Mn(III)-Promoted Synergistic Radical N-Heteroannulation

A plausible mechanism for the formation of fullerotetrahydroquinolines is proposed, and the electrochemical properties of the obtained fullerene adduc...
0 downloads 0 Views 391KB Size
Subscriber access provided by University of Rochester | River Campus & Miner Libraries

Article

Cu(II)/Mn(III)-Promoted Synergistic Radical N-Heteroannulation Reaction: Synthesis of [60]Fullerene-Fused Tetrahydroquinoline Derivatives Qingfeng Liu, Tongxin Liu, Nana Ma, Chenhao Tu, Ruoya Wang, and Guisheng Zhang J. Org. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.joc.9b00929 • Publication Date (Web): 16 May 2019 Downloaded from http://pubs.acs.org on May 16, 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 37 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

The Journal of Organic Chemistry

Cu(II)/Mn(III)-Promoted

Synergistic

Radical

N-Heteroannulation Reaction: Synthesis of [60]Fullerene-Fused Tetrahydroquinoline Derivatives

Qingfeng Liu, Tong-Xin Liu,* Nana Ma, Chenhao Tu, Ruoya Wang, and Guisheng Zhang*

Collaborative Innovation Center of Henan Province for Green Manufacturing of Fine Chemicals, Key Laboratory of Green Chemical Media and Reactions, Ministry of Education, Henan Key Laboratory of Organic Functional Molecule and Drug Innovation, School of Chemistry and Chemical Engineering, Henan Normal University, Xinxiang, Henan 453007, China. E-mail: [email protected] and [email protected]

Abstract _____________________________________________________________________ MeO2C

CO2Me + R1

, ODCB/CH3CN, N2

NH O 2S

Cu(II)/Mn(III)-mediated

CO2Me

CO2Me CuCl2, Mn(OAc)3, Cs2CO3

O 2S

R2

synergistic

R1 N

radical

N-heteroannulation

R2

reaction

of

[60]fullerene with N-sulfonylated o-amino-arylmalonates has been developed for the direct and efficient construction of [60]fullerene-fused tetrahydroquinoline derivatives. A plausible mechanism for the formation of fullerotetrahydroquinolines is proposed, and the electrochemical properties of the obtained fullerene adducts are investigated.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 2 of 37

_____________________________________________________________________ Introduction Fullerenes and their derivatives have attracted the interest of scientists from many disciplines owing to their unique structures and properties.1,2 Chemical modification of fullerenes has proven to be a promising approach for the creation of new nanocarbon-based materials by introducing different functional and structural units on the cage surface or carving carbon cage skeleton itself. Over the past few decades, many efforts have been devoted to this direction and, as a result, different synthetic methods and strategies have been established, thus leading to the formation of diverse fullerene derivatives.3,4 Free radical reactions are among the most powerful tools for the

chemical

modification

of

fullerenes.3e,5

Especially

in

recent

years,

transition-metal-mediated radical reactions, including Mn(III),6 Cu(I)/Cu(II),7 Fe(II)/Fe(III),8 Co(II),9 Ni(II),10 Ag(I),11 W(VI),12 Ir(III),13 and Pb(IV)14, have received growing attention due to their remarkable advantages of high efficiency, high selectivity and high compatibility with substrates and functional groups compared with traditional light- and thermal-induced radical processes. Heterocyclic skeletons are frequently encountered in optoelectronic materials, natural products and bioactive molecules. The incorporation of heteroatoms such as nitrogen or oxygen onto the C60 core to form fullerene-fused heterocycle derivatives results in better electron-accepting ability compared with parent C60, which is beneficial for the preparation of a wide variety of fullerene-based organic photovoltaic devices.15 Therefore, the development of direct and efficient methods for

ACS Paragon Plus Environment

Page 3 of 37 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

The Journal of Organic Chemistry

the preparation of diverse fullerene-fused heterocyclic structures becomes an important subject in fullerene chemistry and has received wide attention.1,3,4 However, among many known fullerene-fused heterocyclic derivatives, reports on the formation of C60-fused tetrahydroquinoline are quite limited, and there are only two currently published approaches allowing for its synthesis.16 The first has been reported by Martín and coworkers, in which C60-fused tetrahydroquinoline was first synthesized via thermal hetero-Diels − Alder reaction of C60 with N-methyl substituted o-aminobenzyl alcohols under metal-free conditions.16a Another known pathway is based on that Wang’s group described Cu2O-assisted hetero-Diels−Alder reaction of C60 with N-(o-chloromethyl)aryl sulfonamides.16b Thus, the exploration of new synthetic routes access to structurally diverse C60-fused tetrahydroquinoline is desirable. Recently, we successfully carried out a range of transition-metal-promoted radical heteroannulation reactions to achieve the synthesis of a series of diverse fullerene-fused heterocycle derivatives.7f,j,8e,11b,17 Among them, the radical reaction of C60 with N-sulfonylated o-amino-aromatic methyl ketones and O-alkyl oximes selectively

furnished

novel

[60]fullerene-fused

tetrahydroazepinones

and

-azepinonimines in the presence of Cu(OAc)2 (Scheme 1).7f Inspired by this work, we were interested to investigate if the readily available, inexpensive and low-toxic Cu(II)-mediated radical protocol could be further developed for the preparation of rare

[60]fullerene-fused

tetrahydroquinolines

by

using

N-sulfonylated

o-amino-2-arylmalonates as reaction partners for C60, as shown in Scheme 1. Scheme

1.

Our

Provious

Work

of

the

Preparation

ACS Paragon Plus Environment

[60]Fullerene-Fused

The Journal of Organic Chemistry 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

Page 4 of 37

Tetrahydroazepinones and -azepinonimines and Current Design for the Synthesis of [60]Fullerene-Fused Tetrahydroquinolines. Our previous work: X +

X Cu(OAc)2, Cs2CO3

R1

, ODCB/CH3CN

NH SO2R2 X = O, N-OMe(Bn)

N SO2R2

R1

New design: CO2R +

RO2C CO2R

CO2R Cu(OAc) , additives 2

R1 NH

SO2R2

R1 N SO2R2

Results and Discussion We commenced the study with easily available N-(4-methylbenzsulfonyl) substituted o-amino-2-phenylmalonate 1a as a model substrate to reaction with C60 under different conditions (Table 1). Initially, the reaction was treated under similar reaction conditions to those of the reported preparation of [60]fullerene-fused tetrahydroazepinones and –azepinonimines.7f However, only a trace amount of the desired product 2a was obtained, even when raising the reaction temperature to 130 °C (Table 1, entries 1 and 2). Further examination of other copper salts, including Cu(OTf)2, CuSO4, CuBr2 and CuCl2, revealed that CuCl2 was a better choice for the transformation, but the product 2a was only obtained in 7% yield (Table 1, entries 3−6). Replacing Cs2CO3 with other bases such as K2CO3, CsOAc, HCOOCs, and 4-dimethylaminopyridine (DMAP), all proved to be ineffective (Table 1, entries 7−10). The above optimization experiments indicated that copper salts were less

ACS Paragon Plus Environment

Page 5 of 37 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

The Journal of Organic Chemistry

effective in the activation of 2-arylmalonates for generating the corresponding C-centered radical compared with aromatic methyl ketones,7a,f which suggests that it is not possible to accomplish this transformation by using the single CuCl2 as an oxidant. Considering that manganese acetate (Mn(OAc)3) has a good capability for abstracting a hydrogen atom from active methylene compounds to form carbon radicals,6,18 we envisioned that the annulation reaction may be achieved through a Cu/Mn-mediated synergistic effect. To our delight, the yield of 2a significantly increased to 36% with specific selectivity when 2 equiv of Mn(OAc)3 was added to the reaction system (Table1, entry 11). We also observed that the reaction did not proceed only in the presence of Mn(OAc)3 without CuCl2, confirming that they are indeed responsible together for the cycloaddition reaction (Table1, entry 12). Prolonging the reaction time or raising the temperature did not give 2a in higher yields (Table 1, entries 13 and 14). Carrying out the reaction in air resulted in an obvious decrease in yield (Table 1, entry 15). Therefore, the molar ratio of 1:3:2:2:1 for the reagents C60, 1a, CuCl2, Mn(OAc)3•2H2O, and Cs2CO3 and the reaction temperature of 130 oC in ODCB/CH3CN under a nitrogen atmosphere were chosen as the optimized reaction conditions. Table 1. Optimization of the Reaction Conditionsa MeO2C

CO2Me CO2Me

+

NHTs

CO2Me

Cu(II), Mn(OAc)3, base 130 oC, ODCB/CH3CN, N2

N Ts

1a

entry

Cu(II)

Mn(OAc)3•2H2O

2a

base

ACS Paragon Plus Environment

molar ratiob

yield (%)c

The Journal of Organic Chemistry 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

a

Page 6 of 37

1

Cu(OAc)2



Cs2CO3

1:3:2:0:1

traced

2

Cu(OAc)2



Cs2CO3

1:3:2:0:1

trace

3

Cu(OTf)2



Cs2CO3

1:3:2:0:1

trace

4

CuSO4



Cs2CO3

1:3:2:0:1

no reaction

5

CuBr2



Cs2CO3

1:3:2:0:1

no reaction

6

CuCl2



Cs2CO3

1:3:2:0:1

7 (37)

7

CuCl2



K2CO3

1:3:2:0:1

trace

8

CuCl2



CsOAc

1:3:2:0:1

trace

9

CuCl2



HCOOCs

1:3:2:0:1

no reaction

10

CuCl2



DMAP

1:3:2:0:1

trace

11

CuCl2

Mn(OAc)3•2H2O

Cs2CO3

1:3:2:2:1

36 (97)

12



Mn(OAc)3•2H2O

Cs2CO3

1:3:0:2:1

trace

13

CuCl2

Mn(OAc)3•2H2O

Cs2CO3

1:3:2:2:1

25 (71) e

14

CuCl2

Mn(OAc)3•2H2O

Cs2CO3

1:3:2:2:1

27 (77)f

15

CuCl2

Mn(OAc)3•2H2O

Cs2CO3

1:3:2:2:1

10 (83)g

All reactions were carried out with a designated molar ratio in co-solvent of

anhydrous ODCB (7 mL) and CH3CN (1 mL) for 3 h under a nitrogen atmosphere unless specified otherwise. b Molar ratio refers to C60/1a/Cu(II)/Mn(OAc)3•2H2O/base. c

Isolated yields; values in parentheses were based on consumed C60.

80 °C. e Reaction for 4 h. f Reaction at 140 °C. g Reaction in air.

ACS Paragon Plus Environment

d

Reaction at

Page 7 of 37 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

The Journal of Organic Chemistry

With the optimized conditions in hand, the substrate scope was investigated by employing a wide array of N-sulfonyl-2-phenylmalonate esters for the synthesis of [60]fullerene-fused

tetrahydroquinolines,

as

shown

in

Table

2.

Different

N-arylsulfonylated substrates 1a−f underwent efficient transformations with C60 to afford the corresponding products 2a−f in 27%−36% yields. Electron-donating (2a and 2c) and -withdrawing (2d) functional groups present in the aryl substituent of the sulfonamides were well tolerated in the developed synergistic annulation system. The reaction of 2-naphthalene sulfonylated ester 1e with C60 also furnished the tetrahydroquinoline derivative 2e in a moderate yield (28%). Furthermore, N-heteroaryl sulfonylated 2-phenylmalonate ester 1f was also a suitable substrate for the reaction and a higher yield of the desired product (40%) was obtained. In addition, substrates 1g−j bearing an electron-donating or -withdrawing group at different positions of the arylmalonate moiety also efficiently participated in the reaction to afford the desired products 2g−j in moderate to good yields. When the methyl ester was changed to an ethyl ester, a comparable yield (31%) was obtained for the transformation of 1k with C60 under the standard conditions. Subsequently, we also explored

the

possibility

for

the

synthesis

of

[60]fullerene-fused

tetrahydrobenzoazepines using N-sulfonylated o-amino-2-benzylmalonates as reaction substrates.

To

our

disappointment,

no

expected

[60]fullerene-fused

tetrahydrobenzoazepine derivative or other fullerene adducts were obtained when dimethyl 2-(2-((4-methylphenyl)sulfonamido)benzyl)malonate was subjected to the standard reaction conditions, and the exact reason is unclear at the present stage.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 8 of 37

Table 2. Cu(II)/Mn(III)-Mediated Synergistic Synthesis of [60]fullerene-Fused Tetrahydroquinolinesa R 3O 2C

CO2R3 CO2R

+ R1

MeO2C

R1 N O 2S

R2

MeO2C

CO2Me

CO2R3

CuCl2, Mn(OAc)3, Cs2CO3 130 oC, ODCB/CH3CN, N2

NH O 2S

3

CO2Me

R2

MeO2C

CO2Me

N

N

N

O S O

O S O

O S O

OCH3

2a, 36% (97%) MeO2C

2b, 36% (96%) MeO2C

CO2Me

CO2Me

2c, 34% (92%) MeO2C

CO2Me

N

N

N

O S O

O S O

O S O S

NO2

2d, 27% (80%) MeO2C

2e, 28% (93%) MeO2C

CO2Me

CO2Me

2f, 40% (92%) MeO2C

CO2Me

N

N

N

O S O

O S O

O S O

2g, 37% (82%) MeO2C

2h, 39% (89%) EtO2C

CO2Me

N O S O

2j, 29% (82%)

CF3

CO2Et

N O S O

2k, 31% (97%)

ACS Paragon Plus Environment

2i, 24% (84%)

OMe

Page 9 of 37 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

The Journal of Organic Chemistry

a

All

reactions

were

carried

out

with

a

molar

ratio

of

C60/1a/CuCl2/Mn(OAc)3•2H2O/Cs2CO3 = 1:3:2:2:1 in co-solvent of anhydrous ODCB (7 mL) and CH3CN (1 mL) for 3 h under a nitrogen atmosphere.

b

Isolated yields;

values in parentheses were based on consumed C60. To investigate the synergistic annulation mechanism, preliminary mechanistic experiments were conducted (Scheme 2). When the reaction of [60]fullerene with 1a was treated with CuCl2/Mn(OAc)3 in the absence of Cs2CO3, only trace amount of the desired product 2a was detected, indicating that Cs2CO3 is crucial for the success of the transformation (Scheme 2i). However, the reaction did not proceeded only in the presence of Cs2CO3, thus the possibility that nitrogen or carbon anion formed in situ directly

attacks

C60

and

subsequently

further

undergoes

other

transformation processes access to the final product 2a could be ruled out (Scheme 2ii). Next, the radical-trapping experiments were also performed (Scheme 2iii). The addition of free radical scavenger 2,2,6,6-tetramethylpiperidine-1-oxy (TEMPO) or N-tert-butyl-α-phenylnitrone (PBN) could obviously retard or completely suppress the heteroannulation reaction depending on the amount of TEMPO or PBN used, suggesting that the current transformation might involve a free radical pathway, although many efforts to detect or isolate the possible radical intermediates failed. Scheme 2. Preliminary Mechanistic Experiments

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

MeO2C

CO2Me

i)

CO2Me

+

NHTs

Page 10 of 37

CO2Me

CuCl2, Mn(OAc)3 130 oC, ODCB/CH3CN, N2

N Ts

(trace)

1a

2a

CO2Me

ii)

CO2Me

+

NHTs

Cs2CO3 130 oC, ODCB/CH3CN, N2

no reaction

1a MeO2C

CO2Me

iii)

CO2Me

+

NHTs

CO2Me

CuCl2, Mn(OAc)3, Cs2CO3 130 oC, ODCB/CH3CN, N2

N Ts

1a

2a radical scavenges TEMPO (2 equiv) (3 equiv) PBN (2 equiv) (3 equiv)

yield 7% trace 19% trace

Based on the above experimental results, a possible radical mechanism involving different reaction paths to rationalize the formation of [60]fullerene-fused tetrahydroquinolines 2 is shown in Scheme 3. In path a, the substrate 1 undergoes one-electron oxidation of the nitrogen to produce an N-centered radical I in the presence of Cu(II),7c,11a,19 which is captured by C60 to form fullerene radical II. Next, the radical II further undergoes Mn(OAc)3-mediated relay hydrogen absorption to generate the C-centered biradical species III,9b,20 and subsequent intramolecular biradical C−C coupling affords the final product 2. In an alternative path b, the C-centered radical IV is formed initially via the reaction of 1 with Mn(OAc)3.6,18 Addition of radical IV to C60 then gives fullerenyl radical V, which is oxidized by Cu(II) with the assistance of base to generate C- and N-centered biradical species VI.7f,j Finally, a similar intramolecular biradical C−N coupling cyclization occurs to afford the product 2. In another alternative route, via path c, the biradical species VII

ACS Paragon Plus Environment

Page 11 of 37

is generated via Cu(II)-Mn(III) co-mediated dehydrogenation. Rapid tautomerization of VII further converts into aza-o-quinone methide intermediate VIII,16,21 which undergoes Diels−Alder reaction with C60 to give the desired product 2.22 Scheme 3. Proposed Reaction Mechanism CO2R3 CO2R

R1 N O2 S

ba se (II ), Cu

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

The Journal of Organic Chemistry

C60 N O2 S

R2

R

CO2R3 NH O2 S 2 R

1

Cu(II), Mn(OAc)3, base

R2

R1

Path c

N O2 S

CO2R3

R1 N O2 S

R2

CO2R3 R

CO2R3

1

NH O2 S 2 R

IV

R 3O 2 C

R 3O 2 C

C60 HN O2 S

V

R1 N O2 S

R2

R2

2 R 3O 2 C

CO2R3 R1

CO2R3

C60

VIII

VII

3

R2

III

CO2R3

CO2R3

R1 N O2 S

CO2R3

Path b M n( O Ac )

Mn(OAc)3

II

CO2R3 1

R1

I

Path a

R3O2C CO2R3

R3O2C CO2R3 3

Cu(II), base

CO2R3 R1

N O2 S

R2

VI

R2

To better understand the mechanism of this transformation, density functional theory (DFT) calculations for the reaction of C60 with 1a were performed at the (U)B3LYP/6-31G(d) level, and the energy profiles are depicted in Scheme 4. Although the C-centered radical species IV is more stable than N-centered radical species I by 3.6 kcal mol−1, the formation of intermediate II requires 26.2 kcal mol−1, 12.9 kcal mol−1 higher than that required for the formation of V. Furthermore, the intermediate II is inclined to return the C-centered radical species IV only overcoming 6.5 kcal mol−1. Thus, the reaction path a via N-centered radical species I should be more favorable than path b. Interestingly, theoretical calculation indicates that path c is also possible under applied reaction conditions. The generated unstable

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 12 of 37

biradical species VII rapidly converts into a dienes species VIII, which subsequently undergoes D-A cycloaddition with C60, climbing energy barrier of 17.5 kcal mol−1 access to 2. It should be pointed out, from the perspective of the formed intermediate, this path can explain why the aforementioned reaction of C60 with dimethyl 2-(2-((4-methylphenyl)sulfonamido)benzyl)malonate

fails

in

producing

the

corresponding tetrahydrobenzoazepine derivative. Therefore, the reaction path c is not ruled out. Scheme 4. DFT Energy Profiles for the Different Reaction Pathway G298K (kcal/mol) CO2Me

19.7

Ts

15.5

II

TS2

I

3.6

CO2Me

0.0

CO2Me

C60

MeO2C CO2Me

HN

CO2Me N Ts

N radical C radical biradical

MeO2C CO2Me

26.2 TS1

5.3

HN Ts

MeO2C CO2Me

TS1 MeO2C CO2Me

2.2 C60

-0.1

N Ts

NH

N Ts

V

Ts

IV

TS2

CO2Me CO2Me

MeO2C CO2Me

20.5

17.5

N

TS3

Ts

N

VII

Ts

CO2Me

TS3

3.4

CO2Me

C60

MeO2C

0.0

CO2Me

N

-10.8

Ts

N

VIII

Ts

2

In order to clearly elucidate the reaction mechanism, we further carried out the following

control

experiments

(Scheme

5).

It

was

found

that

when

4-methyl-N-phenylbenzenesulfonamide 3a was employed under the standard reaction conditions or in the presence of CuCl2/Cs2CO3, no reaction occurred, even if 4-methyl-N-(o-tolyl)benzenesulfonamide 3b with similar electronic and steric effects

ACS Paragon Plus Environment

Page 13 of 37 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

The Journal of Organic Chemistry

to substrate 1a was used as a starting material (Scheme 5i and 5ii). Thus, the possibility that the transformation via path a, involving the direct addition of an N-centered radical species to C60, should be excluded. As a control, the reactions of dimethyl 2-phenylmalonate 4a and dimethyl 2-(2-nitrophenyl)malonate 4b with C60 were also performed (Scheme 5iii and 5iv). Likewise, no fullerene adduct could be detected, and the results also fully supports theoretical research, indicating that the N-heteroannulation reaction is impossible to occur via path b. Therefore, according to the aforementioned experimental results and DFT study, the current reaction is most likely to happen via path c in proposed three reaction paths. Scheme 5. Control Experiments standard conditions

i)

no reaction

+ NHTs

3a

ii)

CuCl2, Cs2CO3 130 oC, ODCB/CH3CN, N2

standard conditions

+

no reaction

no reaction

NHTs

3b CO2Me

iii)

standard conditions

no reaction

CO2Me

+

Mn(OAc)3, Cs2CO3

4a

130 oC, ODCB/CH3CN, N2

no reaction

CO2Me

iv)

+

CO2Me

standard conditions

no reaction

NO2

4b

The electrochemical properties of the products along with [6,6]-phenyl-C61-butyric acid methyl ester (PCBM) and C60 have been investigated by cyclic voltammetry (CV), and their half-wave reduction potentials are summarized in Table 3. The results

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 14 of 37

revealed that the obtained products exhibited irreversible redox processes, which is different from electrochemical behavior observed in reported N-methyl substituted similar structures,16a indicating that the presence of sulfonyl groups with strong electron-withdrawing has significant effect on the redox properties of products and easily leads to cleavage of the C−N bonds of 2a−k upon acceptance of electrons.23 However, the first redox process of the products was reversible at a scanning rate of 100 mV s-1. Furthermore, the first half-wave reduction potentials of products show obvious cathodic shifts for the redox processes compared to the parent fullerene, because one of the C═C double bonds in the fullerene surface is saturated. When compared to PCBM, all the new compounds present a first reduction potential that is anodically shifted, which indicates a small but consistent influence of the electronegativity of the N atom (and may be sulfonyl groups) on the redox properties. Table 3. Half-Wave Reduction Potentials (V) of Products 2, PCBM and C60a Compound

E1

E2

E3

2a

−1.160





2b

−1.141





2c

−1.153





2d

−1.088





2e

−1.129





2f

−1.127





2g

−1.138





2h

−1.133





ACS Paragon Plus Environment

Page 15 of 37 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

The Journal of Organic Chemistry

2i

−1.098





2j

−1.118





2k

−1.140





PCBMb

–1.175

−1.576

–2.103

C60b

−1.075

−1.463

−1.932

a

Versus ferrocene/ferrocenium; experimental conditions: 1 mM of compound 2 and

0.1 M of (n-Bu)4NClO4 in anhydrous o-dichlorobenzene; reference electrode: SCE; working electrode: Pt; auxiliary electrode: Pt wire; scanning rate: 100 mV s-1.

b

Scanning rate: 20 mV s-1. Conclusion In summary, we have successfully disclosed a Cu(II)/Mn(III)-mediated synergistic radical

N-heteroannulation

o-amino-arylmalonates

for

reaction the

direct

of

[60]fullerene and

efficient

with

N-sulfonylated

preparation

of

novel

[60]fullerene-fused tetrahydroquinoline derivatives. The reaction mechanism has been systematically studied based on control experiments and theoretical calculation, and electrochemical properties of the new-obtained fullerene adducts has been investigated.

Experimental Section General Information. Mn(OAc)3•2H2O, CuCl2, CuBr2, and Cs2CO3 were purchased from Sigma-Alderich. o-Dichlorobenzene (ODCB) and CH3CN were treated with CaH2. 1H NMR (400 MHz, referenced to residual CDCl3 at 7.26 ppm) and 13C NMR (100

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 16 of 37

and 150 MHz, referenced to residual CDCl3 at 77.16 ppm) were registered on Bruker 400 and 600 M spectrometers. HRMS were measured on Bruker Ultraflextreme MALDI-TOF/TOF

in

a

positive

mode

using

E-2-[3-(4-tert-butylphenyl)-2-methyl-2-propenylidene]malononitrile (DCTB) as a matrix. UV-vis spectra were recorded on Shimadzu UV-1700 spectrometer with CHCl3 as the solvent. All of the DFT calculations were performed with the Gaussian 09

program.24

The

geometries

optimizations

were

performed

at

the

(U)B3LYP25/6-31G(d) level. The vibrational frequencies were computed at the same level to check whether each optimized structure is an energy minimum or a transition state and to evaluate its thermal corrections at 298 K. The Gibbs free energies (ΔG) are used to discuss the reaction. The starting materials 1a−k were prepared according to the literature reported procedures and their identities were confirmed by comparison of their spectral data with those reported in the literature.26 General Procedure for the Synthesis of Products 2: A dry 25-mL Schlenk tube equipped with a magnetic stirrer was charged with C60 (36.0 mg, 0.05 mmol), 1a (1b−k, 0.15 mmol), CuCl2 (0.10 mmol, 13.4 mg), Mn(OAc)3•2H2O (0.10 mmol, 26.8 mg), and Cs2CO3 (0.05 mmol, 16.3 mg). After dissolving the solids in a mixture of anhydrous ODCB (7 mL) and CH3CN (1 mL) by sonication, the sealed Schlenk tube was stirred in an oil bath pre-set at 130 oC under a nitrogen atmosphere. The reaction mixture was filtered through column chromatography on silica gel to remove any insoluble material. After the solvent had been evaporated under vacuum, the residue was separated through column chromatography on silica gel with CS2 as the eluent to

ACS Paragon Plus Environment

Page 17 of 37 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

The Journal of Organic Chemistry

recover unreacted C60, and then the eluent was switched to CS2/DCM to give products 2a−k. 1-Tosyl-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquino[60]fullerene

2a.

Yield 19.7 mg, 36%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.21−8.19 (m, 1H), 8.02 (d, J = 8.4 Hz, 2H), 7.69−7.66 (m, 1H), 7.63−7.59 (m, 2H), 7.29 (d, J = 8.0 Hz, 2H), 4.12 (br, 3H), 3.88 (s, 3H), 2.41(s, 3H). 13C{1H} NMR (100 MHz, CDCl3, all 1C unless indicated) δ 169.1, 167.7, 152.4, 151.2, 148.3, 147.9, 147.8, 146.8, 146.7 (2C), 146.6, 146.3, 146.28, 146.26, 146.2, 146.1, 146.0, 145.9, 145.7, 145.6, 145.58, 145.56 (2C), 145.5, 145.45, 145.4, 145.35, 145.1, 144.8, 144.7 (2C), 144.6, 144.5, 144.41, 143.40, 143.3, 143.1, 142.9, 142.88, 142.8 (2C), 142.75, 142.7, 142.5, 142.3, 142.1, 141.84, 141.83, 141.6, 141.59, 141.45, 141.41, 141.12, 139.1, 139.06 (2C), 139.0, 138.2, 137.6, 135.2, 134.9, 134.87, 129.9 (3C), 129.5, 129.2 (3C), 129.1, 128.1, 126.1, 82.9, 76.2, 54.3, 53.8, 21.8; FT-IR ν/cm-1 (KBr) 2952 (CH3), 1738 (C=O), 1462, 1432, 1359 (S=O), 1235 (C−O−C), 1167 (S=O), 1086 (C−O−C), 1034 (C−O−C), 753, 664, 566, 527; λmax/nm (CHCl3) 258, 318, 415, 694; MALDI-FT MS m/z calcd for C78H17NO6S [M]+ 1095.0771, found 1095.0778. 1-(Phenylsulfonyl)-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquino[60]fulle rene 2b. Yield 19.4 mg, 36%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.21−8.19 (m, 1H), 8.15 (d, J = 8.0 Hz, 2H), 7.70−7.68 (m, 1H), 7.63−7.61 (m, 2H), 7.57 (d, J = 7.6 Hz, 1H), 7.52−7.48 (m, 2H), 4.13 (s, 3H), 3.88 (s, 3H); 13C{1H} NMR (100 MHz, CDCl3, all 1C unless indicated) δ 169.1, 167.7, 152.4, 151.0, 148.3, 147.9, 147.5, 146.8, 146.7 (2C), 146.6, 146.3, 146.27 (2C), 146.2, 146.1, 146.0, 145.9,

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

145.7, 145.6, 145.56 (2C), 145.55, 145.5, 145.46, 145.3 (2C), 145.1, 144.8, 144.6, 144.55, 144.4, 143.4, 143.2, 143.1, 143.07, 142.9, 142.88, 142.82, 142.80, 142.7 (2C), 142.5, 142.3, 142.1, 142.0, 141.83, 141.81, 141.6 (2C), 141.43, 141.41, 141.1, 139.1, 139.0, 138.9, 138.2, 137.6, 135.2, 135.0, 134.8, 133.7, 129.6, 129.3 (3C), 129.2, 129.1 (3C), 128.2, 126.1, 82.9, 76.2, 54.3, 53.9; FT-IR ν/cm-1 (KBr) 2949 (CH3), 1738 (C=O), 1512, 1481, 1432, 1361 (S=O), 1234 (C−O−C), 1169 (S=O), 1086 (C−O−C), 1034 (C−O−C), 753, 688, 569, 527; λmax/nm (CHCl3) 258, 318, 416, 694; MALDI-FT MS m/z calcd for C77H15NO6S [M]+ 1081.0615, found 1081.0609. 1-[(4-Methoxyphenyl)sulfonyl]-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroqui no[60]fullerene 2c. Yield 18.7 mg, 34%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.20−8.18 (m, 1H), 8.06 (d, J = 8.8 Hz, 2H), 7.75−7.73 (m, 1H), 7.62−7.60 (m, 2H), 6.94 (d, J = 9.2 Hz, 2H), 4.12 (br, 3H), 3.88 (s, 3H), 3.85(s, 3H); 13C{1H}

NMR (150 MHz, CDCl3, all 1C unless indicated) δ 169.0, 167.5, 163.5,

152.2, 151.0, 148.1, 147.7, 147.5, 146.7, 146.5 (2C), 146.4, 146.2, 146.1 (2C), 146.06, 145.9 (2C), 145.7, 145.5, 145.42 (3C), 145.4, 145.34, 145.3, 145.27, 145.2, 144.9, 144.6, 144.4, 144.37, 144.2, 143.2, 143.1, 142.9, 142.8, 142.7, 142.6 (2C), 142.57, 142.5, 142.3, 142.2, 141.9, 141.7, 141.6, 141.5, 141.4, 141.3, 141.26, 141.0, 139.0, 138.95, 138.8, 138.0, 137.4, 134.9, 134.8, 134.7, 133.1, 131.4 (3C), 129.4, 128.9, 127.8, 125.9, 114.2 (2C), 82.7, 76.1, 55.8, 54.1, 53.7; FT-IR ν/cm-1 (KBr) 2944 (CH3), 1740 (C=O), 1593, 1496, 1311 (S=O), 1263 (C−O−C), 1157 (S=O), 1089 (C−O−C), 1025 (C−O−C), 833, 584, 551, 528; λmax/nm (CHCl3) 258, 318, 417, 694; MALDI-FT MS m/z calcd for C78H17NO7S [M]+ 1111.0720, found 1111.0728.

ACS Paragon Plus Environment

Page 18 of 37

Page 19 of 37 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

The Journal of Organic Chemistry

1-[(p-Nitrophenyl)sulfonyl]-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquino [60]fullerene 2d. Yield 15.2 mg, 27%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.36 (s, 4H), 8.27 (dd, J = 7.6, 1.2 Hz, 1H), 7.69−7.59 (m, 2H), 7.49 (d, J = 7.2 Hz, 1H), 4.12 (br, 3H), 3.88 (s, 3H); 13C{1H} NMR (100 MHz, CDCl3, all 1C unless indicated) δ 169.0, 167.5, 152.0, 150.6, 150.4, 148.4, 148.0, 147.7, 147.2, 146.9, 146.7 (2C), 146.66, 146.4, 146.35 (2C), 146.3, 146.0 (2C), 145.8, 145.7, 145.65, 145.63, 145.6 (2C), 145.5, 145.49, 145.4, 145.1, 144.8, 144.62, 144.6, 144.57, 144.3, 143.5, 143.25, 143.2, 143.0, 142.99, 142.93, 142.9, 142.8, 142.7, 142.6, 142.5, 142.3, 142.0, 141.9, 141.8, 141.78, 141.6, 141.5, 141.45, 141.1, 139.3, 139.2, 138.6, 138.2, 137.8, 135.3, 135.2, 134.7, 130.3 (3C), 129.8, 129.7, 128.8, 125.7, 124.5 (3C), 83.3, 75.8, 54.4, 53.9; FT-IR ν/cm-1 (KBr) 2945 (CH3), 1738 (C=O), 1529 (N=O), 1370 (S=O), 1346 (N=O), 1105 (C−O−C), 798, 736, 683, 587, 553, 526; λmax/nm (CHCl3) 258, 318, 416, 692; MALDI-FT MS m/z calcd for C77H14N2O8S [M]+ 1126.0465, found 1126.0455. 1-(Naphthalen-2-ylsulfonyl)-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquin o[60]fullerene 2e. Yield 15.8 mg, 28%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.66 (s, 1H), 8.22−8.16 (m, 2H), 7.98 (d, J = 8.8 Hz, 1H), 7.89 (d, J = 8.4 Hz, 1H), 7.82 (d, J = 8.4 Hz, 1H), 7.66−7.56 (m, 5H), 4.14 (br, 3H), 3.89 (s, 3H); 13C{1H}

NMR (100 MHz, CDCl3/CS2, all 1C unless indicated) δ 168.9, 167.5, 152.2,

151.0, 148.2, 147.7, 147.6, 146.7, 146.6, 146.52, 146.5, 146.2, 146.17, 146.14, 146.1, 146.0, 145.9, 145.8, 145.52, 145.5, 145.45 (2C), 145.43 (2C), 145.4 (2C), 145.23, 145.2, 145.0, 144.7, 144.45, 144.43, 144.3, 143.3, 143.1, 143.0, 142.9, 142.84, 142.8,

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 20 of 37

142.7, 142.68, 142.66, 142.6, 142.4, 142.2, 142.0, 141.7, 141.67, 141.5, 141.4, 141.34, 141.3, 141.0, 139.0, 138.8 (3C), 138.7, 138.0, 137.5, 135.2, 135.0, 134.8, 134.78, 132.0, 131.2, 129.5, 129.49, 129.47, 129.4, 129.1, 128.04, 128.0, 127.95, 125.9, 123.4, 82.9, 76.1, 54.2, 53.7, 53.6; FT-IR ν/cm-1 (KBr) 2948 (CH3), 1761, 1742 (C=O), 1432, 1358 (S=O), 1224 (C−O−C), 1165 (S=O), 1070 (C−O−C), 1034 (C−O−C), 860, 745, 663, 615, 557, 527; λmax/nm (CHCl3) 258, 318, 415, 694; MALDI-FT MS m/z calcd for C81H17NO6S [M]+ 1131.0771, found 1131.0767. 1-(Thiophen-2-ylsulfony)-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquino[60 ]fullerene 2f. Yield 21.5 mg, 40%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.17 (d, J = 7.6 Hz, 1H), 7.91 (d, J = 7.2 Hz, 1H), 7.80 (dd, J = 3.6, 1.2 Hz, 1H), 7.70−7.62 (m, 3H), 7.00−6.99 (m, 1H), 4.11 (br, 3H), 3.89 (s, 3H);

13C{1H}

NMR (100 MHz, CDCl3, all 1C unless indicated) δ 169.2, 167.6, 152.2, 150.4, 148.3, 147.8, 147.1, 146.8, 146.7, 146.66, 146.6, 146.3, 146.27, 146.26, 146.2 (2C), 146.0, 145.9, 145.7, 145.6, 145.5 (4C), 145.5, 145.46 (2C), 145.3, 145.0, 144.7, 144.6, 144.5, 144.4, 143.4, 143.1, 143.0, 142.97, 142.9, 142.85, 142.8, 142.78, 142.7 (2C), 142.5, 142.3, 142.0, 141.8 (2C), 141.6, 141.5, 141.4 (2C), 141.1, 139.2, 139.0, 138.7, 137.7, 137.6, 136.1 (2C), 135.1, 135.0, 134.8, 133.7 (2C), 129.7, 128.9, 128.1, 127.1, 126.1, 82.9, 76.4, 54.3, 53.9, 53.6; FT-IR ν/cm-1 (KBr) 2947 (CH3), 1740 (C=O), 1513, 1482, 1429, 1356 (S=O), 1242 (C−O−C), 1169 (S=O), 1091 (C−O−C), 1014 (C−O−C), 927, 726, 671, 570, 526; λmax/nm (CHCl3) 258, 318, 416, 694; MALDI-FT MS m/z calcd for C75H13NO6S2 [M]+ 1087.0179, found 1087.0184. 7-Methyl-1-tosyl-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquino[60]fullere

ACS Paragon Plus Environment

Page 21 of 37 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

The Journal of Organic Chemistry

ne 2g. Yield 20.5 mg, 37%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.06 (d, J = 8.0 Hz, 1H), 8.01 (d, J = 8.4 Hz, 2H), 7.45 (s, 1H), 7.40 (d, J = 8.0 Hz, 1H), 7.29 (d, J = 8.0 Hz, 2H), 4.12 (br, 3H), 3.86 (s, 3H), 2.49 (s, 3H), 2.41 (s, 3H); 13C{1H}

NMR (100 MHz, CDCl3, all 1C unless indicated) δ 169.3, 167.8, 152.6,

151.3, 148.3, 147.9, 146.8, 146.7 (2C), 146.6, 146.3, 146.26, 146.24, 146.2, 146.1, 146.0, 145.8, 145.6, 145.57, 145.55, 145.54 (2C), 145.48, 145.42, 145.4, 145.3, 145.0, 144.8, 144.7 (2C), 144.6, 144.5, 144.4, 143.4, 143.3, 143.2 (3C), 143.0, 142.9, 142.86, 142.8 (2C), 142.74, 142.7, 142.5, 142.3, 142.1, 141.8 (2C), 141.63, 141.62, 141.4, 141.39, 141.2, 139.6 (2C), 139.1, 139.06, 139.0, 138.8, 138.2, 137.6, 135.0, 134.9, 129.8 (2C), 129.2 (2C), 128.9, 128.8, 126.5, 82.9, 76.1, 54.3, 53.8, 21.8 (2C); FT-IR ν/cm-1 (KBr) 2945 (CH3), 1737 (C=O), 1497, 1429, 1357 (S=O), 1246 (C−O−C), 1164 (S=O), 1085 (C−O−C), 1042 (C−O−C), 822, 773, 727, 678, 650, 559, 526; λmax/nm (CHCl3) 258, 318, 415, 694; MALDI-FT MS m/z calcd for C79H19NO6S

[M]+

1109.0928,

found

1109.0926.

6-Methyl-1-tosyl-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquino[60]fullerene 2h. Yield 21.6 mg, 39%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.01 (d, J = 8.4 Hz, 2H), 7.96 (s, 1H), 7.51 (d, J = 8.0 Hz, 1H), 7.38 (d, J = 7.2 Hz, 1H), 7.28 (d, J = 8.4 Hz, 2H), 4.12 (br, 3H), 3.88 (s, 3H), 2.57 (s, 3H), 2.40 (s, 3H); 13C{1H}

NMR (100 MHz, CDCl3, all 1C unless indicated) δ 169.2, 167.7, 152.4,

151.3, 148.3, 147.9, 146.8, 146.7, 146.6, 146.58, 146.3, 146.25, 146.23, 146.2, 146.1, 146.06, 145.8, 145.6 (2C), 145.57, 145.5 (2C), 145.47, 145.4, 145.38, 145.3, 145.0, 144.8, 144.6, 144.57 (2C), 144.5, 144.4, 143.4, 143.3, 143.2, 143.0, 142.9, 142.8,

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

142.78 (2C), 142.7, 142.69, 142.5, 142.3, 142.1, 141.8 (3C), 141.6, 141.58, 141.4 (3C), 141.1, 139.2, 139.1, 139.0, 138.2, 138.1 (2C), 137.5, 136.2, 135.0, 134.93, 134.9, 130.3, 129.9 (2C), 129.7, 129.1 (2C), 125.6, 82.9, 76.1, 54.3, 53.8, 22.1, 21.8; FT-IR ν/cm-1 (KBr) 2949 (CH3), 1734 (C=O), 1653, 1559, 1507, 1490, 1457, 1431, 1361 (S=O), 1242 (C−O−C), 1167 (S=O), 1085 (C−O−C), 1036 (C−O−C), 811, 706, 660, 577, 526; λmax/nm (CHCl3) 258, 318, 416, 694; MALDI-FT MS m/z calcd for C79H19NO6S [M]+ 1109.0928, found 1109.0933. 7-Methoxy-1-tosyl-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquino[60]fulle rene 2i. Yield 13.5 mg, 24%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.13 (d, J = 8.4 Hz, 1H), 8.03 (d, J = 8.4 Hz, 2H), 7.29 (d, J = 8.4 Hz, 2H), 7.14−7.10 (m, 2H), 4.11 (br, 3H), 3.87 (s, 3H), 3.86 (s, 3H), 2.41 (s, 3H); 13C{1H} NMR (150 MHz, CDCl3, all 1C unless indicated) δ 169.3, 167.9, 160.2, 152.7, 151.2, 148.3, 147.9, 146.8, 146.7 (2C), 146.6, 146.3, 146.29, 146.25, 146.2, 146.1, 145.8 (2C), 145.6, 145.57 (3C), 145.56 (3C), 145.5, 145.4, 145.37, 145.3, 145.1, 144.8, 144.7, 144.6, 144.55, 144.5, 143.4, 143.3, 143.1, 142.9, 142.88, 142.83, 142.82, 142.8, 142.7, 142.5, 142.3, 142.1, 141.9, 141.86, 141.7, 141.66, 141.5, 141.4, 141.2, 140.0, 139.2, 139.1 (2C), 138.4, 137.7, 135.1, 135.0, 130.0, 129.9 (3C), 129.1 (3C), 113.1, 112.4 (2C), 83.0, 76.3, 55.7, 54.3, 53.8, 21.8; FT-IR ν/cm-1 (KBr) 2944 (CH3), 1739 (C=O), 1610, 1500, 1427, 1315 (S=O), 1278 (C−O−C), 1249 (C−O−C), 1199 (S=O), 1131 (C−O−C), 986, 822, 648, 585, 526; λmax/nm (CHCl3) 258, 319, 416, 693; MALDI-FT MS m/z calcd for C79H19NO7S [M]+ 1125.0877, found 1125.0866. 7-(Trifluoromethyl)-1-tosyl-4,4-di(methylcarboxylate)-1,2,3,4-tetrahydroquino

ACS Paragon Plus Environment

Page 22 of 37

Page 23 of 37 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

The Journal of Organic Chemistry

[60]fullerene 2j. Yield 16.7 mg, 29%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.37 (d, J = 8.8 Hz, 1H), 7.98 (d, J = 8.4 Hz, 2H), 7.82 (br, 2H), 7.33 (d, J = 8.4 Hz, 2H), 4.07 (br, 3H), 3.88 (s, 3H), 2.47 (s, 3H);

13C{1H}

NMR (100 MHz,

CDCl3/CS2, all 1C unless indicated) δ 167.7, 167.0, 151.6, 150.5, 148.1, 147.7, 146.7, 146.6, 146.5, 146.47, 146.2, 146.12, 146.11, 146.1, 145.9, 145.7, 145.6, 145.5, 145.47, 145.4, 145.38, 145.3 (3C), 145.2, 144.9, 144.89, 144.86, 144.7, 144.4, 144.3, 144.0, 143.3, 143.1 (4C), 143.0, 142.84, 142.8, 142.7, 142.66 (2C), 142.6, 142.5, 142.3, 142.2, 141.8, 141.75, 141.53, 141.50, 141.33, 141.3, 141.2, 141.0, 139.6, 139.1, 138.9, 138.7, 138.6, 138.3, 137.6, 134.7, 134.5, 131.4 (JC-F = 32.9 Hz), 129.8 (3C), 129.0 (2C), 124.3 (JC-F = 3.6 Hz), 123.4 (JC-F = 271.4 Hz), 122.5 (JC-F = 3.7 Hz), 82.5, 75.6, 54.1, 53.5, 21.7; FT-IR ν/cm-1 (KBr) 2944 (CH3), 1742 (C=O), 1422, 1369 (S=O), 1328, 1223 (C−O−C), 1170 (S=O), 1129 (C−O−C), 1081 (C−O−C), 834, 649, 592, 559, 526; λmax/nm (CHCl3) 258, 319, 418, 692; MALDI-FT MS m/z calcd for C79H16F3NO6S [M]+ 1163.0645, found 1163.0644. 1-Tosyl-4,4-di(ethylcarboxylate)-1,2,3,4-tetrahydroquino[60]fullerene

2k.

Yield 17.4 mg, 31%; brown solid; mp >300 oC; 1H NMR (400 MHz, CDCl3)  8.21 (s, 1H), 8.02 (d, J = 8.0 Hz, 2H), 7.73−7.71 (m, 1H), 7.61−7.59 (m, 2H), 7.28 (d, J = 8.4 Hz, 2H), 4.97−4.79 (m, 1H), 4.44−4.38 (m, 2H), 4.27−4.19 (m, 1H), 2.40 (s, 3H), 1.39 (br, 3H), 1.28 (t, J = 7.2 Hz, 3H);

13C{1H}

NMR (150 MHz, CDCl3, all 1C

unless indicated) δ 168.3, 167.1, 152.5, 151.1, 148.1, 147.7, 146.7, 146.52, 146.51, 146.4, 146.2, 146.1 (2C), 146.06, 146.0, 145.9, 145.7, 145.5, 145.45 (2C), 145.4 (2C), 145.35, 145.32, 145.2, 145.18, 144.9, 144.6, 144.59, 144.5, 144.4, 144.38, 143.3,

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

143.2, 142.9, 142.8, 142.7, 142.6 (2C), 142.58, 142.5, 142.3, 142.2, 141.7, 141.5, 141.46, 141.3 (2C), 141.0, 139.9, 139.0 (2C), 138.8, 138.7, 138.0, 137.4, 135.0, 134.8, 134.7, 129.7 (3C), 129.2, 129.1 (4C), 129.0, 127.7, 126.0, 82.8, 76.0, 63.0 (2C), 21.6, 14.0, 13.9; FT-IR ν/cm-1 (KBr) 2929 (CH3), 1735 (C=O), 1437, 1364 (S=O), 1232 (C−O−C), 1167 (S=O), 1086 (C−O−C), 1034 (C−O−C), 908, 814, 744, 707, 664, 566, 526; λmax/nm (CHCl3) 258, 318, 417, 694; MALDI-FT MS m/z calcd for C80H21NO6S [M]+ 1123.1084, found 1123.1091. Notes The authors declare no competing financial interest. Acknowledgements We are grateful to the NSFC (Nos. 21302044 and U1604285), Program for Science & Technology Innovation Talents in Universities of Henan Province (18HASTIT006 and 18IRTSTHN004), the 111 Project (D17007) and Outstanding Youth Science Project Funding of Henan Normal University (14YQ004) for financial support. Supporting Information Available UV–vis spectra, CVs of compounds, NMR spectra of products 2a−k, and optimized cartesian coordinates. This material is available free of charge via the Internet at http://pubs.acs.org. References (1) (a) Hirsch, A.; Brettreich, M. Fullerenes: Chemistry and Reactions, Wiley-VCH, Weinheim, 2005. (b) Kadish, K. M.; Ruoff, R. S.; Eds. Fullerenes: Chemistry, Physics, and Technology; Wiley, New York, 2000. (c) Langa, F.; Nierengarten, J.-F.,

ACS Paragon Plus Environment

Page 24 of 37

Page 25 of 37 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

The Journal of Organic Chemistry

Eds. Fullerenes: Principles and Applications; RSC Publishing: Cambridge, 2007. (2) For selected reviews, see: (a) Nakamura, E.; Isobe, H. Functionalized Fullerenes in Water. The First 10 Years of Their Chemistry, Biology, and Nanoscience. Acc. Chem. Res. 2003, 36, 807−815. (b) Anilkumar, P.; Lu, F.; Cao, L.; Luo, P. G.; Liu, J.-H.; Sahu, S.; Tackett II, K. N.; Wang, Y.; Sun, Y.-P. Fullerenes for Applications in Biology and Medicine. Curr. Med. Chem. 2011, 18, 2045−2059. (c) Nierengarten, J.-F. Fullerene Hexa-adduct Scaffolding for the Construction of Giant Molecules. Chem. Commun. 2017, 53, 11855−11868; (d) Li, C.-Z.; Yip, H.-L.; Jen, A. K.-Y. Functional Fullerenes for Organic Photovoltaics. J. Mater. Chem. 2012, 22, 4161−4177. (e) Matsuo, Y. Design Concept for High-LUMO-level Fullerene Electron-acceptors for Organic Solar Cells. Chem. Lett. 2012, 41, 754−759. (f) Cui, C.; Li, Y.; Li, Y. Fullerene Derivatives for the Applications as Acceptor and Cathode Buffer Layer Materials for Organic and Perovskite Solar Cells. Adv. Energy Mater. 2017, 7, 1601251. (g) Martín, N.; Giacalone, F.; Eds. Fullerene Polymers: Synthesis, Properties and Applications; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2009. (h) Martín, N.; Nierengarten, J.-F.; Eds. Supramolecular Chemistry of Fullerenes and Carbon Nanotubes; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2012. (i) Zieleniewska, A.; Lodermeyer, F.; Rotha, A.; Guldi, D. M. Fullerenes–How 25 Years of Charge Transfer Chemistry Have Shaped Our Understanding of (Interfacial) Interactions. Chem. Soc. Rev. 2018, 47, 702−714. (3) For selected reviews, see: (a) Thilgen, C.; Diederich, F. Structural Aspects of Fullerene Chemistry-A Journey through Fullerene Chirality. Chem. Rev. 2006, 106,

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

5049−5135. (b) Maroto, E. E.; Izquierdo, M.; Reboredo, S.; Marco-Martínez, J.; Filippone, S.; Martín, N. Chiral Fullerenes from Asymmetric Catalysis. Acc. Chem. Res. 2014, 47, 2660−2670. (c) Matsuo, Y.; Nakamura, E. Selective Multiaddition of Organocopper Reagents to Fullerenes. Chem. Rev. 2008, 108, 3016−3028. (d) Murata, M.; Murata, Y.; Komatsu, K. Surgery of Fullerenes. Chem. Commun. 2008, 6083−6094. (e) Lee, K.; Song, H.; Park, J. T. [60]Fullerene-Metal Cluster Complexes:  Novel Bonding Modes and Electronic Communication. Acc. Chem. Res. 2003, 36, 78−86. (f) Tzirakis, M. D.; Orfanopoulos, M. Radical Reactions of Fullerenes: from Synthetic Organic Chemistry to Materials Science and Biology. Chem. Rev. 2013, 113, 5262−5321. (g) Gan, L. Peroxide-Mediated Selective Cleavage of [60]Fullerene Skeleton Bonds: towards the Synthesis of Open-cage Fulleroid C55O5. Chem. Rec. 2015, 15, 189−198. (h) Itami, K. Molecular Catalysis for Fullerene Functionalization. Chem. Rec. 2011, 11, 226−235. (i) Wang, G.-W. Functionalization of [60]Fullerene via Palladium-Catalyzed C-H Bond Activation. Top. Organomet. Chem. 2016, 55, 119−136. (j) Lin, H.-S.; Matsuo, Y. Functionalization of [60]Fullerene through Fullerene Cation Intermediates. Chem. Commun. 2018, 54, 11244−11259. (4) For selected recent examples, see: (a) Li, Y.; Zhang, G.; Wang, D.; Xu, B.; Xu, D.; Lou, N.; Gan, L. Fullerene-Based Macro-Heterocycle Prepared through Selective Incorporation of Three N and Two O Atoms into C60. Angew. Chem., Int. Ed. 2016, 55, 14590−14594. (b) Xiao, Y.; Zhu, S.-E., Liu, D.-J.; Suzuki, M.; Lu, X.; Wang, G.-W. Regioselective Electrosynthesis of Rare 1,2,3,16-Functionalized [60]Fullerene

ACS Paragon Plus Environment

Page 26 of 37

Page 27 of 37 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

The Journal of Organic Chemistry

Derivatives. Angew. Chem., Int. Ed. 2014, 53, 3006−3010. (c) Nambo, M.; Segawa, Y.; Itami, K. Aziridinofullerene: A Versatile Platform for Functionalized Fullerenes. J. Am. Chem. Soc. 2011, 133, 2402–2405. (d) Cerón, M. R.; Izquierdo, M.; Aghabali, A.; Valdez, J. A.; Ghiassi, K. B.; Olmstead, M. M.; Balch, A. L.; Wudl, F.; Echegoyen, L. Tethered Bisadducts of C60 and C70 with Addends on a Common Hexagonal Face and a 12-Membered Hole in the Fullerene Cage. J. Am. Chem. Soc. 2015, 137, 7502−7508. (e) Li, S.-H.; Li, Z.-J.; Nakagawa, T.; Jeon, I.; Ju, Z.; Matsuo, Y.; Gao, X. Multifunctionalization of C70 at the Two Polar Regions with a High Regioselectivity via Oxazolination and Benzylation Reactions. Chem. Commun. 2016, 52, 5710−5713. (f) Lou, N.; Li, Y.; Gan, L. Synthesis of C70-Based Fluorophores through Sequential Functionalization to Form Isomerically Pure Multiadducts. Angew. Chem., Int. Ed. 2017, 56, 2403−2407. (g) Ueda, M.; Sakaguchi, T.; Hayama, M.; Nakagawa, T.; Matsuo, Y.; Munechika, A.; Yoshida, S.; Yasuda, H.; Ryu, I. Regio- and Stereo-Selective Intermolecular [2+2] Cycloaddition of Allenol Esters with C60 Leading to Alkylidenecyclobutane-Annulated Fullerenes. Chem. Commun. 2016, 52, 13175−13178. (h) Wu, A.-J.; Tseng, P.-Y.; Hsu, W.-H.; Chuang, S.-C. Tricyclohexylphosphine-Catalyzed Cycloaddition of Enynoates with [60]Fullerene and the Application of Cyclopentenofullerenes as n-Type Materials in Organic Photovoltaics. Org. Lett. 2016, 18, 224−227. (i) Li, S.-H.; Li, Z.-J.; Yang, W.-W.; Gao, X. Controlled Synthesis of C70 Equatorial Multiadducts with Mixed Addends from an Equatorial Diadduct: Evidence for an Electrophilic Carbanion. Org. Lett. 2018, 20, 2328−2332. (j) Li, F.; Wang, L.; Wang, J.; Peng, D.; Zhao, Y.; Li, S.; Zhou,

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 28 of 37

H.; Wu, J.; Tian, X.; Tian, Y. KOtBu-Mediated, Three-Component Coupling Reaction of Indoles, [60]Fullerene, and Haloalkanes: One-Pot, Transition-Metal-Free Synthesis of Various 1,4-(3-Indole)(organo)[60]fullerenes. Org. Lett. 2017, 19, 1192−1195. (k) Tuktarov, A. R.; Shakirova, Z. R.; Dzhemilev, U. M. One-Pot Method for the Synthesis of 2,5-Unsubstituted Pyrrolidino[3′,4′:1,9]fullerenes. Org. Lett. 2017, 19, 3863−3866. (l) Hashikawa, Y.; Murata, M.; Wakamiya, A.; Murata, Y. Palladium-Catalyzed Cyclization: Regioselectivity and Structure of Arene-Fused C60 Derivatives. J. Am. Chem. Soc. 2017, 139, 16350−16358. (m) Yang, X.-Y.; Lin, H.-S.; Jeon, I.; Matsuo, Y. Fullerene-Cation-Mediated Noble-Metal-Free Direct Introduction of Functionalized Aryl Groups onto [60]Fullerene. Org. Lett. 2018, 20, 3372−3376. (n) Hu, B.; Liu, T.-X.; Zhang, P.; Liu, Q.; Bi, J.; Shi, L.; Zhang, Z; Zhang, G. N-Heterocyclic

Carbene-Catalyzed

α,β-Unsaturated

Aldehydes

Umpolung

in

Fullerene Chemistry: Construction of [60]Fullerene-Fused Cyclopentan-1-ones and Cyclohex-2-en-1-ones. Org. Lett. 2018, 20, 4801−4805. (5) Morton, J. R.; Negri, F.; Preston, K. F. Addition of Free Radicals to C60. Acc. Chem. Res. 1998, 31, 63−69. (6) For a review, see: (a) Wang, G.-W.; Li, F.-B. Transition Metal Salt-Mediated Radical Reactions of [60]Fullerene. Curr. Org. Chem. 2012, 16, 1109−1127. (b) Li, C.; Zhang, D.; Zhang, X.; Wu, S.; Gao, X. Manganese(III) Acetate-Mediated Free Radical Reactions of [60]Fullerene with β-Dicarbonyl Compounds. Org. Biomol. Chem. 2004, 2, 3464−3469. (c) Liu, T.-X.; Li, F.-B.; Wang, G.-W. Synthesis of [60]Fullerene-Fused Tetrahydronaphthalene and Indane Derivatives via a Pathway

ACS Paragon Plus Environment

Page 29 of 37 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

The Journal of Organic Chemistry

Switched by Aluminum Chloride. Org. Lett. 2011, 13, 6130−6133, and references cited therein. (7) For selected examples, see: (a) Wang, G.-W.; Li, F.-B. Cu(II) Acetate- and Mn(III) Acetate-Mediated Radical Reactions of [60]Fullerene with Ketonic Compounds. Org. Biomol. Chem. 2005, 3, 794−797. (b) Lu, S.; Jin, T.; Kwon, E.; Bao, M.; Yamamoto, Y. Highly Efficient Cu(OAc)2-Catalyzed Dimerization of Monofunctionalized Hydrofullerenes Leading to Single-Bonded [60]Fullerene Dimers. Angew. Chem., Int. Ed. 2012, 51, 802−806. (c) Yang, H.-T.; Liang, X.-C.; Wang, Y.-H.; Yang, Y.; Sun, X.-Q.; Miao, C.-B. CuI-Catalyzed Oxidative [3+2] Reaction of Fullerene with Amidines or Amides Using Air as the Oxidant: Preparation of Fulleroimidazole or Fullerooxazole Derivatives. Org. Lett. 2013, 16, 4650−4653. (d) Xiao, Z.; Matsuo, Y.; Nakamura, E. Copper-Catalyzed Formal [4+2] Annulation between Alkyne and Fullerene Bromide. J. Am. Chem. Soc. 2010, 132, 12234−12236. (e) Si, W.; Lu, S.; Bao, M.; Asao, N.; Yamamoto, Y.; Jin, T. Cu-Catalyzed C−H Amination of Hydrofullerenes Leading to 1,4-Difunctionalized Fullerenes. Org. Lett. 2014, 16, 620−623. (f) Liu, T.-X.; Zhang, Z.; Liu, Q.; Zhang, P.; Jia, P.; Zhang, Z.; Zhang,

G.

Synthesis

of

[60]Fullerene-Fused

Tetrahydroazepinones

and

Azepinonimines via Cu(OAc)2-Promoted N-Heteroannulation Reaction. Org. Lett. 2014, 16, 1020−1023. (g) Jiang, S.-P.; Su, Y.-T.; Liu, K.-Q.; Wu, Q.-H.; Wang, G.-W. Copper(I)-Catalyzed Heteroannulation of [60]Fullerene with Ketoxime Acetates: Preparation of Novel 1-Fulleropyrrolines. Chem. Commun. 2015, 51, 6548−6551. (h) Wu, J.; Liu, C.-X.; Wang, H.-J.; Li, F.-B.; Shi, J.-L.; Liu, L.; Li, J.-X.; Liu, C.-Y.;

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 30 of 37

Huang, Y.-S. Cu(OAc)2-Mediated Reaction of [60]Fullerene with Aldehydes and Primary Amines for the Synthesis of Fulleropyrrolines. J. Org. Chem. 2016, 81, 9296−9307. (i) Yang, H.-T.; Ge, J.; Lu, X.-W.; Sun, X.-Q.; Miao, C.-B. Copper-Catalyzed Functionalizations of C60 with Amino Alcohols. J. Org. Chem. 2017, 82, 5873−5880. (j) Liu, T.-X.; Hua, S.; Ma, N.; Zhang, P.; Bi, J.; Zhang, Z.; Zhang, G. Reactivity and Synthetic Applications of α-Functionalized Oxime Acetates: Divergent

access

to

Fulleropyrrolidines

and

Mono-

and

Disubstituted

1-Fulleropyrrolines via Copper-Catalyzed Redox-Neutral N-Heteroannulation with [60]Fullerene. Adv. Synth. Catal. 2018, 360, 142−152. (8) For selected examples, see: (a) Gan, L.; Huang, S.; Zhang, X.; Zhang, A.; Cheng, B.; Cheng, H.; Li, X.; Shang, G. Fullerenes as a tert-Butylperoxy Radical Trap, Metal Catalyzed Reaction of tert-Butyl Hydroperoxide with Fullerenes, and Formation of the First Fullerene Mixed Peroxides C60(O)(OOtBu)4 and C70(OOtBu)10. J. Am. Chem. Soc. 2002, 124, 13384−13385. (b) Li, F.-B.; Liu, T.-X.; Wang, G.-W. Synthesis of Fullerooxazoles: Novel Reactions of [60]Fullerene with Nitriles Promoted by Ferric Perchlorate. J. Org. Chem. 2008, 73, 6417−6420. (c) Zhai, W.-Q.; Jiang, S.-P.; Peng, R.-F.; Jin, B.; Wang, G.-W. Facile access to Novel [60]Fullerenyl Diethers and [60]Fullerene–Sugar Conjugates via Annulation of Diol Moieties. Org. Lett. 2015, 17, 1862−1865, and references cited therein. (d) Hashiguchi, M.; Obata, N.; Maruyama, M.; Yeo, K. S.; Ueno, T.; Ikebe, T.; Takahashi, I.; Matsuo, Y. FeCl3-Mediated Synthesis of Fullerenyl Esters as Low-LUMO Acceptors for Organic Photovoltaic Devices. Org. Lett. 2012, 14, 3276−3279. (e) Liu, T.-X.; Liu, Y.; Chao, D.; Zhang, P.;

ACS Paragon Plus Environment

Page 31 of 37 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

The Journal of Organic Chemistry

Liu, Q.; Shi, L.; Zhang, Z.; Zhang, G. Iron-Mediated Internal-Oxidant Relay Cascade Reaction: Strategy to Synthesize Fullerenooxazoles and Hydroxyfullerenyl Amides. J. Org. Chem. 2014, 79, 11084−11090. (f) Zhang, X.-F.; Li, F.-B.; Wu, J.; Shi, J.-L.; Liu, Z.; Liu, L. Synthesis of Fullerene-fused Dioxanes/Dioxepanes: Ferric Perchlorate-Mediated One-Step Reaction of [60]Fullerene with Diols. J. Org. Chem. 2015, 80, 6037−6043. (9) (a) Lu, S.; Jin, T.; Bao, M.; Yamamoto, Y. Cobalt-Catalyzed Hydroalkylation of [60]Fullerene with Active Alkyl Bromides: Selective Synthesis of Monoalkylated Fullerenes. J. Am. Chem. Soc. 2011, 133, 12842−12848. (b) Lu, S.; Si, W.; Bao, M.; Yamamoto, Y.; Jin, T. Co-Catalyzed Radical Cycloaddition of [60]Fullerene with Active Dibromides: Selective Synthesis of Carbocycle-Fused Fullerene Monoadducts. Org. Lett. 2013, 15, 4030−4033. (10) Si, W.; Zhang, X.; Asao, N.; Yamamoto, Y.; Jin, T. Ni-Catalyzed Direct 1,4-Difunctionalization of [60]Fullerene with Benzyl Bromides. Chem. Commun. 2015, 51, 6392−6394. (11) (a) He, C.-L.; Liu, R.; Li, D.-D.; Zhu, S.-E.; Wang, G.-W. Synthesis and Functionalization of [60]Fullerene-Fused Imidazolines. Org. Lett. 2013, 15, 1532−1535. (b) Chao, D.; Liu, T.-X.; Ma, N.; Zhang, P.; Fu, Z.; Ma, J.; Liu, Q.; Zhang, F.; Zhang, Z.; Zhang, G. Silver(I)-Mediated Three-Component Annulation Reaction of [60]Fullerene, Sulfonylhydrazones, and Nitriles: Leading to Diverse Disubstituted [60]Fullerene-Fused Dihydropyrroles. Chem. Commun. 2016, 52, 982−985.

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 32 of 37

(12) (a) Tzirakis, M. D.; Orfanopoulos, M. Decatungstate-Mediated Radical Reactions of C60 with Substituted Toluenes and Anisoles:  A New Photochemical Functionalization Strategy for Fullerenes. Org. Lett. 2008, 10, 873−876. (b) Tzirakis, M. D.; Orfanopoulos, M. Acyl Radical Reactions in Fullerene Chemistry: Direct Acylation of [60]Fullerene through an Efficient Decatungstate-Photomediated Approach. J. Am. Chem. Soc. 2009, 131, 4063−4069. (c) Tzirakis, M. D.; Orfanopoulos, M. Photochemical Addition of Ethers to C60: Synthesis of the Simplest [60]Fullerene/Crown Ether Conjugates. Angew. Chem., Int. Ed. 2010, 49, 5891−5893. (d) Tzirakis, M. D.; Alberti, M. N.; Orfanopoulos, M. Hydroxyalkylation of [60]Fullerene: Free Radical Addition of Alcohols to C60. Chem. Commun. 2010, 46, 8228−8230. (13)

Miyake,

Y.;

Ashida,

Y.;

Nakajima,

K.;

Nishibayashi,

Y.

Visible-Light-Mediated Addition of α-Aminoalkyl Radicals to [60]Fullerene by Using Photoredox Catalysts. Chem.-Eur. J. 2014, 20, 6120−6125. (14) (a) You, X.; Li, F.-B.; Wang, G.-W. Synthesis of Ortho Acid Ester-Type 1,3-Dioxolanofullerenes: Radical Reaction of [60]Fullerene with Halocarboxylic Acids Promoted by Lead(IV) Acetate. J. Org. Chem. 2014, 79, 11155−11160. (b) Li, F.-B.; Liu, T.-X.; Huang, Y.- S.; Wang, G.-W. Synthesis of Fullerene-Fused Lactones and Fullerenyl Esters: Radical Reaction of [60]Fullerene with Carboxylic Acids Promoted by Manganese(III) Acetate and Lead(IV) Acetate. J. Org. Chem. 2009, 74, 7743−7749. (c) Chuang, S.-C.; Clemente, F. R.; Khan, S. I.; Houk, K. N.; Rubin, Y. Approaches to Open Fullerenes: a 1,2,3,4,5,6-Hexaadduct of C60. Org. Lett. 2006, 8,

ACS Paragon Plus Environment

Page 33 of 37 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

The Journal of Organic Chemistry

4525−4528. (15) (a) Prato, M. [60]Fullerene Chemistry for Materials Science Applications. J. Mater. Chem. 1997, 7, 1097−1109. (b) Da Ros, T.; Prato, M.; Carano, M.; Ceroni, P.; Paolucci, F.; Roffia, S. Enhanced Acceptor Character in Fullerene Derivatives. Synthesis and Electrochemical Properties of Fulleropyrrolidinium Salts. J. Am. Chem. Soc. 1998, 120, 11645−11648. (c) Imahori, H.; Sakata, Y. Donor-Linked Fullerenes: Photoinduced Electron Transfer and Its Potential Application. Adv. Mater. 1997, 9, 537−546. (d) Delgado, J. L.; Bouit, P.-A.; Filippone, S.; Herranz, M. Á.; Martín, N. Organic Photovoltaics: a Chemical Approach. Chem. Commun. 2010, 46, 4853−4865. (e) Chen, C.-P.; Luo, C.; Ting, C.; Chuang, S.-C. Organic Photovoltaics Incorporating Fulleroisoquinolinones as n-Type Materials. Chem. Commun. 2011, 47, 1845−1847. (16) (a) Martín, N.; Martínez-Grau, A.; Sánchez, L.; Seoane, C.; Torres, M. The First Hetero-Diels-Alder Reaction of C60 with 1-Azadienes. Synthesis of Tetrahydropyrido[2′,3′:1,2][60]fullerene Derivatives. J. Org. Chem. 1998, 63, 8074−8076. (b) Jiang, S.-P.; Lu, W.-Q.; Liu, Z.; Wang, G.-W. Synthesis of Fullerotetrahydroquinolines via [4+2] Cycloaddition Reaction of [60]Fullerene with in situ Generated Aza-o-quinone Methides. J. Org. Chem., 2018, 83, 1959−1968. (17) (a) Liu, T.-X.; Ma, J.; Chao, D.; Zhang, P.; Liu, Q.; Shi, L.; Zhang, Z.; Zhang, G. AgNO2-Mediated Cleavage of the N–N Bond of Sulfonylhydrazones and Oxygen Transfer: Access to Fulleroisoxazolines via Radical Cyclization with [60]Fullerene. Chem. Commun. 2015, 51, 12775−12778. (b) Xia, S.; Liu, T.-X.; Zhang, P.; Ma, J.; Liu, Q.; Ma, N.; Zhang, Z.; Zhang, G. Broad Synthesis of Disubstituted

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 34 of 37

Dihydrofuran-Fused [60]Fullerene Derivatives via Cu(I)/Ag(I)-Mediated Synergistic Annulation Reaction. J. Org. Chem. 2018, 83, 862−870. (18) (a) Iqbal, J.; Bhatia, B.; Nayyar, N. K. Transition Metal-Promoted Free-Radical Reactions in Organic Synthesis: The Formation of Carbon-Carbon Bonds. Chem. Rev. 1994, 94, 519−564. (b) Snider, B. B. Manganese(III)-Based Oxidative Free-Radical Cyclizations. Chem. Rev. 1996, 96, 339−364. (c) Melikyan, G. G. Manganese(III) Mediated Reactions of Unsaturated Systems. Synthesis 1993, 833−850. (d) Mondal, M.; Bora, U. Recent Advances in Manganese(III) Acetate Mediated Organic Synthesis. RSC Adv. 2013, 3, 18716−18754. (e) Pan, X.-Q.; Zou, J.-P.; Zhang, G.-L.; Zhang, W. Manganese(III)-Mediated Direct Phosphonation of Arylalkenes and Arylalkynes. Chem. Commun. 2010, 46, 1721−1723. (f) Zhou, P.; Hu, B.; Li, L.; Rao, K.; Yang, J.; Yu, F. Mn(OAc)3-Promoted Oxidative Csp3−P Bond Formation

through

Csp2−

Csp2

and

P−H

Bond

Cleavage:

Access

to

β-Ketophosphonates. J. Org. Chem. 2017, 82, 13268−13276. (19) (a) Sherman, E. S.; Chemler, S. R.; Tan, T. B.; Gerlits, O. Copper(II) Acetate Promoted Oxidative Cyclization of Arylsulfonyl-o-allylanilines. Org. Lett. 2004, 6, 1573−1575. (b) Chemler, S. R. Evolution of Copper(II) as a New Alkene Amination Promoter and Catalyst. J. Organomet. Chem. 2011, 696, 150−158. (c) Liwosz, T. W.; Chemler, S. R. Copper-Catalyzed Oxidative Amination and Allylic Amination of Alkenes. Chem.−Eur. J. 2013, 19, 12771−12777. (d) Tang, D.; Wu, P.; Liu, X.; Chen, Y.-X.; Guo, S.-B.; Chen, W.-L.; Li, J.-G.; Chen, B.-H. Synthesis of Multisubstituted Imidazoles via Copper-Catalyzed [3 + 2] Cycloadditions. J. Org. Chem. 2013, 78,

ACS Paragon Plus Environment

Page 35 of 37 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

The Journal of Organic Chemistry

2746−2750. (e) Yang, H.-T.; Liang, X.-C.; Wang, Y.-H.; Yang, Y.; Sun, X. Q.; Miao, C.-B. CuCl2-Mediated Reaction of [60]Fullerene with Amines in the Presence or Absence of Dimethyl Acetylenedicarboxylate: Preparation of Fulleropyrroline or Aziridinofullerene Derivatives. J. Org. Chem. 2013, 78, 11992−11998. (f) Yang, H.-T.; Tan, Y.-C.; Yang, Y.; Sun, X.-Q.; Miao, C.-B. Cu(OAc)2-Mediated Reaction of C60 with Ureas for the Preparation of Fulleroimidazolidinones. J. Org. Chem. 2016, 81, 1157−1163. (g) Jiang, S.-P.; Liu, Z.; Lu, W.-Q.; Wang, G.-W. Synthesis of Fullerotetrahydropyridazines

via

the

Copper-Catalyzed

Heteroannulation

of

[60]Fullerene with Hydrazides. Org. Chem. Front. 2018, 5, 1188−1193. (20) (a) Hatzimarinaki, M.; Roubelakis, M. M.; Orfanopoulos, M. Biradical Intermediate in the [2+2] Photocycloaddition of Dienes and Alkenes to [60]Fullerene. J. Am. Chem. Soc. 2005, 127, 14182−14183. (b) Liu, T.-X.; Ma, J.; Chao, D.; Zhang, P.; Ma, N.; Liu, Q.; Shi, L.; Zhang, Z; Zhang, G. Synthesis of C60-Fused Tetrahydrocarbazole/Dibenzothiophene/Benzothiophene

and

Dibenzofuran

Derivatives via Metal-Free Oxidative Dehydrogenative Carboannulation. Org. Lett. 2016, 18, 4044−4047. (21) (a) Yang, Q.-Q.; Xiao, C.; Lu, L.-Q.; An, J.; Tan, F.; Li, B.-J.; Xiao, W.-J. Synthesis of Indoles through Highly Efficient Cascade Reactions of Sulfur Ylides and N-(ortho-Chloromethyl)aryl Amides. Angew. Chem., Int. Ed. 2012, 51, 9137−9140. (b) Ito, Y.; Miyata, S.; Nakatsuka, M.; Saegusa, T. Synthesis of Nitrogen-containing Polycycles on the Basis of a New Method of o-Quinonemethide Imine Generation. J. Am. Chem. Soc. 1981, 103, 5250−5251. (c) Lee, A.; Younai, A.; Price, C. K.;

ACS Paragon Plus Environment

The Journal of Organic Chemistry 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

Page 36 of 37

Izquierdo, J.; Mishra, R. K.; Scheidt, K. A. Enantioselective Annulations for Dihydroquinolones by in situ Generation of Azolium Enolates. J. Am. Chem. Soc. 2014, 136, 10589−10592. (d) Huang, H.; Yang, Y.; Zhang, X.; Zeng, W.; Liang, Y. Transition-Metal-Free

Approach

to

Synthesis

of

Indolines

from

N-(ortho-Chloromethyl)aryl Amides and Iodonium Ylides. Tetrahedron Lett. 2013, 54, 6049−6052. (e) Zhan, G.; Shi, M.-L.; He, Q.; Du, W.; Chen, Y.-C. [4+3] Cycloadditions with Bromo-Substituted Morita-Baylis-Hillman Adducts of Isatins and N-(ortho-Chloromethyl)aryl Amides. Org. Lett. 2015, 17, 4750−4753. (f) Steinhagen, H.; Corey, E. J. A Convenient and Versatile Route to Hydroquinolines by Inter- and Intramolecular Aza-Diels-Alder Pathways. Angew. Chem., Int. Ed. 1999, 38, 1928−1931. (g) Liu, J.-Y.; Lu, H.; Li, C.-G.; Liang, Y.-M.; Xu, P.-F. Lewis Base and Brønsted Base Dual-Catalyzed Formal [4+3] Cyclo-Addition Reaction: Synthesis of Aza-Spirocycloheptane Oxindole. Synlett 2016, 27, 1287−1291. (h) Hovey, M. T.; Check, C. T.; Sipher, A. F.; Scheidt, K. A. N-Heterocyclic-Carbene-Catalyzed Synthesis of 2-Aryl Indoles. Angew. Chem., Int. Ed. 2014, 53, 9603−9607. (22) Hudhomme, P. Diels-Alder Cycloaddition as an Efficient Tool for Linking π-Donors onto Fullerene C60. C. R. Chimie, 2006, 9, 8812−891. (23) (a) Hou, H.-L.; Li, Z.-J.; Gao, X. Reductive Benzylation of C60 Imidazoline with a Bulky Addend. Org. Lett. 2014, 16, 712−715. (b) Liu, T.-X.; Li, F.-B.; Zhou, D.-B.; Wang, G.-W. Ferric Perchlorate Promoted Reaction of [60]Fullerene with N−Sulfonyl Aldimines: Synthesis and Functionalization of Fulleroxazolidines. J. Org. Chem. 2015, 80, 11986−11992.

ACS Paragon Plus Environment

Page 37 of 37 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

The Journal of Organic Chemistry

(24) Frisch, M. J.; et al. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, 2013. (25) (a) Becke, A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648−5652. (b) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. (26) (a) Yoshida, M.; Maeyama, Y.; Shishido, K. Regio- and Enantioselective Synthesis

of

Functionalized

Tetrahydroquinolines

by

Palladium-Catalyzed

Cyclization of 2-Amidophenylmalonates with Allylic Bisacetates. Tetrahedron 2012, 68, 9962−9972. (b) Sriramurthy, V.; Kwon, O. Diphosphine-Catalyzed Mixed Double-Michael Reaction: A Unified Synthesis of Indolines, Dihydropyrrolopyridines, Benzimidazolines,

Tetrahydroquinolines,

Tetrahydroisoquinolines,

Dihydrobenzo-1,4-oxazines, and Dihydrobenzo-3,1-oxazines. Org. Lett. 2010, 12, 1084−1087.

ACS Paragon Plus Environment