Chem. Commun. 2015, 51, 16545–16548. (12) There are two formal [3 + 3] cyclization reactions of CF3CHN2 reported before, but none was proceeding via direct [3 + 3] DPC process: (a) Mao, H.; Lin, A.; Tang, Z.; Hu, H.; Zhu, C.; Cheng, Y. Organocatalytic One-Pot Synthesis of Highly Substituted Pyridazines

from Morita-Baylis-Hillman Carbonates and Diazo Compounds. Chem. Eur. J. 2014, 20, 2454–2458. (b) Feraldi-Xypolia, A.; Fredj, G.; Tran, G.; Tsuchiya, T.; Vors, J.-P.; Mykhailiuk, P.; Pardo, D. G.; Cossy, J. Synthesis of Functionalized 4-Fluoropyridazine. Asian J. Org. Chem. 2017, 6, 927–935. (13) (a) Roughley, S. D.; Jordan, A. M. The Medicinal Chemist’s Toolbox: An Analysis of Reactions Used in the Pursuit of Drug Candidates. J. Med. Chem. 2011, 54, 3451–3479. (b) Prachayasittikul, S.; Pingaew, R.; Worachartcheewan, A.; Sinthupoom, N.; Prachayasittikul, V.; Ruchirawat, S.; Prachayasittikul, V. Roles of Pyridine and Pyrimidine Derivatives as Privileged Scaffolds in Anticancer Agents. Mini-Rev. Med. Chem. 2017, 17, 869–901. (c) Burriss, A.; Edmunds, A. J. F.; Emery, D.; Hall, R. G.; Jacob, O.; Schaetzer, J. The importance of trifluoromethyl pyridines in crop protection. Pest Manag. Sci. 2018, 74, 1228–1238. (14) (a) Liu, C.-B.; Meng, W.; Li, F.; Wang, S.; Nie, J.; Ma, J.-A. A Facile Parallel Synthesis of Trifluoroethyl-Substituted Alkynes. Angew. Chem., Int. Ed. 2012, 51, 6227–6230. (b) Guo, R.; Lv, N.; Zhang, F.G.; Ma, J.-A. Zinc-Mediated Mannich-Type Reaction of 2,2,2Trifluorodiazoethane with Imines: Access to β-CF3-Amines. Org. Lett. 2018, 20, 6994–6997. (c) Zhang, F.-G.; Zeng, J.-L.; Tian, Y.-Q.; Zheng, Y.; Cahard, D.; Ma, J.-A. Telescoping Reactions with Trifluorodiazoethane-Derived Aza-Wittig Reagents and Allenyl esters. Chem.-Eur. J. 2018, 24, 7749–7754. (d) Zhang, F.-G.; Lv, N.; Zheng, Y.; Ma, J.-A. Phosphine-Relayed Aldehyde-Olefination and AzaWittig Reaction with 2,2,2-Trifluorodiazoethane. Chin. J. Chem. 2018, 36, 723–730. (15) For recent reviews on DPC reactions with azomethine ylides, see: (a) Adrio, J.; Carretero, J. C. Recent advances in the catalytic asymmetric 1,3-dipolar cycloaddition of azomethine ylides. Chem. Commun. 2014, 50, 12434–12446. (b) Fang, X.; Wang, C.-J. Catalytic asymmetric construction of spiropyrrolidines via 1,3-dipolar cycloaddition of azomethine ylides. Org. Biomol. Chem. 2018, 16, 2591–2601. (c) Tang, S.; Zhang, X.; Sun, J.; Niu, D.; Chruma, J. J. 2‑Azaallyl Anions, 2‑Azaallyl Cations, 2‑Azaallyl Radicals, and Azomethine Ylides. Chem. Rev. 2018, 118, 10393−10457. For selected examples: (d) Na, R.; Jing, C.; Xu, Q.; Jiang, H.; Wu, X.; Shi, J.; Zhong, J.; Wang, M.; Benitez, D.; Tkatchouk, E.; Goddard, W. A., III; Guo, H.; Kwon, O. Phosphine-Catalyzed Annulations of Azomethine Imines: Allene-Dependent [3 + 2], [3 + 3], [4 + 3], and [3 + 2 + 3] Pathways. J. Am. Chem. Soc. 2011, 133, 13337–13348. (e) Meng, W.; Zhao, H.-T.; Nie, J.; Zheng, Y.; Fu, A.; Ma, J.-A. Allenoate-derived 1,5-, 1,7-, and 1,9-zwitterions as highly versatile coupling-partners for phosphinetriggered cycloaddition reactions. Chem. Sci. 2012, 3, 3053–3057. (f) Li, Q.-H.; Wei, L.; Wang, C.-J. Catalytic Asymmetric 1,3-Dipolar [3 + 6] Cycloaddition of Azomethine Ylides with 2-Acyl Cycloheptatrienes: Efficient Construction of Bridged Heterocycles Bearing Piperidine Moiety. J. Am. Chem. Soc. 2014, 136, 8685–8692. (g) Teng, H.-L.; Yao, L.; Wang, C.-J. Cu(I)-Catalyzed Regio- and Stereoselective [6 + 3] Cycloaddition of Azomethine Ylides with Tropone: An Efficient Asymmetric Access to Bridged Azabicyclo[4.3.1]decadienes. J. Am. Chem. Soc. 2014, 136, 4075–4080. (h) Tong, M.-C.; Chen, X.; Tao, H.Y.; Wang, C.-J. Catalytic Asymmetric 1,3-Dipolar Cycloaddition of Two Different Ylides: Facile Access to Chiral 1,2,4-Triazinane Frameworks. Angew. Chem., Int. Ed. 2013, 52, 12377–12380. (i) Guo, H.; Liu, H.; Zhu, F.-L.; Na, R.; Jiang, H.; Wu, Y.; Zhang, L.; Li, Z.; Yu, H.; Wang, B.; Xiao, Y.; Hu, X.-P.; Wang, M. Enantioselective Copper-Catalyzed [3 + 3] Cycloaddition of Azomethine Ylides with Azomethine Imines. Angew. Chem., Int. Ed. 2013, 52, 12641–12645. (16) For selected reviews, see: (a) Vocadlo, D. J.; Hang, H. C.; Kim, E.-J.; Hanover, J.-A.; Bertozzi, C. R. A Chemical Approach for Identifying O-GlcNAc-Modified Proteins in Cells. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 9116–9121. (b) Ramil, C. P.; Lin, Q. Bioorthogonal Chemistry: Strategies and Recent Developments. Chem. Commun. 2013, 49, 11007–11022. (c) Takaoka, Y.; Ojida, A.; Hamachi, I. Protein Organic Chemistry and Applications for Labeling and Engineering in Live-Cell Systems. Angew. Chem., Int. Ed. 2013, 52, 4088–4106. (d) McKay, C. S.; Finn, M. G. Click Chemistry in Complex Mixtures: Bioorthogonal Bioconjugation. Chem. Biol. 2014, 21, 1075– 1101. (e) Spicer, C. D.; Davis, B. G. Selective Chemical Protein Modification. Nat. Commun. 2014, 5, 4740–4754. (f) Yang, M.-Y.; Li,

ACS Paragon Plus Environment

ACS Catalysis 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

J.; Chen, P. R. Transition Metal-Mediated Bioorthogonal Protein Chemistry in Living Cells. Chem. Soc. Rev. 2014, 43, 6511−6526. (g) Li, J.; Chen, P. R. Development and Application of Bond Cleavage Reactions in Bioorthogonal Chemistry. Nat. Chem. Biol. 2016, 12, 129–137. (h) Zhang, S.; He, D.; Lin, Z.; Yang, Y.; Song, H.-P.; Chen, P. R. Conditional Chaperone−Client Interactions Revealed by Genetically Encoded Photo-cross-linkers. Acc. Chem. Res. 2017, 50, 1184−1192. (i) Wu, H.; Devaraj, N. K. Advances in Tetrazine Bioorthogonal Chemistry Driven by the Synthesis of Novel Tetrazines and Dienophiles. Acc. Chem. Res. 2018, 51, 1249–1259. (j) Row, R. D.; Prescher, J. A. Constructing New Bioorthogonal Reagents and Reactions. Acc. Chem. Res. 2018, 51, 1073–1081. (k) N. K. Devaraj, The Future of Bioorthogonal Chemistry. ACS Cent. Sci. 2018, 4, 952– 959. (17) (a) Morandi, B.; Carreira, E. M. Iron-Catalyzed Cyclopropanation with Trifluoroethylamine Hydrochloride and Olefins in Aqueous Media: In Situ Generation of Trifluoromethyl Diazomethane. Angew. Chem., Int. Ed. 2010, 49, 938–941. (b) Morandi, B.; Carreira, E. M. Rhodium‐Catalyzed Cyclopropenation of Alkynes: Synthesis of Trifluoromethyl-Substituted Cyclopropenes. Angew. Chem., Int. Ed. 2010, 49, 4294–4296. (c) Morandi, B.; Mariampillai, B.; Carreira, E. M. Enantioselective Cobalt-Catalyzed Preparation of Trifluoromethyl-Substituted Cyclopropanes. Angew. Chem. Int. Ed. 2011, 50, 1101–1104. (d) Morandi, B.; Carreira, E. M. Synthesis of Trifluoroethyl-Substituted Ketones from Aldehydes and Cyclohexanones. Angew. Chem., Int. Ed. 2011, 50, 9085–9088. (18) Mykhailiuk, P. K. Generation of C2F5CHN2 In Situ and Its First Reaction: [3 + 2] Cycloaddition with Alkenes. Chem. Eur. J. 2014, 20, 4942–4947. (b) Mykhailiuk, P. K. In Situ Generation of Difluoromethyl Diazomethane for [3 + 2] Cycloadditions with Alkynes. Angew. Chem., Int. Ed. 2015, 54, 6558–6561. (19) Chen, Z.; Zhang, Y.; Nie, J.; Ma, J.-A. Transition-Metal-Free [3 + 2] Cycloaddition of Nitroolefins and Diazoacetonitrile: A Facile Access to Multisubstituted Cyanopyrazoles. Org Lett. 2018, 20, 2120– 2124. (20) The X-ray crystallographic structures for 2b and 4g have been deposited at the Cambridge Crystallographic Data Centre (CCDC), under deposition numbers CCDC 1408486 and 1572352. The data can be obtained free of charge from the Cambridge Crystallographic Data Centre via http://www.ccdc. cam.ac.uk/data_request/cif. (21) Perfluorocarbons (PFCs) have been potent molecular probes for magnetic resonance imaging (MRI). For selected reviews, see: (a) Riess, J. G. Oxygen Carriers (“Blood Substitutes”) Raison d'Etre, Chemistry, and Some Physiology Blut ist ein ganz besondrer Saft. Chem. Rev. 2001, 101, 2797–2920. (b) Tirotta, I.; Dichiarante, V.; Pigliacelli, C.; Cavallo, G.; Terraneo, G.; Bombelli, F.; Metrangolo, B. P.; Resnati, G. 19F Magnetic Resonance Imaging (MRI): From Design of Materials to Clinical Applications. Chem. Rev. 2015, 115, 1106– 1129. For our previous work, see: (c) Shi, H.; Lai, B.; Chen, S.; Zhou, X.; Nie, J.; Ma, J.-A. Facile Synthesis of Novel PerfluorocarbonModulated 4-Anilinoquinazoline Analogues. Chin. J. Chem. 2017, 35, 1693–1700. (22) The tetraphenylethene (TPE) skeleton is one of the most important luminogens in the realm of aggregation-induced emission (AIE), see: (a) Hong, Y.; Lam, J. W. Y.; Tang, B. Z. AggregationInduced Emission. Chem. Soc. Rev. 2011, 40, 5361–5388. (b) Aggregation-Induced Emission: Together We Shine, United We Soar! Mei, J.; Leung, N. L. C.; Kwok, R. T. K.; Lam, J. W. Y.; Tang, B. Z. Chem. Rev. 2015, 115, 11718–11940. (23) All attempts failed in the use of the mass spectrometry (MS) for the further identification of this putative intermediate. (24) For selected previous mechanistic studies on silver catalysis, see (a) Pellissier, H. Enantioselective Silver-Catalyzed Transformations. Chem. Rev. 2016, 116, 14868–14917. (b) FunesArdoiz, I.; Maseras, F. Oxidative Coupling Mechanisms: Current State

Page 8 of 8

of Understanding. ACS Catal. 2018, 8, 1161−1172. (c) Jia, F.; Luo, J.; Zhang, B. Using a traceless directing group for the silver-mediated synthesis of 3-trifluoromethylpyrazoles: a computational study on the mechanism and origins of regioselectivity. Org. Chem. Front. 2018, 5, 3374−3381. (25) DFT calculations suggest that a stepwise pathway is more likely for the formation of intermediate B during the cycloaddition process. Please see Figure S6 in the Supporting Information for details. (26) For a previous study on the reaction of 1,2,4-triazines and 2,5norbornadiene to afford pyridines, see: (a) Stanforth, S. P.; Tarbit, B.; Watson, M. D. Synthesis of pyridine and 2,2’-bipyridine derivatives from the aza Diels–Alder reaction of substituted 1,2,4-triazines. Tetrahedron 2004, 60, 8893−8897. For selected examples on the reactions of 1,2,4-triazines and arynes to afford isoquinolines: (b) Dhar, R.; Hühnermann, W.; Kampchen, T.; Overheu, W.; Seitz, G. Oxepin und 2,7-Dimethyloxepin als Dienophile bei Diels-Alder Cycloadditionen mit inversem Elektronenbedarf. Chem. Ber. 1983, 116, 97−102. (c) Gonsalves, A M. R.; Melo, T. M. V. D. P.; Gilchrist, T. L. Synthesis of isoquinolines by cycloaddition of arynes to 1,2,4-triazines. Tetrahedron 1992, 48, 6821−6826. (d) Kopchuk, D. S.; Nikonov, I. L.; Zyryanov, G. V.; Kovalev, I. S.; Rusinovl, V. L.; Chupakhin, O. N. Preparation of 3-Cyano-1-(2-Pyridyl)Isoquinolines by Using Aryne Intermediates. Chem. Heterocycl. Compd. 2014, 50, 907−910. (e) Kopchuk, D. S.; Nikonov, I. L.; Khasanov, A. F.; Giri, K.; Santra, S.; Kovalev, I. S.; Nosova, E. V.; Gundala, S.; Venkatapuram, P.; Zyryanov, G. V.; Majee, A.; Chupakhin, O. N. Studies on the interactions of 5-R-3-(2-pyridyl)-1,2,4-triazines with arynes: inverse demand aza-Diels–Alder reaction versus aryne-mediated domino process. Org. Biomol. Chem. 2018, 16, 5119−5135. (27) (a) Boger, D. L.; Panek, J. S. Diels-Alder Reaction of Heterocyclic Azadienes. Thermal Cycloaddition of 1,2,4-Triazine with Enamines: Simple Preparation of Substituted Pyridines. J. Org. Chem. 1981, 46, 2179−2182. (b) Glinkerman, C. M.; Boger, D. L. Catalysis of Heterocyclic Azadiene Cycloaddition Reactions by Solvent Hydrogen Bonding: Concise Total Synthesis of Methoxatin. J. Am. Chem. Soc. 2016, 138, 12408–12413. (c) Suleymanova, A. F.; Kozhevnikov, D. N.; Prokhorov, A. M. The use of the 1,2,4-triazine method of pyridine ligand synthesis for the preparation of a luminescent Pt(II) labeling agent. Tetrahedron Lett. 2012, 53, 5293– 5296. (d) Kovalev, I. S.; Kopchuk, D. S.; Khasanov, A. F.; Zyryanov, G. V.; Rusinov, V. L.; Chupakhin, O. N. The synthesis of polyarenemodified 5-phenyl-2,2'-bipyridines via the SNH methodology and azaDiels–Alder reaction. Mendeleev Commun. 2014, 24, 117–118. (e) Krinochkin, A. P.; Kopchuk, D. S.; Kim, G. A.; Ganebnykh, I. N.; Kovalev, I. S.; Santra, S.; Majee, A.; Zyryanov, G. V.; Rusinov, V. L.; Chupakhin, O. N. Synthesis and photophysical studies of new organicsoluble lanthanide complexes of 4-(4-alkoxyphenyl)-2,2´-bipyridine6-carboxylic acids. J. Mol. Struct. 2019, 1176, 583–590. (f) Starnovskaya, E. S.; Kopchuk, D. S.; Khasanov, A. F.; Tanya, O. S.; Santra, S.; Giric, K.; Rahman, M.; Kovalev, I. S.; Zyryanov, G. V.; Majeed, A.; Charushin, V. N. Synthesis and photophysics of new unsymmetrically substituted 5,5´-diaryl 2,2´-bypiridine-based “pushpull” fluorophores. Dyes Pigm. 2019, 162, 324–330. (28) (a) Kamber, D. N.; Liang, Y.; Blizzard, R. J.; Liu, F.; Mehl, R. A.; Houk, K. N.; Prescher, J. A. 1,2,4-Triazines are Versatile Bioorthogonal Reagents. J. Am. Chem. Soc. 2015, 137, 8388–8391. (b) Horner, K. A.; Valette, N. M.; Webb, M. E. Strain-Promoted Reaction of 1,2,4-Triazines with Bicyclononynes. Chem. Eur. J. 2015, 21, 14376–14381. (c) Siegl, S. J.; Dzijak, R.; V´azquez, A.; Pohl, R.; Vrabel, M. The Discovery of Pyridinium 1,2,4-Triazines with Enhanced Performance in Bioconjugation Reactions. Chem. Sci. 2017, 8, 3593–3598. (d) Peewasan, K.; Wagenknecht, H.-A. 1,2,4-TriazineModified 2′-Deoxyuridine Triphosphate for Efficient Bioorthogonal Fluorescent Labeling of DNA. ChemBioChem 2017, 18, 1473–1476.

ACS Paragon Plus Environment