Sunteți pe pagina 1din 44

Subscriber access provided by UNIV NAC AUT DE MEXICO UNAM

Article
Achieving Nickel Catalyzed C-S Cross Coupling under
Mild Conditions using Metal-Ligand Cooperativity
Rina Sikari, Suman Sinha, Siuli Das, Anannya Saha, Gargi Chakraborty, Rakesh Mondal, and Nanda D. Paul
J. Org. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.joc.9b00075 • Publication Date (Web): 11 Mar 2019
Downloaded from http://pubs.acs.org on March 14, 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 43 The Journal of Organic Chemistry

1
2
3
4 Achieving Nickel Catalyzed C-S Cross Coupling under Mild
5
6 Conditions using Metal-Ligand Cooperativity
7
8
9
10
11
12 Rina Sikari,† Suman Sinha,† Siuli Das,† Anannya Saha,‡ Gargi Chakraborty,† Rakesh Mondal,†
13
14 and Nanda D. Paul†,*
15
16
17
18
19

20 Department of Chemistry, Indian Institute of Engineering Science and Technology, Shibpur,
21
22 Botanic Garden, Howrah 711103, India
23
24
25 ‡
Department of Chemical Sciences, Indian Institute of Science Education and Research Kolkata,
26
27
Mohanpur 741246, India
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 Paragon Plus Environment
The Journal of Organic Chemistry Page 2 of 43

1
2
3 ABSTRACT: A simple and efficient approach of C−S cross coupling of a wide variety of
4
5
6
(hetero)aryl thiols and (hetero)aryl halides under mild conditions, mostly at room temperature;
7
8 catalyzed by well defined singlet di-radical Ni(II)-catalysts bearing redox noninnocent ligands is
9
10 reported. Taking advantage of ligand centered redox events; the high-energetic Ni(0)/Ni(II) or
11
12 Ni(I)/Ni(III) redox steps were avoided in the catalytic cycle. The cooperative participation of
13
14
both nickel and the coordinated ligands during oxidative addition/reductive elimination steps
15
16
17 allowed us to perform the catalytic reactions under mild conditions.
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 Paragon Plus Environment
Page 3 of 43 The Journal of Organic Chemistry

1
2
3
INTRODUCTION. Formation of C−S bonds remains one of the most important challenges in
4
5
6 organic synthesis because of the prevalence of aromatic thioethers in a wide spectrum of natural
7
8 and synthetic organic compounds (Figure 1).1 Over the decades, notable progress have been
9
10
made in developing catalytic and stoichiometric C − S bond forming reactions via direct
11
12
13 functionalization of aryl halides, paricularly aryl bromides and iodides (Scheme 1).1-6 Majority of
14
15 the already developed transition metal catalyzed strategies for C−S cross coupling reactions are
16
17
18 based on expensive and relatively less abundant palladium based catalysts which requires typical
19
20 ligand design to achieve efficient catalysis.2,3
21
22
23 In recent years photoredox catalysis has emerged as a powerful alternative method for
24
25 carrying out wide variety of reactions under mild conditions, including C−S bond formation.4
26
27
28 However, these methods are associated with expensive ruthenium or iridium based photo-
29
30 sensitizers which may introduce a bar in terms of sustainability and scalability of these methods.
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52 Figure 1. Selected examples of medicinally and biologically important thioethers.
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 4 of 43

1
2
3 In this perspective, the use of catalysts based on cheap and abundant base-metals is
4
5
6
especially beneficial for large-scale and industrial applications, where the paucity and cost of the
7
8 noble metals like palladium, iridium or ruthenium can be problematic. However, the base metals
9
10 tend to react via single electron transfer process and the corresponding two-electron transfer
11
12 processes are mostly enegetically uphill requiring high temperatures.7 In recent yeras,
13
14
notwworthy progress has also been achieved with cheap copper1d,5 and nickel6 catalysts, however
15
16
17 these protocols almost always requires high temperatures for the reaction to proceed.
18
19
20 Scheme 1
21
22 Previous work
S
23 X + HS
Pd Catalyst

24 R R' Base, Solvent R R'


25 expensive & scarce catalyst
26 Cu or Ni Catalyst S
27 X + HS
R R' Base, Solvent R
28 R'
required elevated temperature
29 works best with aryl bromides and iodides
30
Photoredox S
31 X + HS
catalysis
32 R R' Base, Solvent
R
RT R'
33 required expensive photo-catalyst
34 Present work
35 NiII Catalyst S
36 X + HS
R R' Base, Solvent R
37 X=I, Br, Cl
R'
cheap & earth abundant catalyst
38 easy to prepare; air stable
39 efficient with aryl iodides and bromides
moderate with aryl chlorides
40
mild reaction conditions
41 RT (for X = Br, I); 550 C (for X = Cl)
42 broad substrate scope (> 40)
43
44
45 The use of redox noninnocent ligands in combination with the base-metals has emerged
46
47 as an attractive alternative.7,8 The ability of redox noninnocent ligands to store and release
48
49 electrons when required, allows multi-electron transformations (mostly preferred for noble
50
51
52
metals like palladium) with inexpensive 3d-base metals. In recent times, using the redox
53
54 noninnocent properties of imine, diamine and aminophenol based scaffolds, various new
55
56 stoichiometric and catalytic reactions were developed where these organic scaffolds were found
57
58
59
60 ACS Paragon Plus Environment
Page 5 of 43 The Journal of Organic Chemistry

1
2
3 to participate actively during electron transfer processes involved during catalytic turnover.8
4
5
6
However, no systematic studies were carried out to explore catalytic C−S coupling reactions
7
8 using transition metal complexes of redox noninnocent ligands.
9
10
11 Herein we report the synthesis of a wide variety (47 examples) of aromatic thioethers via
12
13 cross coupling of (hetero)aryl halides(ArX, where X = I, Br, or Cl) and (hetero)aryl thiols
14
15 catalyzed coordinatively unsaturated Ni(II)-complexes containing redox noninnocent diamine
16
17
ligands where the redox processes are expected to be in part (or entirely) ligand centered. To
18
19
20 avoid high-energy Ni(0)/Ni(II) or Ni(I)/Ni(III) redox steps during catalysis we wanted to (partly)
21
22 utilize the low enegy ligand centered redox processes, by either leaving nickel in the +II
23
24 oxidation state throughout the entire catalytic cycle or enabling less energetic nickel(II)-
25
26 nickel(III) oxidative addition/reductive elimination steps. Notably, the reactions were found to
27
28
29
proceed efficiently at room temperature with aryl bromides and iodides affording the
30
31 corresponding thioethers in moderate to good yields. With aryl chlorides, the C-S corss coupling
32
33 reaction did nor proceed at room temperature. At slightly higher temparture (55 °C) the desired
34
35 products were obtained in moderate yields wheras the highest yield was obtained at 80 °C (Table
36
37 2, entry 1) with aryl chlodide.
38
39
40
41
42 RESULT AND DISCUSSION. Three tetracoordinate singlet di-radical Ni(II)-complexes, 1-3
43
44
45
featuring two antiferromagnetically coupled one-electron oxidized redox noninnocent dimaine
46
47 type ligands {(L1-3)•-} with the valence configuration of [NiII{(L1-3)•-}2] (1-3) were employed to
48
49 study the above mentioned C−S cross coupling reactions (Figure 2).9 In all the three catalysts,
50
51
the central Ni(II)-center is coordinated by two o-diiminosemiquinonato type ligands in a
52
53
54 distorted square planner geometry. In catalyst 1, and 2, the Ni(II) ion is bound to two one
55
56 electron oxidized o-phenylenediamine (L1) and two N-phenyl-o-phenylenediamine (L2) ligands
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 6 of 43

1
2
3 respectively, whereas in catalyst 3, two one electron oxidized tridentate N-(2-
4
5
6
aminophenyl)benzene-1,2-diamine (L3) ligands are bound to the Ni(II) centre in a bidentate
7
8 mode with one uncoordinated amine arm from each of the ligands.9c
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
Figure 2. Catalysts used in this study.
26
27
28
29
30 During transition metal catalyzed cross coupling of thiols with aryl halides, mostly aryl
31
32 iodides, and bromides are used and the use of aryl chlorides as the substrate is limited because of
33
34 their poor reactivity.6a,b Therefore, during our initial studies we focused on the possibility of
35
36
cross coupling of thiols with chloro-, bromo- and iodobenzene separately. The reactions of
37
38
39 thiophenol (4a) with chloro- (5a'), bromo- (5a") and iodobenzene (5a) were performed
40
41 independently under various reaction conditions to obtain the optimal conditions and to check
42
43 the effects of ligands and solvents during the catalytic C−S bond formation reactions using
44
45
46 Ni(II)-complexes, 1-3 as catalysts.
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 7 of 43 The Journal of Organic Chemistry

1
2
3 Table 1: Optimization of the Reaction Conditions.a-d
4
5
6 S
7 Ni-Catalyst (mol%)
SH I
8 Base, Solvent
9
Temperature (oC)
10 (4a) (5a) (6aa)
11
12
13
Entry Ni-catalyst (mol%) Solvent Base Time (h) Yield (%)
14
15 1 3 (5.0mol%) DMF NaOtBu 24 68
16 2 3 (5.0mol%) CH3CN/DMF NaOH 24 28
17 3 3 (5.0mol%) CH3CN/DMF Cs2CO3 24 NR
18 4 3 (5.0mol%) CH3CN/DMF K2CO3 24 NR
19
5 3 (5.0mol%) CH3CN/DMF K3PO4 24 NR
20
21 6 3 (5.0mol%) CH3CN/DMF KOH 24 NR
t
22 7 3 (5.0mol%) CH3CN/DMF KO Bu 24 78
23 8 3 (5.0mol%) CH3CN/DMF NaOtBu 24 87
24 9 3 (5.0mol%) CH3CN/DMF NaO But
24 69
25 t
10 3 (5.0mol%) CH3CN NaO Bu 24 Trace
26
27 11 3 (5.0mol%) Toluene NaOtBu 24 NR
t
28 12 3 (5.0mol%) Dioxane NaO Bu 24 NR
29 13 3 (5.0mol%) THF NaOtBu 24 NR
30 14 2 (5.0mol%) CH3CN/DMF NaO But
24 67
31 t
15 1 (5.0mol%) CH3CN/DMF NaO Bu 24 NR
32
33 16 L3 CH3CN/DMF NaOtBu 24 NR
d t
34 17 Ni(COD)2/DPPF CH3CN/DMF NaO Bu 24 <10
35 18d Ni(COD)2/Xanthphos CH3CN/DMF NaOtBu 24 NR
36 19 d
Ni[P(OPh)3]4/DPPF CH3CN/DMF NaO But
24 <10
37 d t
20 Ni[P(OPh)3]4/Xanthphos CH3CN/DMF NaO Bu 24 Trace
38
39 21 Ni(ClO4)2.6H2O (10.0 mol%) CH3CN/DMF NaOtBu 24 NR
t
40 22 NiCl2.6H2O (10.0 mol%) CH3CN/DMF NaO Bu 24 NR
41 23 3(5.0 mol%) + 2 equiv. TEMPO CH3CN/DMF NaOtBu 24 NR
42 24 3(5.0 mol%) + 2 equiv. DPPH CH3CN/DMF NaO But
24 NR
43 a
Stoichiometry: thiol (4a) (1.0 mmol; 1.0 equiv.) and aryl halide (5a) (1.2 mmol; 1.2 equiv). b1.0
44
45
equiv of base. cTemperature: RT. dNickel salts: 10 mol%; ligands: 20mol%.
46
47
48
49
50 Using 3 as the catalyst, the cross-coupling reactions of thiophenol (4a) with bromo- (5a")
51
52 and iodobenzene (5a) are found to proceed efficiently at room temperature in CH3CN/DMF (5:1)
53
54 solvent mixture. Among the series of bases tested, NaOtBu was found to be most efficient. At
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 8 of 43

1
2
3 room temperature, using iodobenzene a maximum of 87% of diphenylsulfide (6aa) was isolated
4
5
6
with 5.0 mol% of catalyst loading in presence 1.0 mmol of NaOtBu as base (Table 1, entry 8)
7
8 whereas bromobenzene afforded 60% of 6aa. However, chlorobenzene requires slightly higher
9
10 temperature, at 55°C, diphenylsulfide (6aa) was isolated in 51% yield. Interestingly, with
11
12 increase in the reaction temperature, the yield of 6aa increases with bromo- and chlorobenzene.
13
14
With bromobenzene, at 50°C, the yield of 6aa increases to 75%. However, a maximum of 70%
15
16
17 6aa was isolated at 80°C using chlorobenzne as the electrophile (Table 2, entry 1). In all the
18
19 cases, the yield of 6aa decreases significantly upon lowering the catalyst loading below 5.0
20
21 mol%, however, no significant increase of yields were noticed with higher catalyst loading.
22
23 Among the three different Ni(II)-catalysts studied, best results were obtained with catalyst 3,
24
25
whereas using catalyst 2 only 67% of the desired C−S coupled product was obtained (Table 1,
26
27
28 entry 14) at room temperature starting from thiophenol (4a) and iodobenzene (5a).
29
30 Control experiments were performed to check the role of base as well as role of the
31
32 coordinated redox active ligands during the catalytic C-S cross coupling. In absence of catalyst, 3
33
34 and NaOtBu, no desired C-S coupled product was obtained. In absence of 3, when the reactions
35
36
37
were carried out in presence of only NaOtBu, no C−S coupled products were obtained while
38
39 using unsubstituted aryl halides or aryl halides containing electron donating groups. However,
40
41 with aryl halides having electron withdrawing substituents the corresponding C−S coupled
42
43 products were isolated in trace amount (<10%) at RT in presence of only NaOtBu under our
44
45
optimized reaction conditions. No desired products were also obtained only in presence of the
46
47
48 ligand L3. Other Ni-salts like NiCl2.6H2O, Ni(ClO4)2.6H2O, Ni(COD)2 and Ni[P(OPh)3]4 when
49
50 used as the catalyst, no desired C−S coupled products were obtained at room temperature under
51
52 our optimal reaction conditions. No desired C−S coupled products were also obtained when the
53
54 cross coupling of thiophenol (4a) and iodobenzene (5a) were carried out using these nickel salts
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 9 of 43 The Journal of Organic Chemistry

1
2
3 in presence of ligands like DPPF and Xantphos under our optimized conditions. Using these
4
5
6
ligands, reactions also did not proceed at room temperature under the optimized reaction
7
8 conditions reported by Stewart and co-workers (Table 1, entries 17-20).6a Literature data reveals
9
10 that with other nickel catalysts containing various redox non-active phosphines, N-heterocyclic
11
12 carbene based ligands almost always require high temperature (1100C-800C) for C-S cross-
13
14
15 coupling reactions (Figure 1).6a-i Interestingly, C−S cross coupling reactions requires high
16
17 temperature to proceed even with nickel complexes bearing redox active ligands such as
18
19 bipyridine,6k diiminopyridine,6l etc where the C−S cross coupling involves only the nickel
20
21 centered redox events and the coordinated ligands remain as spectator.6j-l Literature data reveals
22
23
that irrespective of the nature of the coordinated ligands (redox active or redox non-active) the
24
25
26 nickel catalyzed C-S cross coupling reactions which involve only the nickel centered redox
27
28 events (Ni(0)/Ni(II) or Ni(I)/Ni(III)) almost always require high temperature (Figure 3).6
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 Paragon Plus Environment
The Journal of Organic Chemistry Page 10 of 43

1
2
3 S
Ni-Catalyst
4 SH X
Base, Solvent
5 X=Cl, Br
6
7 Ligand Pre-Catalyst type Ligand Pre-Catalyst type
8
9 NiBr2-bpy [L/NiCl2]
N N
10 O O
11 Condition Additive PR2 PR2

12 110oC Zn R=Ph, Me, iPr, tBu Reference : 6e


13
14 Reference : 6h Condition
80oC
15
16
17 Ligand Pre-Catalyst type Ligand Pre-Catalyst type
18
19 [L/NiX2] [L/Ni(COD)2]
Cl-
20 N N O
21 P P
O H H O Reference : 6c
22 Reference : 6d Ph Ph
23
Condition o Condition
24 110 C
80oC
25
26 Ligand Pre-Catalyst type Ligand Pre-Catalyst type
27
[L/NiCl2] [L/NiCl2]
28
29 N N
H
Reference : 6i R2P PR2 Reference : 6g
30 N N
Arf Arf t
31 R=Ph, Cy, Bu
32 Condition Condition
33 110oC-160oC
800C
34
35 Ligand Pre-Catalyst type Ligand Pre-Catalyst type
36 L/Ni[P(Op-Tolyl)3]4 L/Ni[P(Op-Tolyl)3]4
Ph2P PPh2
37
38 Condition Additive O
110oC PR2 PR2 Reference : 6a
39 Zn t
40 R=Ph, Bu
Reference : 6a
41 Condition
42 110oC
43
44
45
46 Figure 3. Examples of some nickel catalyzed C−S cross-coupling reactions involving metal
47
48 (Ni)-centered redox events.
49
50
51 Having the optimized conditions in hand, we studied the substrate scope of this nickel
52
53 catalyzed C−S cross coupling reactions. A diverse range of thiophenols bearing electron
54
55
donating, withdrawing, and heterocyclic functionalities were reacted with iodo-, bromo- and
56
57
58
59
60 ACS Paragon Plus Environment
Page 11 of 43 The Journal of Organic Chemistry

1
2
3 chlorobenzene separately. It is worth mentioning that the reactions of iodobenzene with thiols
4
5
6
were performed at room temperature while the reactions of bromo- and chlorobenzenes were
7
8 carried out at two different temperatures and the results are summarized in Table 2.
9
10
11 Aryl thiols containing both electron donating and withdrawing functionalities were found
12
13 to be effective in the catalytic C−S cross coupling reactions with aryl halides (X = I, Br, Cl)
14
15 affording the corresponding C−S-coupled products in moderate to good yields. S−arylated
16
17
products were obtained in higher yield with aryl halides bearing electron donating groups at the
18
19
20 ortho-, para- or meta- positions (Table 2, entries 2-5). At room temperature the reaction of iodo-
21
22 and bromobenzene with ortho-, meta-, and para-thiocresol produced the corresponding
23
24 diarylsulfides (6ab-ad) in 67%, 78% and 82% (with iodobenzene) and 45%, 61% and 63% (with
25
26 bromobenzene) yields respectively. The same reactions with bromo- (at 50°C) and
27
28
29
chlorobenzene (at 55°C) afforded the corresponding diarylsulfides in 54%, 69% and 71% (with
30
31 bromobenzene) and 42%, 58% and 65% (with chlorobenzene) yields respectively.
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 Paragon Plus Environment
The Journal of Organic Chemistry Page 12 of 43

1
2
3 Table 2: Substrate Scope of Thiols with Aryl Chlorides/Bromides/Iodides.a-g
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 Aryl thiols containing electron withdrawing groups were also found to be suitable,
42
43 although the corresponding S−arylated products were obtained in moderate to good yield (Table
44
45 2, entries 10-12). It is worth to mention that aryl thiols bearing free amine or hydroxyl groups
46
47 were also tolerated under the optimal conditions (Table 2, entry 6, 8, 9). Excellent
48
49
chemoselectivity was observed when ortho- and para-hydroxy thiols were reacted with aryl
50
51
52 iodide and bromide under the optimized reaction conditions; only the C−S coupled product was
53
54 formed, and no C−O bond formation was observed. However, with chlorobenzene at 80°C, other
55
56 than C−S, small amount of C−O coupled product was also isolated. 2-naphthalenethiol was also
57
58
59
60 ACS Paragon Plus Environment
Page 13 of 43 The Journal of Organic Chemistry

1
2
3 found to be compatible, affording the desired C−S coupling product in 65% isolated yield at RT
4
5
6
with iodobenzene (Table 2, entry 13). Heterocyclic functionalities were also tolerated under the
7
8 optimized reactions conditions producing the corresponding diarylsulfides in moderate to good
9
10 isolated yields (Table 2, entries 14, 15). The C−S cross coupling reactions also proceeded with
11
12 alkyl thiols and iodobenzene (Table 2, entry 16).
13
14
15 To further substantiate the substrate scope, various aryl iodides, bromides and chlorides
16
17 were used as the electrophile to couple with thiophenol. Aryl halides bearing different electronic
18
19 functionalities were all found to be suitable yielding the corresponding diarylsulfides in good to
20
21 moderate yields (Table 3). Both activated and unactivated aryl halides are equally effective
22
23
24
affording the corresponding diarylsulfides in almost comparable yields. At room temperature, the
25
26 reactions of aryl iodides containing electron donating groups at ortho-, para-, and meta-
27
28 positions produced the corresponding diarylsulfides in 63-85% yields (Table 3, entries 17-20).
29
30 Almost identical yields were obtained with aryl bromides at 50°C, whereas slightly lower yields
31
32 were obtained with chlorobenzene even at 80°C. Napthyl and pyrene iodides were also found to
33
34
35 be compatible, yielding the corresponding diarylsulfides in 61 and 65% yields at room
36
37 temperature (Table 3, entries 34-36) respectively. Heterocyclic iodides such as, 2-iodothiophene,
38
39 3-iodopyridine produced the corresponding diarylsulfides in 60% and 65% yields respectively
40
41 (Table 3, entries 37, 38). Long chain alkyl iodides were also tolerated under our optimal
42
43
conditions (Table 3, entry 39).
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 14 of 43

1
2
3 Table 3: Substrate Scope aryl Chlorides/Bromides/Iodides.a-g
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 Since arylthioethers are ubiquitous in various pharmaceuticals, we explored the potential
49
50 of our methodology to bring about late stage thiolation of some selected pharmaceutically
51
52 relevent molecules such as Hydrochlorothiazide, Fenofibrate, Monoclobemide and Indometacin
53
54 under our optimized reaction conditions. To our delight the desired thiolated compounds (Table
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 15 of 43 The Journal of Organic Chemistry

1
2
3 4, entries 40-45) were obtained in 52%−72% yield. We also synthesized the key structure of 11β-
4
5
6
HSD1 inhibitors (Table 4, entry 46) in 50% yield via the reaction of 2-chlorothiophenol and 4-
7
8 bromoacetophenone under our optimized reaction conditions.
9
10
11
12
13
14 Table 4: Late Stage Functionalization of Some Pharmaceutically Relevant Molecules.
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 Finally to understand the mechanism of the developed catalytic methodology and to
42
43 confirm our hypothesis of using ligand centered redox events avoiding Ni(0)/Ni(II) or
44
45 Ni(I)/Ni(III) redox couples during catalytic turnover, some control experiments were carried out.
46
47
48 Since the catalysts used in this study are all redox active and the redox processes are expected to
49
50 occur at the ligand centers generating organic radicals upon oxidation or reduction, we decided to
51
52 check the catalytic reaction in presence of a radical inhibitor to confirm the involvement of
53
54 ligand centered redox as well as to check the involvement of any organic radicals during catalytic
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 16 of 43

1
2
3 turnover. Interestingly, no C−S coupled product was obtained in presence of TEMPO or DPPH
4
5
6 (∼2.0 equiv with respect to the catalyst; TEMPO = 2,2,6,6-tetramethyl piperidinoxyl; DPPH =
7
8 2,2-diphenyl-1-picrylhydrazyl; Table 1, entries 19-20).
9
10 An immediate change of color of the catalysts 2 and 3 was observed upon addition of
11
12
13
NaOtBu. Since, alkali metal tertiary butoxides, other than acting as base can also act as a
14
15 reducing agent,10 the observed color change may lead to confusion whether it is due to
16
17 deprotonation of any of the −NH group of the coordinated ligand or it is due to the reduction of
18
19 the coordinated ligands. Recently, in our work with catalysts 2 and 3, we revealed that in
20
21
presence of KOtBu the catalysts 2 and 3 undergo simple deprotonation.9d Our present
22
23
24 experimental results in presece of NaOtBu also indicates that the base seems to act as a base
25
26 (deprotonating agent) rather than a reducing agent in these reactions and the deprotonated
27
28 complex [3]─ possibly plays the role of the active catalyst. Interestingly, when the stoichiometric
29
30 reaction of iodobenzene and thiophenol was performed in absence of any added base using the
31
32
33
preformed [3]─ as the catalyst, 80% of diphenylthiol (6aa) was isolated confirming the active
34
35 participation of [3]─ during catalysis.
36
37 Next, to explore the feasibility of single electron transfer from [3]─ to the aryl halide,
38
39 generating aryl radicals, a control reaction was carried out with thiophenol (4a) and 1-(allyloxy)-
40
41
42 2-iodobenzene (5y) under the optimized reaction conditions (Scheme 2).4b Characterization of
43
44 the reaction mixture confirms the formation of the C−S coupling product 9-(2-(allyloxy)phenyl)-
45
46 thiophenol (6ya) and no intramolecular cyclization product (that would have been formed if the
47
48 reaction proceeds through single electron transfer pathway) was found. 9-(2-(allyloxy)phenyl)-
49
50
thiophenol (6ya) was isolated in 52% yield and ~44% of unreacted 1-(allyloxy)-2-iodobenzene
51
52
53 (5y) was recovered from the reaction mixture. This result indeed eliminates mechanistic
54
55 pathways involving single electron transfer from the catalyst to the aryl halides and it was also
56
57
58
59
60 ACS Paragon Plus Environment
Page 17 of 43 The Journal of Organic Chemistry

1
2
3 conclude that our catalyst was efficient to from C-S coupled product when iodo substrate was 1-
4
5
6
(allyloxy)-2-iodobenzene.
7
8
9
10 SPh
11 Catalyst 3 (5.0 mol%)
I
12 NaOtBu (1.0 equiv.)
O O
47
13 CH3CN/DMF (5:1) O
O
14 RT, 24h Unreacted starting materials
15 (5y)
SPh
16 (6ya) 52%
17
18 Scheme 2. Control experiment to check the possibility of single electron transfer pathway.
19
20
21
22
23 For further understanding of the reaction mechanism and to check the role of the ligand
24
25 centered redox events, DFT studies were carried out at the B3LYP level using LANL2DZ basis-
26
27 set for Ni, 6-311+G(d) basis-set for N and the 6-31G(d,p) basis-set for C, H. (see experimental
28
29
section and the Supporting Information for details).12-17 Broken-symmetry calculations were
30
31
32 performed following the formalism as introduced by Noodleman18 to achieve the spin-polarized
33
34 symmetry-broken solutions for all the intermediates involved in the catalytic reactions.19,20
35
36 Efforts were given to understand the plausible electronic structures of the reaction intermediates
37
38 as well as to reveal the involvement of the redox couples involved in the oxidative
39
40
41
addition/reductive elimination steps. Computed low-energy barriers are displayed in Figure 4.21
42
43 As evident from the experimental results described above, it seems that in presence of
44
45 NaOtBu, the complex 3 undergoes deprotonation9d and the deprotonated complex [3]─ plays the
46
47
role of the active catalyst. Therefore, electronic structure of [3]─ was first studied. As was
48
49
50 reported with 3,9 the deprotonated complex [3]─ was also found to be singlet di-radical species
51
52 containing two one electron oxidized semiquinone radicals coupled antiferromagnetically with
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 18 of 43

1
2
3 an electronic structure description of [NiII(L3)•−{(L3)−}•−]− (one extra negetive charge is for the
4
5
6
initial deprotonation of L3) (See SI).
7
8 The activation of the Ar-I bond was then studied. Upon introduction of PhI moiety to 3 in
9
10 presence of tBuO─, the catalyst 3 undergoes deprotonation to form the active species [3]─ which
11
12
then losses its square plannar geometry and both the ligand become orthogonal to each other
13
14
15 which indeed creates some extra space for PhI to aproach close to [3]─ and its C–I bond gets
16
17 activated in transition state (3ATS). The transition state 3ATs for oxidative addition was found to
18
19 have a barrier of 26.3 kcalmol-1 relative to 3. The intermediate 3B, formed after oxidative
20
21 addition was found to be penta-coordinate with one iodide anion at a distance of 4.5 Å from the
22
23
24
nickel center. Formation of 3B is exergonic by –12.8 kcalmol-1 relative to 3. Transition state
25
26 barrier was also calculated starting with the deprotonated species [3]─ to check the feasibility of
27
28 the step-wise deprotonation of 3 to form [3]─ followed by oxidative addtion of PhI on the
29
30 preformed [3]─ (Figure S6). In both the cases the activation of Ph-I goes through the same
31
32 activation energy of 26.3 kcalmol-1 relative to 3, however, with respect to the preformed [3]─, the
33
34
35 transition state barrier was found to be slightly higher 29.2 kcalmol-1. On the other hand, the
36
37 direct activation (oxidative addition) of Ph-I on 3, goes through a transition state (3ATS') energy
38
39 barrier of 29.1 kcalmol-1 (Figure S7). This indeed indicates that the base (tBuO─) plays an
40
41 important role during catalysis and upon deprotonation the oxidative addition is favored.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 19 of 43 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 Figure 4. The relative free energy profile with scehmatic presentation for C–S crosscoupling
36
37
38 reaction of PhI and PhSH catalyzed by 3.
39
40
41
42
43
It is important to note that during oxidative addition the two electron transfer from the
44
45 catalyst can occur either involving energetically demanding exclusively nickel centered redox-
46
47 couple, Ni(II)/Ni(IV) or both Ni(II) and one of the coordinated ligand can participate in a
48
49 synergistic way by loosing one electron each, thus leading to comparitively low energy
50
51 Ni(II)/Ni(III) and {(L3)•−}/{(L3)0} or {(L3)•2−}/{(L3)−}(extra negative charge results from the
52
53
54 initial deprotonation) redox couples. The possibility of two electron transfer from the two
55
56 coordinated ligands can also not be excluded. Therefore, to understand the possible redox
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 20 of 43

1
2
3 couples involved during the oxiative addition, the electronic structures of 3ATs and the
4
5
6
intermediate (3B) obtained after the oxidative addition were analysed throughly. In the transition
7
8 state 3ATs, the nickel center possesses +1.44 spin and the deprotonated ligand contains –0.16
9
10 spin whereas –0.82 spin resides over the other ligand.
11
12
Analysis of eletronic configuration of 3B reveals following facts: i) +0.83 spin is located
13
14
15 on the nickel center which supports the +3 oxidation state of nickel suggesting the involvement
16
17 of the Ni(II)/Ni(III) redox couple during oxidative addition, (ii) the deprotonated ligand contains
18
19 only +0.11 spin, indicating the existence of fully oxidized quinone form [(L3)]− of the ligand
20
21 indicating involvement of {(L3)•2−}/{(L3)−} redox couple during oxidative addition, here extra
22
23
24
negative charge results from the initial deprotonation (iii) the other ligand contains –0.87 spin
25
26 which supports monoanionic semiquinone radical oxidation state [(L3)•−] (See SI, Figure S6, S8).
27
28 The electronic configuration of 3B can be expressed as {[NiIII{(L3)•−}{(L3)−}(Ph)]I}−. Thus DFT
29
30 study indicates that oxidative addition involves Ni(II)/Ni(III) and {(L3)•2−}/{(L3)−} redox
31
32 couples demonstrating metal-ligand cooperative catalysis. Therefore, computational studies
33
34
35 reveal that oxidative addition is favored upon ligand -NH deprotonatio by tBuO−. Possibly,
36
37 because of deprotonation, the coordinated ligand becomes more electron rich which makes it
38
39 more susceptible towards oxidation and that in turn promotes the C–I bond activation under mild
40
41 conditions in combination with Ni(II)/Ni(III) redox couple.
42
43
44 In the next step, the nucleophile PhSH coordinates to Ni(III) center as PhS– and the
45
46 proton is transferred to the doubly oxidized ligand leading to 3C which is exergonic by –16.0
47
48 kcalmol-1 compared to 3B. The energy difference for the step 3B to 3C was calculated as (3C+I–)
49
50 – (3B + PhSH). In 3C, nickel center remains in +3 oxidation state as evident from +0.70 spin on
51
52
53
nickel. No substantial amount of spin (0.01) was found on the fully oxidized quinone form of the
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 21 of 43 The Journal of Organic Chemistry

1
2
3 ligand, as expected. The other ligand contains –0.88 spin indicating its monoanionic
4
5
6
semiquinone radical oxidation state [(L3)•−] as was there in 3B.
7
8 Next, attention was paid to understand the reductive elimination step. Reductive
9
10 elimination was found to pass through the transition state 3CTS with energy barrier of 8.3
11
12 kcalmol−1. The TS energy barrier for the non-catalyzed C–S coupling reaction is calculated to be
13
14
29.9 kcalmol−1 which is 21.6 kcalmol−1 higher than the catalyzed reaction. It is important to note
15
16
17 that the transition state energy barrier (26.3 kcalmol−1) for oxidative addition is larger than the
18
19 energy barrier for reductive elimination (8.3 kcalmol−1). This suggests that the oxidative addition
20
21 is the rate determining step for this reaction.
22
23
24
After reductive elimination it leads to an intermediate 3D where nickel center is reduced
25
26 to Ni(II) as revealed from its spin population of +1.47. Along with metal center reduction, the
27
28 deprotonated doubly oxidized ligand in 3D also gets reduced to its monoanionic radical state
29
30 (spin population is –0.73) during C–S coupling. This also supports synergestic metal-ligand
31
32 participation in the catalytic cycle. The coupled product diphenylsulfide is released in this
33
34
35 process. 3D was found to be exergonic by –28.9 kcalmol−1 relative to 3C.
36
37 Thus, computational study performed herein reinforces the following facts; i) initial
38
39 t
BuO− mediated ligand NH deprotonation facilitates the activation of Ph–I via oxidative addition
40
41
under mild conditions, ii) the catalytic cycle utilizes combined ligand and Ni(II)/Ni(III) redox
42
43
44 events, iii) the oxidative addition was found to be rate limiting step in the catalytic cycle.
45
46 Based on the above experimental results and available literature a plausible mechanism is
47
48 outlined in Scheme 3. The reaction is supposed to progress via tBuO– mediated ligand -NH
49
50
51 deprotonation of catalyst 3 followed by oxidative addition of Ph-I to form the intermediate 3B.
52
53 The increased electron density in [3]─ likely enables the oxidative addition of the ArI substrate
54
55 under mild conditions. During oxidative addition the nickel center is oxidized by one electron to
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 22 of 43

1
2
3 Ni(III) and another electron is provided by the deprotonated ligand. In the next step PhSH
4
5
6
coordinate to Ni(III) center as PhS– and the proton is transferred to the doubly oxidized ligand to
7
8 form intermediate 3C which upon reductive elimination produce the diarylsulfide.
9
10
11
12
13
14
15 - Na+
16 R
17 H
N II N
R
H
NaOtBu
18 Ni N II N
- tBuOH Ni I
19 N
H
N
N N
20 R
R
21 3 [3]-
22
R = H, NH2
23 NaOtBu
24 S
25
26 R -
I
27 R S N III
N III Ni NH
28 Ni NH N
29 N N
H
30 N 3B
3C
31 R
R

32
33 SH
34
35
36
37 Scheme 3. Proposed Mechanism.
38
39
40
41
42 Conclusion. In summary, we have reported a simple and efficient approach of C–S bond
43
44
formation of a wide variety of (hetero)aryl thiols and (hetero)aryl halides under mild conditions
45
46
47 mostly at room temperature (with aryl iodides/bromides), catalyzed by simple, well defined, easy
48
49 to prepare and cheap Ni(II)-catalysts bearing redox noninnocent ligands. Taking advantage of
50
51 ligand centered redox events; high-energetic Ni(0)/Ni(II) or Ni(I)/Ni(III) redox steps were
52
53 avoided in the catalytic cycle via synergistic participation of both nickel and the coordinated
54
55
56
ligands during oxidative addition/reductive elimination steps. As revealed from DFT studies,
57
58
59
60 ACS Paragon Plus Environment
Page 23 of 43 The Journal of Organic Chemistry

1
2
3 both nickel and one of the coordinated ligand undergo one electron trasfer during oxidative
4
5
6
addition/reductive elimination. The successful development of similar approaches avoiding high
7
8 energetic Ni(I)/Ni(III) redox couple is expected to open many new research avenues aimed at the
9
10 future replacement of expensive noble metal catalysts in various cross-coupling reactions. Our
11
12 studies in this area is in progress and will be reported in due course.
13
14
15
16
17 EXPERIMENTAL SECTION
18
19
20 General Information. Unless otherwise stated, all the reactions were performed under argon
21
22 atmosphere using standard Schlenk technique. All the solvents were dried following standard
23
24 drying procedure. Aryl halides and substitutedthiophenols werepurchased from Sigma-Aldrich,
25
26 TCI and used without further purification. Analytical TLC was performed on a Merck 60 F254
27
28
silica gel plate (0.25mm thickness) and column chromatography was performed on Merck 60
29
30
31 silica gel (60-120 mesh).NMR spectra were recorded on a Bruker DPX-300(300 MHz), Bruker
32
33 DPX-400(400 MHz) and Bruker DPX-500(500 MHz) spectrometers. TMS (tetramethylsilane)
34
35 was used as an internal standard.
36
37
38
Catalyst Synthesis. The catalysts 1-3 were synthesized following the available literature
39
40 procedures.9
41
42 General Procedure for Catalytic C-S Coupling. Under argon atmosphere, Ni-catalyst (5.0 mol
43
44
%) and NaOtBu (1.0 equiv.) were added to an oven dried 25mL Schlenk tube containing a Teflon
45
46
47 coated magnetic stir-bar. To this Schlenk tube, respective aryl halides (1.10 mmol) and
48
49 thiophenols (1.0 mmol) were added via syringe. A mixture of solvent CH3CN/DMF (5:1) was
50
51 added to the reaction vessels through a syringe. Using a rubber septum the Schlenk tube was
52
53 capped and was tightly wrapped with Teflon. The reaction mixture was then degassed with argon
54
55
56
for 10 min and the reaction mixture was kept for stirring at room temperature for 24 hours. The
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 24 of 43

1
2
3 reaction was monitored using TLC. After the reaction was completed, the solvent was removed
4
5
6
through vacuum and the product was purified through column chromatography using a mixture
7
8 of petroleum ether and hexane as an eluent.
9
10 Computational Studies. All DFT calculation were done using theGaussian 16 package11 with
11
12
the Lee–Yang–Parr correlation function (B3LYP).12-13 Grimme’s D3-dispersion correction14 was
13
14
15 employed for the dispersion correction. For Ni and iodide, the Hay and Wad basis-set LANL2DZ
16
17 pseudopotential15 was used, 6-311+G(d)16 basis-set was employed for N and the 6-31G(d,p)17
18
19 basis-set was used for C, H. For all the optimized structures, broken-symmetry states were found
20
21 to be lower in energy compared to their corresponding high-spin states. Broken-symmetry
22
23
24
calculations were performed following the formalism as introduced by Noodleman18 to achieve
25
26 the spin-polarized symmetry-broken solutions.19,20 For complex 3, two unpaired electrons on
27
28 both the ligands could not be separated by broken-symmetry. This could be due to strong anti-
29
30 ferromagnetic coupling between both the ligands in square planar geometry. For all other
31
32 structures, broken-symmetry solutions were obtained as expected. The nature of the ground-state
33
34
35 geometries of each optimized structure, were evaluated by frequency calculations at the same
36
37 level of theory, as that of geometry optimization. No imaginary frequency was obtained for the
38
39 intermediates whereas transition states have only one imaginary frequency. The intrinsic reaction
40
41 coordinate (IRC) calculations revealed that the transition states connect to the right minima. All
42
43
energies reported in the reaction coordinate are Gibbs free energy. All the optimization of the
44
45
46 intermediates and transition states were performed with CPCM polarizable conductor calculation
47
48 model22 employing acetonitrile as solvent. Spin-density plots were made using the Chemcraft
49
50 Visualization program.
51
52
53
Characterization Data of the Isolated Compounds.
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 25 of 43 The Journal of Organic Chemistry

1
2
3 Diphenylsulfane (6aa).4h,5d Eluent: petroleum ether. Colorless oil. Yield c87% (162 mg, for X =
4
5
6
I); c60% & d75% (112 mg & 140 mg, for X = Br); e51% & f70% (95mg & 130mg, for X=Cl).
7
1
8 HNMR (400 MHz, CDCl3): δ = 7.36-7.25 (m, 10H). 13C{1H} NMR (100 MHz, CDCl3) = 135.0,
9
10 130.7, 129.0, 127.1.
11
12
Phenyl (o-tolyl)sulfane (6ab).4g,5d,6c Eluent: petroleum ether. Colorless oil. Yield c67% (134 mg,
13
14
15 for X = I), c45% & d54% (90 mg & 108 mg, for X = Br); e42% (84 mg, for X = Cl). 1HNMR
16
13
17 (400 MHz, CDCl3): δ = 7.38-7.10 (m, 9H), 2.46 (s, 3H). C{1H} NMR (100 MHz, CDCl3) =
18
19 139.9, 138.9, 136.1, 133.8, 133.0, 131.2, 131.0, 130.6, 129.7, 129.3, 129.1, 127.9, 127.2, 127.0,
20
21 126.79, 126.7, 126.3, 21.0.
22
23
24 Phenyl (m-tolyl)sulfane (6ac).4h,5d,6c Eluent: petroleum ether. Colorless oil. Yield c78% (156 mg,
25
26 for X = I), c61% & d69% (122 mg & 138 mg, for X = Br); e58% (116 mg, for X = Cl). 1HNMR
27
28 (400 MHz, CDCl3): δ = 7.33-7.26 (m, 4H), 7.23-7.12 (m, 4H), 7.05 (d, 1H, J = 6.8 Hz). 13C{1H}
29
30
NMR (100 MHz, CDCl3) = 139.0, 136.1, 135.2, 131.9, 130.8, 129.2, 129.1, 128.4, 128.1, 126.9,
31
32
33 21.3.
34
35 Phenyl (p-tolyl)sulfane (6ad).4h,5d,6c Eluent: petroleum ether. Colorless oil. Yield c82% (164 mg,
36
37 for X = I), c63% & d71% (126 mg & 142 mg, for X = Br); e65% (130 mg, for X = Cl). 1H NMR
38
39
40
(400 MHz, CDCl3): δ = 7.29-7.26 (m, 2H), 7.25-7.20 (m, 4H), 7.20-7.13 (m, 1H), 7.11 (d, J = 4
41 13
42 Hz, 2H), 2.23 (s, 3H). C{1H} NMR (100 MHz, CDCl3) = 137.6, 137.2, 132.3, 131.3, 130.1,
43
44 129.8, 129.1, 126.4, 21.2.
45
46
4-(methoxyphenyl) (phenyl)sulfane (6ae).4e,4h,5d Eluent: petroleum ether/ethyl acetate (50:1).
47
48
49 Light yellow solid. Yield c85% (184 mg, for X = I), c61% & d72% (132 mg, 156 mg, for X = Br);
50
e
51 65% (141 mg, for X = Cl). 1H NMR (400 MHz, CDCl3): δ = 7.42 (d, J = 6.4 Hz, 2H), 7.23-7.13
52
53 (m, 5H), 6.88 (d, J = 8.4 Hz, 2H), 2.83 (s, 3H). 13C{1H} NMR (100 MHz, CDCl3) = 160.2,
54
55 138.8, 135.5, 129.1, 128.4, 126.0, 124.4, 115.2, 55.5.
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 26 of 43

1
2
3 4-(Phenylthio)benzenamine (6af).4e Eluent: petroleum ether/ethyl acetate (50:1). White solid.
4
5
6
Yield c83% (167 mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 7.17 (dd, J1 = 12 Hz, J2 = 8
7
8 Hz, 2H), 7.05 (t, J = 8 Hz, 2H), 7.00 (d, J = 8 Hz, 2H), 6.95 (t, J = 8 Hz, 1H), 6.47 (d, J = 8 Hz,
9
13
10 2H), 3.60 (br, s, 2H). C{1H} NMR (100 MHz, CDCl3) = 147.3, 139.9, 136.3, 129.0, 127.3,
11
12 125.4, 120.1, 116.0.
13
14
15 Methyl 2-(phenylthio)benzoate (6ag).4h Eluent: petroleum ether/ethyl acetate (50:1). Yellow oil.
16
17 Yield c52% (127 mg, for X = I), c39% & d45% (96 mg & 110 mg, for X = Br); e41%
18
19 (101 mg, for X = Cl). 1H NMR (400 MHz, CDCl3): δ = 7.97 (d, J = 7.91 Hz, 1H), 7.56-7.53 (m,
20
21 2H), 7.43-7.41 (m, 3H), 7.23 (t, J = 8.26 Hz, 1H), 7.14 (t, J = 7.90 Hz, 1H), 6.83 (d, J = 8.25 Hz,
22
23
24
1H), 3.95 (s, 3H). 13C{1H} NMR (100 MHz, CDCl3) = 167.1, 143.7, 134.7, 132.8, 132.4, 131.5,
25
26 130.0, 129.2, 127.5, 127.2, 124.7, 52.0.
27
28 2-(phenylthio)phenol (6ah).6h Eluent: petroleum ether/ethyl acetate (50:1). Yellow oil. Yield
29
30 c
57% (115 mg, for X = I), c37% & d43% (75 mg & 87 mg, for X = Br); e40% (81 mg, for X =
31
32
33 Cl). 1H NMR (400 MHz, CDCl3): δ = 7.47-7.45 (m, 1H), 7.33-7.28 (m, 1H), 7.18-7.14 (m, 2H),
34
13
35 7.09-7.05 (m, 1H), 7.02-6.99 (m, 3H), 6.90-6.86 (m, 1H), 6.45 (s, 1H). C{1H} NMR (100
36
37 MHz, CDCl3) = 157.2, 136.9, 135.8, 132.3, 129.2, 126.8, 126.1, 121.3, 116.2, 115.5.
38
39
40
4-(phenylthio)phenol (6ai).4e,5d Eluent: petroleum ether/ethyl acetate(40:1). Yellow oil. Yield
41 c
42 69% (140 mg, for X = I), c47% & d60% (95 mg & 122 mg, for X = Br), e52% (106 mg, for X =
43
44 Cl). 1H NMR (400 MHz, CDCl3): δ = 7.43-7.35 (m, 2H), 7.27-7.13 (m, 5H), 6.86-6.82 (m, 2H),
45
46 4.98 (br, s, 1H). 13
C{1H} NMR (100 MHz, CDCl3) = 156.0, 138.7, 135.7, 129.0, 128.8, 125.8,
47
48 124.5, 116.5.
49
50
51 4-(nitrophenyl)(phenyl)sulfane (6aj).4g,4h Eluent: petroleum ether (50:1). Yellow solid. Yield
52
c
53 55% (128 mg, for X = I), c35% & d45% (81 mg & 104 mg, for X = Br), e25% (58 mg, for X =
54
55 Cl). 1H NMR (400 MHz, CDCl3): δ = 7.94-7.91 (m, 2H), 7.43-7.40 (m, 2H), 7.36-7.32 (m, 3H),
56
57
58
59
60 ACS Paragon Plus Environment
Page 27 of 43 The Journal of Organic Chemistry

1
2
3 7.06-7.03 (m, 2H). 13C{1H} NMR (100 MHz, CDCl3) = 148.5, 145.4, 134.8, 130.4, 130.0, 129.7,
4
5
6
126.6, 124.0.
7
8 4-(chlorophenyl)(phenyl)sulfane (6ak).5d,6c Eluent: petroleum ether. Colorless oil. Yield c59%
9
10 (131 mg, for X = I), c35% & d42% (78 mg & 93 mg, for X = Br), e30% (67 mg, for X = Cl). 1H
11
12
NMR (400 MHz, CDCl3): δ = 7.33-7.31 (m, 6H), 7.28-7.25 (m, 3H). 13C{1H} NMR (100 MHz,
13
14
15 CDCl3) = 135.3, 134.8, 133.1, 132.2, 131.5, 129.5, 129.5, 127.6.
16
17 4-(bromophenyl)(phenyl)sulfane (6al).6c Eluent: petroleum ether. Colorless oil. Yield c61% (162
18
19 mg, for X = I), c43% & d49% (114 mg & 130 mg, for X = Br), e35% (93 mg, for X = Cl). 1H
20
21
22
NMR (400 MHz, CDCl3): δ = 7.36-7.31 (m, 6H), 7.29-7.21 (m, 3H). 13C{1H} NMR (100 MHz,
23
24 CDCl3) = 135.3, 134. 9, 133.3, 132.3, 131.6, 129.5, 129.5, 127.6.
25
26 (Napthalen-6-yl)(Phenyl)sulfane (6am).4h,5d,6c Eluent: petroleum ether/ethyl acetate (50:1).
27
28
Yellow solid. Yield c65% (154 mg, for X = I), c40% & d55% (95 mg & 130 mg, for X = Br),
29
30 e
31 50% (118 mg, for X = Cl). 1H NMR (400 MHz, CDCl3): δ = 7.80-7.63 (m, 4H), 7.40-7.33 (m,
32
33 5H), 7.30-7.15 (m, 3H). 13C{1H} NMR (100 MHz, CDCl3) = 136.0, 133.9, 133.2, 132.4, 131.1,
34
35 130.0, 129.4, 129.0, 128.9, 127.9, 127.6, 127.2, 126.8, 126.4.
36
37
38
2-(phenylthio)pyridine (6an).4h,5d Eluent: petroleum ether/ethyl acetate (10:1). Yellow oil. Yield
39
c
40 55% (103 mg, for X = I), c40% & d45% (75 mg & 85 mg, for X = Br), e15% (28 mg, for X =
41
42 Cl). 1H NMR (400 MHz, CDCl3): δ = 8.33 (d, J = 4 Hz, 1H), 7.51-7.48 (m, 2H), 7.37-7.30 (m,
43
44 4H), 6.90-6.87 (m, 1H), 6.78 (d, J = 8.4 Hz, 1H). 13
C{1H} NMR (100 MHz, CDCl3) = 161.5,
45
46
149.5, 136.8, 135.0, 131.0, 129.6, 129.1, 121.3, 119.9.
47
48
49 2-(phenylthio)benzo[d] thioazole (6ao).4e,5d Eluent: petroleum ether/ethyl acetate (40:1). White
50
51 solid. Yield c53% (129 mg, for X = I), c40% & d45% (98 mg, 110 & for X = Br), e40% (98 mg,
52
53 for X = Cl). 1H NMR (400 MHz, CDCl3): δ = 7.97 (d, J = 13 Hz, 1H), 7.66 (d, J = 9.6 Hz, 2H),
54
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 28 of 43

1
2
3 7.54 (d, J = 9.6 Hz, 1H), 7.42-7.20 (m, 4H), 7.17 (t, J = 5.6 Hz, 1H). 13C{1H} NMR (100 MHz,
4
5
6
CDCl3) = 169.6, 153.9, 135.6, 135.3, 130.1, 129.9, 126.6, 124.7, 121.9, 120.9.
7
8 Benzyl(phenyl)sulfane (6ap).3h Eluent: petroleum ether. Yellow solid. Yield e66% (132 mg, for
9
10 X = I), d60% (120 mg, for X = Br), e47% (94 mg, for X = Cl). 1H NMR (400 MHz, CDCl3): δ =
11
12 13
7.31-7.17 (m, 10H), 4.10 (s, 2H). C{1H} NMR (100 MHz, CDCl3) = 138.5, 137.7, 129.8,
13
14
15 129.4, 129.2, 128.8, 127.7, 127.3, 39.5.
16
17 2-(methoxyphenyl)(phenyl)sulfane (6ba).4h,5d Eluent: petroleum ether/ethyl acetate (50:1).
18
19 Yellow oil. Yield c63% (137 mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 7.32-7.26 (m, 4H),
20
21 13
22
7.26-7.20 (m, 2H), 7.07-7.01 (m, 1H), 6.90-6.85 (m, 2H), 3.71 (s, 3H). C{1H} NMR (100
23
24 MHz, CDCl3) = 157.3, 134.7, 131.2, 130.5, 130.3, 128.1, 127.3, 125.1, 120.9, 109.8, 54.7.
25
26 3, 5-dimethylphenyl(phenyl)sulfane (6ea).4g,4h,5d Eluent: petroleum ether. Colorless oil. Yield
27
28 c
75% (161 mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 7.30-7.28 (m, 2H), 7.25-7.21 (m,
29
30 13
31 2H), 7.18-7.14 (m, 1H), 6.98 (s, 2H), 6.85 (s, 1H), 2.22 (s, 6H). C{1H} NMR (100 MHz,
32
33 CDCl3) = 137.8, 135.4, 133.7, 129.4, 128.1, 125.6, 20.1.
34
35 Ethyl 4-(phenylthio)benzoate (6fa).5i Eluent: petroleum ether. Colorless oil. Yield c81% (210
36
37
38
mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 7.86 (d, J = 4.4 Hz, 2H), 7.47-7.45 (m, 2H),
39
40 7.38-7.35 (m, 3H), 7.21 (d, J = 4 Hz, 2H), 4.31-4.26 (q, J = 7.2 Hz, 2H), 1.34-1.29 (m, 3H).
41
13
42 C{1H} NMR (100 MHz, CDCl3) = 166.1, 144.0, 133.5, 132.5, 130.0, 129.5, 128.6, 127.9,
43
44 127.6, 60.9, 14.3.
45
46
47 1-(2-(phenylthio)phenyl)ethanone (6ga).4h Eluent: petroleum ether/ethyl acetate (50:1). Yellow
48
49 solid. Yield c52% (119 mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 7.82 (d, J = 8.8 Hz, 2H),
50
13
51 7.54-7.50 (m, 2H), 7.44-7.40 (m, 3H), 7.26-7.17 (m, 2H), 6.88 (d, J = 7.6 Hz, 1H). C{1H}
52
53 NMR (100 MHz, CDCl3) = 199.8, 142.9, 135.0, 134.8, 133.2, 132.5, 130.4, 129.7, 129.0, 128.2,
54
55
56
124.4, 29.4.
57
58
59
60 ACS Paragon Plus Environment
Page 29 of 43 The Journal of Organic Chemistry

1
2
3 1-(3-(phenylthio)phenyl)ethanone (6ha).3c Eluent: petroleum ether/ethyl acetate (50:1). Yellow
4
5
6
solid. Yield c70% (160 mg, for X = I), d50% (114 mg, for X = Br). 1H NMR (400 MHz, CDCl3):
7
8 δ = 7.81 (s, 1H), 7.71 (d, J = 7.6 Hz, 1H), 7.38 (d, J = 8.0 Hz, 1H), 7.31-7.17 (m, 6H), 2.46 (s,
9
13
10 3H). C{1H} NMR (100 MHz, CDCl3) = 197.5, 137.9, 137.5, 134.6, 134.4, 131.9, 130.0, 129.5,
11
12 129.4, 127.8, 126.6, 26.7.
13
14
15 1-(4-(phenylthio)phenyl)ethanone (6ia).5d,6h Eluent: petroleum ether/ethyl acetate (30:1). Yellow
16
17 solid. Yield c75% (172 mg, for X = I), d67% (153 mg, for X = Br). 1H NMR (400 MHz, CDCl3):
18
19 δ = 7.80 (d, J = 7.6 Hz, 2H), 7.51-7.47 (m, 2H), 7.41-7.38 (m, 3H), 7.20 (d, J = 5.6 Hz, 2H),
20
21 2.54 (s, 3H). 13
C{1H} NMR (100 MHz, CDCl3) = 197.2, 145.0, 134.6, 134.0, 132.3, 129.8,
22
23
24
129.0, 128.9, 127.6, 26.7.
25
26 4-(phenylthio) benzoic acid (6ja).5d Eluent: petroleum ether. Colorless oil. Yield c82% (189 mg,
27
28 for X = I), d,e65% (150 mg, for X = Br & Cl). 1H NMR (400 MHz, CDCl3): δ = 7.95 (d, J = 8Hz,
29
30 13
2H), 7.35-7.33 (m, 2H), 7.25-7.21 (m, 4H), 7.19-7.15 (m, 2H). C{1H} NMR (100 MHz,
31
32
33 CDCl3) = 171.0, 145.9, 135.0, 131.9, 130.7, 129.0, 127.9, 127.1, 126.5.
34
35 2-(phenylthio)benzenamine (6la).4g Eluent: petroleum ether. Colorless oil. Yield c65% (131 mg,
36
37 for X = I), e57% (115 mg, for X = Cl). 1H NMR (400 MHz, CDCl3): δ = 7.36 (d, J = 7.6Hz, 1H),
38
39
40
7.15-7.10 (m, 3H), 7.02-6.98 (m, 3H), 6.68-6.63 (m, 2H), 4.18 (br, s, 2H). 13C{1H} NMR (100
41
42 MHz, CDCl3) = 148.9, 137.5, 136.8, 131.2, 129.0, 126.5, 125.4, 118.8, 115.4, 114.3.
43
44 2-(nitrophenyl)phenylsulfane (6ma).6h Eluent: petroleum ether/ethyl acetate (50:1). Colorless
45
46
oil. Yield c69% (160 mg, for X = I), e53% (123 mg, for X = Cl). 1H NMR (400 MHz, CDCl3): δ
47
48 13
49 = 7.46-7.40 (m, 1H), 7.24-7.11 (m, 3H), 7.10-7.01 (m, 3H), 6.76-6.69 (m, 2H). C{1H} NMR
50
51 (100 MHz, CDCl3) = 148.9, 137.4, 136. 9, 130.9, 129.3, 126.7, 125.3, 118.9, 114.6.
52
53 4-(fluorophenyl) (phenyl)sulfane (6oa).3h,5d,6h Eluent: petroleum ether(50:1). Colorless oil. Yield
54
55 c
56
75% (154 mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 7.36-7.32 (m, 2H), 7.29-7.20 (m,
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 30 of 43

1
2
3 5H), 7.00-6.98 (m, 2H). 13C{1H} NMR (100 MHz, CDCl3) = 163.6, 161.4, 136.6, 134.2, 134.0,
4
5
6
130.2, 129.9, 129.3, 126.7, 116.4, 116.3.
7
8 2-(phenylthio)benzonitrile (6pa).4h,5d Eluent: petroleum ether/ethyl acetate (30:1). Yellow solid.
9
10 Yield c67% (142 mg, for X = I), d42% (89 mg, for X = Br). 1H NMR (400 MHz, CDCl3): δ =
11
12
7.63 (d, J = 7.80 Hz, 1H), 7.48-7.45 (m, 2H), 7.43-7.36 (m, 4H), 7.26 (t, J = 7.79 Hz, 1H), 7.12
13
14
15 (t, J = 8.24 Hz, 1H), 6.88 (d, J = 8.24 Hz, 1H), 2.65 (s, 3H). 13C{1H} NMR (100 MHz, CDCl3) =
16
17 142.6, 133.7, 133.6, 133.0, 131.4, 130.0, 129.9, 129.0, 126.3, 117.0, 112.9.
18
19 3-(phenylthio)benzonitrile (6qa).5d Eluent: petroleum ether/ethyl acetate (50:2). Colorless oil.
20
21
22
Yield c79% (167 mg, for X = I), d58% (123 mg, for X = Br). 1H NMR (400 MHz, CDCl3): δ =
23
24 7.33-7.18 (m, 9H). 13C{1H} NMR (100 MHz, CDCl3) = 139.9, 133.4, 132.9, 132.2, 131.6, 129.9,
25
26 129.7, 129.5, 128.9, 118.3, 113.3.
27
28
4-(phenylthio)benzonitrile (6ra).5d Eluent: petroleum ether/ethyl acetate (50:1). Yellow solid.
29
30
31 Yield c84% (178 mg, for X = I), d70% (148 mg, for X = Br), e53% (112 mg, for X = Cl). 1H
32
33 NMR (400 MHz, CDCl3): δ = 7.53-7.34 (m, 7H), 7.18-7.14 (m, 2H). 13C{1H} NMR (100 MHz,
34
35 CDCl3) = 145.8, 134.5, 132.4, 130.8, 129.9, 129.4, 127.3, 118.8, 108.6.
36
37
38
(Napthalen-5-yl)(Phenyl)sulfane (6sa).4g,4h,5d Eluent: petroleum ether/ethyl acetate
39
40 (50:1).Yellow solid. Yield c61% (145 mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 8.38-8.35
41
42 (m, 1H), 7.92-7.74 (m, 2H), 7.63 (dd, J = 6.4 Hz, 1H), 7.47-7.41 (m, 2H), 7.36 (dd, J = 7.6 Hz,
43
44 1H), 7.20-7.05 (m, 5H). 13C{1H} NMR (100 MHz, CDCl3) = 137.0, 134.4, 133.7, 132.7, 131.3,
45
46
129.4, 129.2, 129.1, 128.7, 127.1, 126.6, 126.3, 126.0, 125.8.
47
48
49 (1-dihydropyren-3-yl)(phenyl)sulfane (6ua).4e Eluent: petroleum ether/ethyl acetate (30:1).
50
51 Light yellow solid. Yield c65% (203 mg, for X = I). 1H NMR (CDCl3, 400 MHz): δ = 8.68 (d, J =
52
53 9.2 Hz, 1H), 8.25-8.20 (m, 2H), 8.17 (d, J = 2 Hz, 1H), 8.10-8.01 (m, 4H), 7.93 (d, J = 3.6 Hz,
54
55
56
1H), 7.39-7.36 (m, 2H), 7.00-6.83 (m, 2H). 13C{1H} NMR (CDCl3, 100 MHz) = 144.5, 138.2,
57
58
59
60 ACS Paragon Plus Environment
Page 31 of 43 The Journal of Organic Chemistry

1
2
3 132.6, 132.5, 131.7, 131.3, 131.0, 130.7, 129.4, 129.2, 128.8, 128.7, 128.3, 128.0, 127.3, 127.3,
4
5
6
126.4, 126.2, 125.8, 125.8, 125.5, 125.4, 125.0, 124.5, 115.0, 114.5.
7
8 2-(phenylthio)thiophene (6va).3h,4g Eluent: petroleum ether/ethyl acetate (30:1). Colorless oil.
9
10 Yield c60% (116 mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 7.49 (d, J = 3.6 Hz, 2H), 7.46-
11
12 13
7.37 (m, 2H), 7.35-7.18 (m, 3H), 7.08 (d, J = 8 Hz, 1H). C{1H} NMR (100 MHz, CDCl3) =
13
14
15 138.9, 136.6, 131.7, 131.5, 129.5, 128.8, 127.8, 126.5.
16
17 3-(phenylthio)pyridine (6wa).5d,3h Eluent: petroleum ether/ethyl acetate (50:1). Yellow oil. Yield
18
19 c
65% (122 mg, for X = I). 1H NMR (400 MHz, CDCl3): δ = 8.50 (d, J = 1.8 Hz, 1H), 8.41 (dd, J
20
21
22
= 4.8, 1.2 Hz, 1H), 7.53 – 7.49 (m, 1H), 7.34 – 7.29 (m, 2H), 7.19 – 7.13 (m, 3H), 2.35 (s, 3H)
23
13
24 C{1H} NMR (100 MHz, CDCl3) = 148.5, 136.8, 134.6, 129.8, 129.7, 128.7, 128.1, 125.6,
25
26 121.5.
27
28
Butyl (phenyl) sulfane (6xa).6c Eluent: petroleum ether. Yellow oil. Yield f49% (82 mg, for X =
29
30
31 I). 1H NMR (400 MHz, CDCl3): δ = 7.25-7.16 (m, 4H), 7.09-7.04 (m, 1H), 2.83 (t, 2H, J = 7.2
32
33 Hz), 1.58-1.51(m, 2H), 1.41-1.33 (m, 2H), 0.85-0.82 (m, 3H). 13C{1H} NMR (100 MHz, CDCl3)
34
35 = 137.0, 128.8, 128.8, 125.6, 33.2, 31.2, 22.0, 13.7.
36
37
38
6-((4-Methylphenyl)thio)-3,4-dihydro-2H-benzo[e][1,2,4]thiadiazine-7-sulfonamide1,1-dioxide
39
40 (6a ).4e Eluent: petroleum ether/ethyl acetate (40:1). White solid. Yield 72% (278 mg, for X =
41
42 Cl). 1H NMR (400 MHz, d6-DMSO) = 7.92 (s, 1H), 7.73 (t, J = 7.6 Hz, 2H), 7.52 (dd, J = 6.4
43
44 Hz, J = 4.8 Hz, 2H), 7.37 (s, 2H), 7.11 (dd, J = 5.2 Hz, 3.2 Hz, 2H), 6.14 (s, 1H), 4.67 (dd, J =
45
46
8.4 Hz, 3.2 Hz, 2H), 2.39 (s, 3H).
47
48
49 6-(thio)-3,4-dihydro-2H-benzo[e][1,2,4]thiadiazine-7-sulfonamide 1,1-dioxide (6b).4e Eluent:
50
51 petroleum ether/ethyl acetate (40:5). White solid. Yield 69% (257 mg, for X = Cl). 1H NMR (400
52
53 MHz, d6-DMSO) = 7.92 (s, 1H), 7.72 (t, J = 7.6 Hz, 2H), 7.56 (d, J = 6.8 Hz, 2H), 7.36 (s, 2H),
54
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 32 of 43

1
2
3 7.11 (d, J = 4.8 Hz, 2H), 6.14 (s, 1H), 4.58 (dd, J = 5.6 Hz, 5.2 Hz, 2H). 13C{1H} NMR (100
4
5
6
MHz, d6-DMSO) = 151.3, 145.2, 144.6, 137.9, 128.0, 125.2, 121.1, 117.3, 116.3, 113.5, 61.0.
7
8 6-((4-Methoxyphenyl)thio)-3,4-dihydro-2H-benzo[e][1,2,4]thiadiazine-7-sulfonamide1,1-
9
10 dioxide (6c).4e Eluent: petroleum ether/ethyl acetate (40:1). White solid. Yield 68% (273 mg, for
11
12
X = Cl). 1H NMR (400 MHz, CDCl3) = 7.92 (s, 1H), 7.00 (t, J = 7.6 Hz, 2H), 7.51 (dd, J = 7.6
13
14
15 Hz, 2.0 Hz, 2H), 7.37 (s, 2H), 7.12 (dd, J = 8.4 Hz, 3.2 Hz, 2H), 6.12 (s, 1H), 4.67 (dd, J = 8.4
16
17 Hz, 3.7 Hz, 2H), 3.84 (s, 3H). 13C{1H} NMR (100 MHz, CDCl3) = 161.0, 146.0, 144.5, 138.0,
18
19 127.7, 125.5, 120.4, 116.6, 115.3, 113.6, 56.0, 54.7.
20
21
22
Isopropyl 2-methyl-2-(4-(4-(p-tolylthio)benzoyl)phenoxy)propanoate (6d).4e Eluent: petroleum
23
24 ether/ethyl acetate (40:5). White solid. Yield 52% (242 mg, for X = Cl). 1H NMR (CDCl3, 400
25
26 MHz): δ = 7.74-7.69 (m, 2H), 7.64-7.60 (m, 2H), 7.44-7.40 (m, 2H), 7.21-7.16 (m, 2H), 7.24-
27
28 7.16 (m, 2H), 6.86-6.83 (m, 2H), 5.12-5.05 (m, 1H), 3.84 (s, 3H), 1.64 (s, 6H), 1.20 (s, 6H).
29
30 13
C{1H} NMR (100 MHz, CDCl3) = 194.5, 173.1, 160.5, 158.9, 145.5, 130.7, 130.4, 125.5,
31
32
33 121.1, 117.2, 115.0, 79.4, 69.8, 54.8, 25.8, 21.5.
34
35 4-((4-Methoxyphenyl)thio)-N-(2-morpholinoethyl)benzamide (6e).4e Eluent:
36
37 dicholoromethane/methanol (10:1). White solid. Yield 61% (228 mg, for X = Cl). 1H NMR (400
38
39
40
MHz, CDCl3): δ = 7.63 (d, J = 3.6 Hz, 2H), 7.48 (d, J = 3.6 Hz, 2H), 7.17 (d, J = 6.0 Hz, 2H),
41
42 6.97 (d, J = 6.8 Hz, 2H), 6.73 (s, 1H), 3.84 (s, 3H), 3.74 (t, J = 2.8 Hz, 4H), 3.55 (d, J = 7.6 Hz,
43
44 2H), 2.65 (t, J = 5.2 Hz, 6.8 Hz, 3H), 2.66 (t, J =5.2 Hz, 6.8 Hz, 4H). 13C{1H} NMR (100 MHz,
45
46 CDCl3) = 166.5, 160.8, 144.5, 136.7, 131.8, 126.9, 122.5, 115.2, 66.8, 56.9, 53.2, 36.0.
47
48
49 2-(5-Methoxy-1-(4-((4-methoxyphenyl)thio)benzoyl)-2-methyl-1H-indol-3-yl)acetic acid (6f).4e
50
51 Eluent: dicholoromethane/methanol (10:1). Yellow solid. Yield 68% (314 mg, for X = Cl). 1H
52
53 NMR (CDCl3, 400 MHz): δ = 7.56-7.48 (m, 4H), 7.12 (dd, J = 6.0 Hz, 4.4 Hz, 2H), 6.98-6.90
54
55 (m, 4H), 6.68 (dd, J = 6.8 Hz, 4.4 Hz, 1H), 3.85 (s, 3H), 3.81 (s, 3H), 3.67 (s, 2H), 2.37 (s, 3H).
56
57
58
59
60 ACS Paragon Plus Environment
Page 33 of 43 The Journal of Organic Chemistry

1
2
3 13
C{1H} NMR (100 MHz, CDCl3) = 176.9, 168.8, 160.8, 155.7, 147.9, 136.9, 136.6, 131.6,
4
5
6
131.2, 130.9, 130.0, 130.0, 125.8, 121.2, 15.5, 115.0, 111.6, 111.2, 101.1, 55.5, 30.0, 21.6, 13.0.
7
8 1-(4-((2-Chlorophenyl)thio)phenyl)ethan-1-one (6g).4e Eluent: petroleum ether/ethyl acetate
9
10 (40:1). White solid. Yield 50% (131 mg, for X = Br). 1H NMR (CDCl3, 400 MHz): δ = 7.89-7.85
11
12
(m, 2H), 7.51-7.48 (m, 1H), 7.43-7.39 (m, 1H), 7.34-7.29 (m, 1H), 7.29-7.22 (m, 3H), 2.23 (s,
13
14
15 3H). 13C{1H} NMR (100 MHz, CDCl3) = 197.0, 142.0, 137.2, 135.2, 134.6, 132.3, 130.0, 129.8,
16
17 129.2, 128.8, 128.7, 127.3, 26.2.
18
19 2-(Allyloxy)phenylphenyl sulfide (6ya).3c Eluent: petroleum ether/ethyl acetate (50:2). Colorless
20
21
22
oil. Yield c52% (126 mg, X = I). 1HNMR(400 MHz, CDCl3): δ = 7.42-7.34 (m, 2H), 7.34-7.28
23
24 (m, 2H), 7.28-7.23 (m, 1H), 7.21 (ddd, J = 1.6 Hz, 7.6 Hz, 10.0 Hz, 1H), 7.09-7.01 (m, 1H),
25
26 6.94-6.85 (m, 2H), 6.04-5.90 (m, 1H), 5.38-5.36 (m, 1H), 5.24-5.20 (m,1H), 4.58 (td, J = 5.2 Hz,
27
28 6Hz, 2H ). 13C{1H} NMR (100 MHz, CDCl3) =156.1, 134.7, 133.0, 131.9, 129.2, 129.1, 128.1,
29
30
127.2, 125.1, 121.3, 117.5, 112.6, 69.0.
31
32
33 ASSOCIATED CONTENT
34
35 Supporting Information
36
37
38 Reaction optimization, EPR studies, 1H and 13C NMR spectral data, spin density plots, optimized
39
40 coordinates are provided. This material are available free of charge via the Internet at
41
42 http://pubs.acs.org.
43
44
45 AUTHOR INFORMATION
46
47 Corresponding Author
48
49
50 *E-mail: ndpaul@gmail.com; ndpaul2014@chem.iiests.ac.in
51
52 ORCID:
53
54
55
Nanda Dulal Paul: 0000-0002-8872-1413
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 34 of 43

1
2
3 Notes
4
5
6
The authors declare no competing financial interest.
7
8 ACKNOWLEDGMENT
9
10
The research was supported by DST, Govt. of India (Project: YSS/2015/001552). We thank Prof.
11
12
13 Bas de Bruin, University of Amsterdam for helpful discussion and for EPR experiments.
14
15
16
17
18
19
20
21
22
23
24
25
26
27
REFERENCES
28
29
30 (1) (a) Rakitin, O. A.; In Science of Synthesis; Georg Thieme Verlag: New York, 2007; 31a,
31
32 975−1000. (b) Feng, M.; Tang, B.; Liang, S.; Jiang, X. Sulfur Containing Scaffolds in Drugs:
33
34 Synthesis and Application in Medicinal Chemistry. Curr. Top. Med. Chem. 2016, 16, 1200–1216.
35
36
37 (c) Boyd, D. A. Sulfur and Its Role In Modern Materials Science. Angew. Chem. Int. Ed. 2016,
38
39 55, 15486–15502. (d) Evano, G.; Theunissen, C.; Pradal, A. Impact of Copper-Catalyzed Cross-
40
41 Coupling Reactions in Natural Product Synthesis: the Emergence of New Retrosynthetic
42
43 Paradigms. Nat. Prod. Rep. 2013, 30, 1467–1489. (e) Sheehan, J. C.; H.-Logan, K. R. The Total
44
45
Synthesis of Penicillin V. J. Am. Chem. Soc. 1959, 81, 3089–3094. (f) Raghavan, S.; Sridhar, B.
46
47
48 Asymmetric Synthesis of the Potent HIV-Protease Inhibitor, Nelfinavir. J. Org. Chem. 2010, 75,
49
50 498–501. (g) Ilardi, E. A.; Vitaku, E.; Njardarson, J. T. Data-Mining for Sulfur and Fluorine: An
51
52 Evaluation of Pharmaceuticals To Reveal Opportunities for Drug Design and Discovery. J. Med.
53
54 Chem. 2014, 57, 2832–2842.
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 35 of 43 The Journal of Organic Chemistry

1
2
3 (2) (a) Bhunia, S .; Pawar, G. G.; Kumar, S. V. ; Jiang, Y. ; Ma, D. Selected Copper-Based
4
5
6
Reactions for C-N, C-O, C-S, and C-C Bond Formation. Angew. Chem. Int. Ed. 2017, 56,
7
8 16136–16179. (b) Desnoyer, A. N.; Love, J. A. Recent Advances in Well-defined, Late Transition
9
10 Metal Complexes that Make and/or Break C–N, C–O and C–S Bonds. Chem. Soc. Rev. 2017, 46,
11
12 197–238. (c) Shen, C.; Zhang, P.; Sun, Q.; Bai, S.; Hor, T. S. A.; Liu, X. Recent advances in C–S
13
14
bond formation via C–H bond functionalization and decarboxylation. Chem. Soc. Rev. 2015, 44,
15
16
17 291–314. (d) Lee, C.-F.; Liu, Y.-C.; Singh, S. B. Transition-Metal-Catalyzed C–S Bond Coupling
18
19 Reaction. Chem. Asian J. 2014, 9, 706–722. (e) Timpa, S. D.; Pell, C. J.; Ozerov, O. V. A Well-
20
21 Defined (POCOP)Rh Catalyst for the Coupling of Aryl Halides with Thiols. J. Am. Chem. Soc.
22
23 2014, 136, 14772–14779. (f) Beletskaya, I. P.; Ananikov, V. P. Transition-Metal-Catalyzed C–S,
24
25
C–Se, and C–Te Bond Formation via Cross-Coupling and Atom-Economic Addition Reactions.
26
27
28 Chem. Rev. 2011, 111, 1596–1636. (g) Alvaro, E.; Hartwig, J. F. Resting State and Elementary
29
30 Steps of the Coupling of Aryl Halides with Thiols Catalyzed by Alkylbisphosphine Complexes
31
32 of Palladium. J. Am. Chem. Soc. 2009, 131, 7858–7868. (h) Hartwig, J. F. Evolution of a Fourth
33
34 Generation Catalyst for the Amination and Thioetherification of Aryl Halides. Acc. Chem. Res.
35
36
37
2008, 41, 1534–1544.
38
39 (3) (a) Yugandar, S.; Konda, S.; Ila, H. Synthesis of Substituted Benzo[b]thiophenes via
40
41 Sequential One-Pot,Copper-Catalyzed Intermolecular C−S Bond Formation and Palladium-
42
43
Catalyzed Intramolecular Arene−Alkene Coupling of Bis(het)aryl/alkyl-1,3-monothiodiketones
44
45
46 and o–Bromoiodoarenes. Org. Lett. 2017, 19, 1512–1515. (b) Mao, J.; Jia, T.; Frensch, G.;
47
48 Walsh, P. J. Palladium-Catalyzed Debenzylative Cross-Coupling of Aryl Benzyl Sulfides with
49
50 Aryl Bromides: Synthesis of Diaryl Sulfides. Org. Lett. 2014, 16, 5304–5307. (c) Fernández-
51
52 Rodríguez, M. A.; Hartwig, J. F. One-Pot Synthesis of Unsymmetrical Diaryl Thioethers by
53
54
55
Palladium-Catalyzed Coupling of Two Aryl Bromides and a Thiol Surrogate. Chem. Eur. J. 2010,
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 36 of 43

1
2
3 16, 2355–2359. (d) Hartwig, J. F. Carbon–Heteroatom Bond Formation Catalysed by
4
5
6
Organometallic Complexes. Nature. 2008, 455, 314–322. (e) Lee, J-Y.; Lee, P. H. Palladium-
7
8 Catalyzed Carbon-Sulfur Cross-Coupling Reactions with Indium Tri(organothiolate) and Its
9
10 Application to Sequential One-Pot Processes. J. Org. Chem. 2008, 73, 7413–7416. (f)
11
12 Fernández-Rodríguez, M. A.; Shen, Q.; Hartwig, J. F. A General and Long-Lived Catalyst for the
13
14
Palladium-Catalyzed Coupling of Aryl Halides with Thiols. J. Am. Chem. Soc. 2006, 128, 2180–
15
16
17 2181. (g) Itoh, T.; Mase, T. A General Palladium-Catalyzed Coupling of Aryl Bromides/Triflates
18
19 and Thiols. Org. Lett. 2004, 6, 4587–4590. (h) Kondo, T.; Mitsudo, T. Metal-Catalyzed Carbon-
20
21 Sulfur Bond Formation. Chem. Rev. 2000, 100, 3205–3220. (i) Fu, C.-F.; Liu, Yi-H.; Peng, S.-
22
23 M.; Liu, S.-T. C–S Bond Formation Catalyzed by N-Heterocylic Carbene Palladium Phosphine
24
25
Complexes. Tetrahedron 2010, 66, 2119–2122.
26
27
28 (4) (a) Wimmer, A.; König, B. Photocatalytic formation of carbon–sulfur bonds. Beilstein J. Org.
29
30 Chem. 2018, 14, 54–83 and reference therein. (b) Shaw, M. H.; Twilton, J.; MacMillan, D. W. C.
31
32 Photoredox Catalysis in Organic Chemistry. J. Org. Chem. 2016, 81, 6898–6926 and references
33
34
35 therein. (c) Jiang, M.; Li, H.; Yang, H.; Fu, H. Room-Temperature Arylation of Thiols:
36
37 Breakthrough with Aryl Chlorides. Angew. Chem. Int. Ed. 2017, 56, 874–879. (d) Oderinde, M.
38
39 S.; Frenette, M.; Robbins, D. W.; Aquila, B.; Johannes, J. W. Photoredox Mediated Nickel
40
41 Catalyzed Cross-Coupling of Thiols With Aryl and Heteroaryl Iodides via Thiyl Radicals. J. Am.
42
43
Chem. Soc. 2016, 138, 1760–1763. (e) Johnson, M.; Hannoun, K.; Tan, Y.; Fu, G.; Peters, J. C.; A
44
45
46 Mechanistic Investigation of the Photoinduced, Copper-Mediated Cross-Coupling of An Aryl
47
48 Thiol with An Aryl Halide. Chem. Sci. 2016, 7, 4091–4100. (f) Uyeda,C.; Tan, Y.; Fu, G. C.;
49
50 Peters, J. C. A New Family of Nucleophiles for Photoinduced, Copper-Catalyzed Cross-
51
52 Couplings via Single-Electron Transfer: Reactions of Thiols with Aryl Halides Under Mild
53
54
55
Conditions (0 °C). J. Am. Chem. Soc. 2013, 135, 9548–9552.
56
57
58
59
60 ACS Paragon Plus Environment
Page 37 of 43 The Journal of Organic Chemistry

1
2
3 (5) (a) Wang, M.; Wei, J.; Fana, Q.; Jiang, X. Cu(II)-Catalyzed Sulfide Construction: both Aryl
4
5
6
Groups Utilization of Intermolecular and Intramolecular Diaryliodonium Salt. Chem. Commun.
7
8 2017, 53, 2918–2921. (b) Soria-Castro, S. M.; Andrada, D. M.; Caminos, D. A.; Argüello, J.
9
10 E.; Robert, M.; Peñéñory, A. B. Mechanistic Insight into the Cu-Catalyzed C−S Cross-Coupling
11
12 of Thioacetate with Aryl Halides: A Joint Experimental−Computational Study. J. Org.
13
14
Chem. 2017, 82, 11464–11473. (c) Font, M.; Parella, T.; Costas, M.; Ribas, X. Catalytic C−S,
15
16
17 C−Se, and C−P Cross-Coupling Reactions Mediated by a CuI/CuIII Redox Cycle.
18
19 Organometallics 2012, 31, 7976–7982. (d) Xu, H.-J.; Zhao, Y. -Q.; Feng, T.; Feng, Y.-S.
20
21 Chan−Lam-Type S-Arylation of Thiols with Boronic Acids at Room Temperature. J. Org. Chem.
22
23 2012, 77, 2878–2884. (e) Ranjit, S.; Lee, R.; Heryadi, D.; Shen, C.; Wu, Ji’En.; Zhang, P.;
24
25
Huang, K.-W.; Liu, X. Copper-Mediated C−H Activation/C−S Cross-Coupling of Heterocycles
26
27
28 with Thiols. J. Org. Chem. 2011, 76, 8999–9007. (f) Prasad, D. J. C.; Sekar, G.; Cu-Catalyzed
29
30 One-Pot Synthesis of Unsymmetrical Diaryl Thioethers by Coupling of Aryl Halides Using a
31
32 Thiol Precursor. Org. Lett. 2011, 13, 1008–1011. (g) Herrero, M. T.; SanMartin, R.; Domínguez,
33
34 E. Copper(I)-Catalyzed S-Arylation of Thiols with Activated Aryl Chlorides on Water.
35
36
37
Tetrahedron 2009, 65, 1500–1503. (h) Sperotto, E.; van Klink, G. P. M.; de Vries, J. G.; van
38
39 Koten, G. Ligand-Free Copper-Catalyzed C-S Coupling of Aryl Iodides and Thiols. J. Org.
40
41 Chem. 2008, 73, 5625–5628. (i) Wong, Y-C.; Jayanth, T. T.; Cheng, C-H. Cobalt Catalyzed
42
43 Aryl−Sulfur Bond Formation. Org. Lett. 2006, 8, 5613–5616.
44
45
46 (6) (a) Jones, K. D.; Power, D. J.; Bierer, D.; Gericke, K. M.; Stewart, S. G. Nickel
47
48 Phosphite/Phosphine-Catalyzed C–S Cross-Coupling of Aryl Chlorides and Thiols. Org. Lett.
49
50 2018, 20, 208–211. (b) Gogoi, P.; Hazarika, S.; Sarma, M. J.; Sarma, K.; Barman, P.; Nickel–
51
52 Schiff Base Complex Catalyzed C–S Cross-Coupling of Thiols with Organic Chlorides.
53
54
55
Tetrahedron 2014, 70, 7484–7489. (c) Wellala, N. P. N.; Guan, H. A Diphenyl Ether Derived
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 38 of 43

1
2
3 Bidentate Secondary Phosphine Oxide As a Preligand for Nickel catalyzed C–S Cross-Coupling
4
5
6
Reactions. Org. Biomol. Chem. 2015, 13, 10802–10807. (d) Martin, A. R.; Nelson, D. J.;
7
8 Meiries, S.; Slawin, A. M. Z.; Nolan, S. P. Efficient C–N and C–S Bond Formation Using the
9
10 Highly Active [Ni(allyl)Cl(IPr*OMe)] Precatalyst. Eur. J. Org. Chem. 2014, 3127–3131. (e)
11
12 Zhang, J.; Medley, C. M.; Krause, J. A.; Guan, H. Mechanistic Insights into C-S Cross-Coupling
13
14
Reactions Catalyzed by Nickel Bis(phosphinite) Pincer Complexes. Organometallics. 2010, 29,
15
16
17 6393–6401. (f) Zhang, Y.; Ngeow, K. C.; Ying, J. Y. The First N-Heterocyclic Carbene-Based
18
19 Nickel Catalyst for C-S Coupling. Org. Lett. 2007, 9, 3495–3498. (g) Venkanna, G. T.; Arman,
20
21 H. D.; Tonzetich, Z. J. Catalytic C−S Cross-Coupling Reactions Employing Ni Complexes of
22
23 Pyrrole-Based Pincer Ligands. ACS Catal. 2014, 4, 2941–2950. (h) Taniguchi, N. Alkyl- or
24
25
Arylthiolation of Aryl Iodide via Cleavage of the S-S Bond of Disulfide Compound by Nickel
26
27
28 Catalyst and Zinc. J. Org. Chem. 2004, 69, 6904-6906. (i) Baldovino-Pantaleón, O.; Hernández-
29
30 Ortega, S.; Morales-Morales, D. Alkyl- and Arylthiolation of Aryl Halides Catalyzed by
31
32 Fluorinated Bis-Imino-Nickel NNN Pincer Complexes [NiCl2{C5H3N-2,6-(CHNArf)2}] Adv.
33
34 Synth. Catal. 2006, 348, 236–242.
35
36
37 (7) (a) de Bruin, B.; Gualco, P.; Paul, N. D. Ligand Design in Metal Chemistry: Reactivity and
38
39 Catalysis; John Wiley & Sons, Ltd.: Chichester, UK, 2016; pp 176−204. (b) Special Issue:
40
41 Cooperative and Redox Non-Innocent Ligands in Directing Organometallic Reactivity. Eur. J.
42
43
Inorg. Chem. 2012, 340. DOI: 10.1002/ejic.201290001. (c) Forum on Redox-ActiveLigands.
44
45
46 Inorg. Chem. 2011, 50, 9737. DOI: 10.1021/ic201881k. (d) Dzik, W. I.; van der Vlugt, J. I.;
47
48 Reek, J. N. H.; de Bruin, B. Ligands that Store and Release Electrons during Catalysis. Angew.
49
50 Chem. Int. Ed. 2011, 50, 3356–3358. (e) Lyaskovskyy, V.; de Bruin, B. Redox Non-Innocent
51
52 Ligands: Versatile New Tools to Control Catalytic Reactions. ACS Catal. 2012, 2, 270–279. (f)
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 39 of 43 The Journal of Organic Chemistry

1
2
3 Chirik, P. J.; Wieghardt, K. Radical Ligands Confer Nobility on Base-Metal Catalysts. Science.
4
5
6
2010, 327, 794–795.
7
8 (8) (a) van der Vlugt, J. I. Radical-type Reactivity and Catalysis via Single-Electron Transfer to
9
10 or from Redox-Active Ligands. Chem. Eur. J. 2018, doi.org/10.1002/chem.201802606. (b) Du,
11
12
H-Y.; Chen, S-C.; Su, X-J.; Jiao, L.; Zhang, M-T. Redox-Active Ligand Assisted Multi electron
13
14
15 Catalysis: A Case of CoIII Complex as Water Oxidation Catalyst. J. Am. Chem. Soc. 2018, 140,
16
17 1557−1565. (c) Pramanick, R.; Bhattacharjee, R.; Sengupta, D.; Datta, A.; Goswami, S. An
18
19 Azoaromatic Ligand as Four Electron Four Proton Reservoir: Catalytic Dehydrogenation of
20
21 Alcohols by Its Zinc(II) Complex. Inorg. Chem. 2018, 57, 6816–6824. (d) Broere, D. J. L.;
22
23
24
Plessius, R.; van der Vlugt, J. I. New Avenues for Ligand-Mediated Processes – Expanding
25
26 Metal Reactivity by The Use of Redox-Active Catechol, o-Aminophenol and o-
27
28 Phenylenediamine Ligands. Chem. Soc. Rev. 2015, 44, 6886–6915. (e) Chirik, P. J. Iron- and
29
30 Cobalt-Catalyzed Alkene Hydrogenation: Catalysis with Both Redox-Active and Strong Field
31
32 Ligands. Acc. Chem. Res. 2015, 48, 1687–1695. (f) Schmidt, V. A.; Hoyt, J. M.; Margulieux, G.
33
34
35 W.; Chirik, P. J. Cobalt-Catalyzed [2π + 2π] Cycloadditions of Alkenes: Scope, Mechanism, and
36
37 Elucidation of Electronic Structure of Catalytic Intermediates. J. Am. Chem. Soc. 2015, 137,
38
39 7903–7914. (g) Broere, D. L. J.; de Bruin, B.; Reek, J. N. H.; Lutz, M.; Dechert, S.; van der
40
41 Vlugt, J. I. Intramolecular Redox-Active Ligand-to-Substrate Single-Electron Transfer: Radical
42
43
Reactivity with a Palladium(II) Complex. J. Am. Chem. Soc. 2014, 136, 11574–11577. (h) Luca,
44
45
46 O. R.; Crabtree, R. H. Redox-Active Ligands in Catalysis. Chem. Soc. Rev. 2013, 42, 1440–
47
48 1459. (i) Sengupta, D.; Ghosh, P.; Chatterjee, T.; Datta, H.; Paul, N. D.; Goswami, S. Ligand-
49
50 Centered Redox in Nickel(II) Complexes of 2‑(Arylazo)pyridine and Isolation of 2‑Pyridyl-
51
52
53 Substituted Triaryl Hydrazines via Catalytic N–Arylation of Azo-Function. Inorg.
54
55 Chem. 2014, 53, 12002–12013. (j) Sinha, S.; Sikari, R.; Sinha, V.;, Jash, U.; Das, S.; Brandão, P.;
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 40 of 43

1
2
3 Demeshko, S.; Meyer, F.; de Bruin, B.; Paul, N. D. Iron-Catalyzed/Mediated C-N Bond
4
5
6
Formation: Competition between Substrate Amination and Ligand Amination. Inorg Chem. 2019,
7
8 58, 1935–1948.
9
10 (9) (a) Stiefel, E. I.; Waters, J. H.; Billig, E.; Gray, H. B. The Myth of Nickel(III) and Nickel(IV)
11
12
13 in Planar Complexes. J. Am. Chem. Soc. 1965, 87, 3016–3017. (b) Chaudhuri, P.; Verani, C. N.;
14
15 Bill, E.; Bothe, E.; Weyhermüller, T.; Wieghardt, K. Electronic Structure of Bis(o-
16
17 iminobenzosemiquinonato)metal Complexes (Cu, Ni, Pd). The Art of Establishing Physical
18
19 Oxidation States in Transition-Metal Complexes Containing Radical Ligands. J. Am. Chem. Soc.
20
21
22
2001, 123, 2213–2223. (c) Sikari, R.; Sinha, S.; Jash, U.; Das, S.; Brandão, P.; de Bruin, B.; Paul,
23
24 N. D. Deprotonation Induced Ligand Oxidation in a NiII Complex of a Redox Noninnocent N1-
25
26 (2-Aminophenyl)benzene-1,2-diamine and Its Use in Catalytic Alcohol Oxidation. Inorg. Chem.
27
28 2016, 55, 6114–6123. (d) Chakraborty, G.; Sikari, R.; Das, S.; Mondal, R.; Sinha, S.; Banerjee,
29
30
S.; Paul, N. D. Dehydrogenative Synthesis of Quinolines, 2-Aminoquinolines, and Quinazolines
31
32
33 Using Singlet Diradical Ni(II)-Catalysts. J. Org. Chem. DOI: 10.1021/acs.joc.8b03070.
34
35 (10) (a) Barham, J. P.; Coulthard, G.; Emery, K. J.; Doni, E.; Cumine, F.; Nocera, G.; John, M. P.;
36
37 Berlouis, L. E. A.; McGuire, T.; Tuttle, T.; Murphy, J. A. KOtBu: A Privileged Reagent for
38
39
40
Electron Transfer Reactions? J. Am. Chem. Soc. 2016, 138, 7402–7410. (b) Yi, H.; Jutand, A.;
41
42 Lei, A. Evidence for the Interaction Between tBuOK and 1,10-Phenanthroline to form the 1,10-
43
44 Phenanthroline Radical Anion: A Key Step for the Activation of Aryl Bromides by Electron
45
46 Transfer. Chem. Commun. 2015, 51, 545–548. (c) Das, S.; Sinha, S.; Jash, U.; Sikari, R.; Saha,
47
48 A.; Barman, S.; Brandaõ, P.; Paul. N. D. Redox-Induced Interconversion and Ligand-Centered
49
50
51 Hemilability in NiII–Complexes of Redox-Noninnocent Azo-Aromatic Pincers. Inorg.
52
53 Chem. 2018, 57, 5830–5841. (d) Sinha, S.; Das, S.; Sikari, R.; Parua, S.; Brandão, P.; Demeshko,
54
55 S.; Meyer, F.; Paul, N. D. Redox Noninnocent Azo-Aromatic Pincers and Their Iron Complexes.
56
57
58
59
60 ACS Paragon Plus Environment
Page 41 of 43 The Journal of Organic Chemistry

1
2
3 Isolation, Characterization, and Catalytic Alcohol Oxidation. Inorg. Chem. 2017, 56, 14084–
4
5
6
14100.
7
8 (11) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J.
9
10 R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A.
11
12
V.; Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. V.;
13
14
15 Izmaylov, A. F.; Sonnenberg, J. L.; Williams-Young, D.; Ding, F.; Lipparini, F.; Egidi, F.;
16
17 Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.;
18
19 Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.;
20
21 Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.;
22
23
24
Montgomery, J. A.; Peralta, Jr., J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.;
25
26 Kudin, K. N.; Staroverov, V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.;
27
28 Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.;
29
30 Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J.
31
32 B.; Fox D. J.; Gaussian 16, revision B.01; Gaussian, Inc.: Wallingford, CT, 2016.
33
34
35
(12) (a) Becke A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J.
36
37
38 Chem. Phys, 1993, 98, 5648–5652. (b) Becke, A. D. Density-Functional Exchange-Energy
39
40 Approximation with Correct Asymptotic Behavior. Phys. Rev. A, 1988, 38, 3098–3100.
41
42 (13) Lee, C.; Yang, W.; Parr, W. R. G. Development of the Colle-Salvetti Correlation-Energy
43
44
45
Formula into A Functional of the Electron Density. Phys. Rev. B: Condens. Matter, 1988, 37,
46
47 785–789.
48
49 (14) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate ab Initio
50
51
Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu.
52
53
54 J. Chem. Phys. 2010, 132, 154104 - 154119.
55
56
57
58
59
60 ACS Paragon Plus Environment
The Journal of Organic Chemistry Page 42 of 43

1
2
3 (15) Hay, P. J.; Wadt, W. R. Ab Initio Effective Core Potentials Formolecular Calculations.
4
5
6
Potentials for the Transition Metal Atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270−283.
7
8 (16) Petersson, G. A.; Bennett,A.; Tensfeldt, T. G.; Al-Laham, M. A.; Shirley, W. A.; Mantzaris,
9
10 J. Acomplete Basis Set Model Chemistry. I. The Total Energies of Closed Shell Atoms and
11
12
Hydrides of the First-Row Elements. J. Chem. Phys.1988, 89, 2193−2218.
13
14
15 (17) Petersson, G. A.; Al-Laham, M. A. A Complete Basis Set Model Chemistry. II. Open-Shell
16
17 Systems and the Total Energiesof the First-Row Atoms. J. Chem. Phys.1991, 94, 6081−6090.
18
19
20 (18) Noodleman, L. Valence Bond Description of Antiferromagnetic Coupling in Transition
21
22 Metal Dimers. J. Chem. Phys. 1981, 74, 5737−5743. (b) Noodleman, L.; Davidson, E. R. Ligand
23
24 Spin Polarization and Antiferromagnetic Coupling in Transition Metal Dimers.Chem. Phys.
25
26 1986, 109, 131−143.
27
28
29 (19) Bachler, V.; Olbrich, G.; Neese, F.; Wieghardt, K. Theoretical Evidence for the Singlet
30
31 Diradical Character of Square Planar Nickel Complexes Containing Two o-Semiquinonato Type
32
33 Ligands. Inorg. Chem. 2002, 41, 4179−4193.
34
35
36 (20) Rajput, A.; Sharma, A. K.; Barman, S. K.; Koley, D.; Steinert, M.; Mukherjee, R. Neutral,
37
38 Cationic, and Anionic Low-Spin Iron(III) Complexes Stabilized by Amidophenolate and
39
40 Iminobenzosemiquinonate Radical in N,N,O Ligands. Inorg. Chem. 2014, 53, 36−48.
41
42
(21) The DFT barriers obtained at the B3LYP level are slightly higher than the experimental
43
44
45 temperature. Because of several reasons such as presence (lower energy) dormant species that
46
47 were not explored computationally, the experimental reaction can proceed at room temperature.
48
49 The computational barriers might also be somewaht overestimated. Higher level calculations
50
51 perhaps lead to somewhat lower barriers.
52
53
54 (22) Barone,V.; Cossi, M.; Rega, N.; Scalmani, G. Energies, Structures, and Electronic Properties
55
56 of Molecules in Solution with the C-PCM Solvation Model. J. Comp. Chem, 2003, 24, 669-681.
57
58
59
60 ACS Paragon Plus Environment
Page 43 of 43 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 ACS Paragon Plus Environment

S-ar putea să vă placă și