Base Metal-Catalyzed Direct Olefinations of Alcohols with Sulfones

3 hours ago - ABSTRACT: Herein, a base-metal nickel-catalyzed direct olefination of alcohols with sulfones is reported. The reaction operates under lo...
0 downloads 0 Views 942KB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 7082−7087

http://pubs.acs.org/journal/acsodf

Base Metal-Catalyzed Direct Olefinations of Alcohols with Sulfones Satyadeep Waiba, Animesh Das, Milan K. Barman, and Biplab Maji* Department of Chemical Sciences, Indian Institute of Science Education and Research Kolkata, Mohanpur 741246, India

ACS Omega 2019.4:7082-7087. Downloaded from pubs.acs.org by 81.22.46.223 on 04/18/19. For personal use only.

S Supporting Information *

ABSTRACT: Herein, a base-metal nickel-catalyzed direct olefination of alcohols with sulfones is reported. The reaction operates under low catalyst loading and does not require an external redox reagent. A wide range of trans-stilbenes and styrenes were synthesized in good yields and selectivities. Biologically active stilbene DMU-212 could also be synthesized in a single step under these conditions. Mechanistic studies involving kinetic isotope effect, deuterium labeling experiments, and catalytic and stoichiometric reactions with possible catalytic intermediates were performed to elucidate a plausible mechanism.

1. INTRODUCTION

2. RESULTS AND DISCUSSION Proceeding on our quest of developing sustainable and environmentally friendly chemistry by using first-row transition metals,19 we speculated whether or not we could synthesize olefins by olefinating alcohols with sulfones. Alcohols can be easily obtained from various industrial sources or can be derived from many sustainable precursors. Using alcohol, we had earlier reported an olefination protocol of methylsubstituted heteroarenes catalyzed by manganese.19a Although a wide variety of olefins were efficiently synthesized by following this methodology, the need of heteroarenes limited the scope to a certain aspect. In this work, we chose a different path of synthesizing olefins using readily available Julia olefinating reagent sulfones. Olefination of alcohols with sulfones has been performed earlier by using homogeneous ruthenium pincer complex20 and a heterogeneous platinum catalyst21a in the presence of excess amount of strong base (Scheme 1). The reaction has also been performed in the presence of an excess base in the presence of alcohol itself as the solvent.21b Although this protocol provides a promising aspect as a metal-free methodology, the use of excess alcohol as the solvent does not provide the sustainability and can be pointed out as one of the limitations of the said procedure. Herein, we report on the direct olefination of alcohols in the presence of more accessible nickel and diversely available nitrogen ligand. To the best of our knowledge, a base-metalcatalyzed oxidant-free Julia olefination of alcohols has not been reported thus far. We started our investigation by taking (benzylsulfonyl)benzene 1a and benzyl alcohol 2a as the model substrate (Table 1). To test the efficacy of various Ni catalysts, various commercially available ligands were tested with NiBr2 in the presence of KOH in dioxane at 140 °C for 24 h. To our delight, 1,10-phenanthroline L1 gave the desired stilbene 3a in

Olefins or alkenes are one of the most important classes of organic compounds with a wide range of applications.1 Many styrene derivatives displayed critical biological activities such as resistance to cardiovascular and cancerous diseases.2 They also find use as sensors and in vivo imaging agents.3 These compounds are also widely used for sensitization, electroluminescence, and photochromism, as well as in organic dyes.4 The importance of this class of compounds can be easily illustrated by the number of synthetic strategies that have been developed for their synthesis. Among such methodologies, the leading one is carbonyl olefination using various carbanions equipped with leaving functionalities,5 such as PR3 in Witting reaction,6 P(O)(OR)2 in Horner−Wadsworth−Emmons reaction,7 SiR3 in Peterson olefination,8 or SO2R in Julia olefination.9 More effective strategies like catalytic Heck reactions, Suzuki coupling,10 Tebbe olefination,11 olefin metathesis,12 etc. are developed. However, the need for prefunctionalized substrates and/or precious metals as catalysts is their significant drawback. For the last few decades, acceptorless dehydrogenative coupling of alcohols has developed as an attractive tool for synthesizing unsaturated compounds. All of these reactions are not only atom-economic and environmentally benign, but also redox-economic as an oxidation step is saved in the process.13 These reactions were performed by using highly toxic and costly noble-metal catalysts.14 Their replacement by more available base metals is more sought after as they provide not only sustainability and economic relief, but have also been found to show novel reactivity, which is evident by the significant advances made in this field.15 In this aspect, nickel, which is one of the abundant 3d metals, has been applied in many cross-coupling16 and photoredox catalysis.17 However, to date, nickel-catalyzed hydrogen autotransfer or borrowing hydrogenation reaction to produce unsaturated compounds remains underexplored.18 © 2019 American Chemical Society

Received: February 28, 2019 Accepted: April 9, 2019 Published: April 18, 2019 7082

DOI: 10.1021/acsomega.9b00567 ACS Omega 2019, 4, 7082−7087

ACS Omega

Article

(Table 2) and we were delighted to find that the efficiency of NiBr2/L2 in the prototype system could also be expanded to

Scheme 1. Synthesis of Olefins

Table 2. Scope of Nickel-Catalyzed Olefination of Primary Alcohols 2a−l with 1aa

Table 1. Key Optimization Studiesa

a

entry

R

product

yield (%)

1 2 3 4 5 6 7 8 9 10 11 12

C6H5 (2a) 4-OMeC6H4 (2b) 4-PhC6H4 (2c) 4-MeC6H4 (2d) 4-ClC6H4 (2e) 3-ClC6H4 (2f) 4-BrC6H4 (2g) 3-IC6H4 (2h) 4-CF3C6H4 (2i) 2-napthyl (2j) 3-pyridyl (2k) 2-thiphenyl (2l)

3a 3b 3c 3d 3e 3f 3g 3h 3i 3j 3k 3l

89 92 68 53 66 70 57 72 49 62 83 54

Reaction conditions: Table 1, entry 2. Isolated yields.

various alcohols. Importantly, the reaction tolerates a number of halo-substituted benzyl alcohols 2e−h delivering the desired stilbenes with yields up to 72% with the retention of the halofunctionalities and thus leaving room for further functionalization. Alcohols with electron-rich (2b) and electron-withdrawing (2i) substituents also responded well to the optimized conditions delivering the stilbenes in satisfactory yields. Naphthyl (2j) and heteroaromatic alcohols 2k−l were also explored. The pyridyl and thiophenyl groups, which have the potential of binding with the nickel center and thereby poisoning it, were also found to be viable for the olefination reaction delivering the product in 83 and 54% yields, respectively. Various other sulfones 1b−i could also be utilized as the olefinating agent (Table 3). When a series of halo-substituted benzyl sulfones 1b−d were employed in the nickel-catalyzed olefination reaction, the desired stilbenes were isolated in up to 81% yield with the complete retention of the halogen functionality. Similarly, electron rich 4-OMe (1e)-, 4-Me (1f)-, 4-Ph (1g)-, and 2-naphthyl (1h)-substituted sulfones also responded well to the reaction conditions, giving the desired olefins in 52−73% yields and complete E-selectivities. The protocol could also be extended for the synthesis of styrenes using (methylsulfonyl)benzene 1i as the olefinating agent (entries 14−20, Table 3). Thus, primary alcohols 2b−m containing various electronically biased groups, halogens, and naphthyl and thiophenyl functional groups were treated with 1i under standard conditions. Gratifyingly, the desired functionalized styrenes (3q−w) were obtained in 45−68% yields. Note that the undesired α-mehtylstyrenes, which were observed as a byproduct (up to 1:8 ratio) in the previous rutheniumcatalyzed process, were not observed in the present reaction conditions.20 The developed protocol was further applied for the preparation of DMU-212, which inhibits tumor growth and also found to show anticarcinogenic activities.22 The treatment of the alcohol 2n with the sulfone 1e under the standard reaction condition delivered DMU-212 in 76% isolated yield with complete E-selectivity (Scheme 2). Note that DMU-212

a

Reaction conditions: 1a (0.15 mmol), 2a (0.1 mmol), Ni salt (0.005 mmol), ligand (0.006 mmol), and KOH (0.1 mmol) under argon atmosphere at 140 °C in 1,4-dioxane (0.5 mL) for 24 h in a Schlenk tube. Yields were determined by gas chromatography−mass spectrometry (GC−MS) using mesitylene as an internal standard (isolated yields are in parentheses).

85% yield (entry 1). Changing the ligand to neocuproine L2 further improved the efficiency (entry 2), and the desired product could be isolated in 89% yield. Other nitrogen ligands (L3−L4) and several phosphine ligands (L5−L8) were also explored, but none of them could surpass the efficacy of NiBr2/ L2 system. Furthermore, other Ni(II) precursors were found to be less efficient (entries 9−11). The influence and scope of solvents, bases, and base loading were also examined, and it did not improve the reaction efficiency (see Supporting Information (SI) for details). The control experiment demonstrated that only 12% of 3a was formed in the absence of NiBr2/L2 system. Under these optimized conditions, we explored the scope of this protocol. Initially, various primary alcohols were tested 7083

DOI: 10.1021/acsomega.9b00567 ACS Omega 2019, 4, 7082−7087

ACS Omega

Article

Table 3. Scope of Nickel-Catalyzed Olefination Reactions of Primary Alcohols 2 with Sulfones 1a

entry

R

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

C6H5 (1a) 4-ClC6H4 (1b) 4-ClC6H4 (1b) 4-ClC6H4 (1b) 4-BrC6H4 (1c) 4-BrC6H4 (1c) 3-lC6H4 (1d) 4-OMeC6H4 (1e) 4-MeC6H4 (1f) 4-MeC6H4 (1f) 4-PhC6H4 (1g) 4-PhC6H4 (1g) 2-naphthyl (1h) H (1i) H (1i) H (1i) H (1i) H (1i) H (1i) H (1i)

R′ C6H5 (2a) C6H5 (2a) 4-OMeC6H4 (2b) 4-PhC6H4 (2c) C6H5 (2a) 4-OMeC6H4 (2b) C6H5 (2a) C6H5 (2a) C6H5 (2a) 4-PhC6H4 (2c) C6H5 (2a) 4-MeC6H4 (2d) C6H5 (2a) 4-OMeC6H4 (2b) 4-MeC6H4 (2d) 4-ClC6H4 (2e) 3-IC6H4 (2h) 2-napthyl (2j) 2-thiphenyl (2l) benzo[d][1,3]dioxolyl (2m)

product

yield (%)

3a 3e 3m 3n 3g 3o 3h 3b 3d 3p 3c 3p 3j 3q 3r 3s 3t 3u 3v 3w

89 81 62 66 68 55 51 73 52 65 70 75 52 55b 48b 45b 68b 65 56b 45b

Scheme 3. Proposed Reaction Mechanism

and thus it provides the opportunity of reusability. Further, the aldehyde intermediate could be detected when the reaction was conducted without 1a (see the SI). Then, we did some deuterium labeling experiments to shed light on the mechanism involved in the reaction (Scheme 4). Scheme 4. Mechanistic Experiments

a

Reaction conditions: Table 1, entry 2. Isolated yields. bNMR yields with mesitylene as an internal standard.

Scheme 2. Synthesis of DMU-212 and Resveratrol

The olefination of [α,α-D2]biphenyl methanol 2c-d2 (94% D) with 1f was studied under standard reaction condition, and the analysis of the product distribution using 1H NMR revealed 78% D incorporation (83% retention) at the α-position along with 17% D incorporation at the β-position (Scheme 4a). Then, a parallel experiment of 2c and 2c-d2 with 1f was performed to perceive about the kinetic isotope effect (KIE). The observed product ratio based on 1H NMR analysis pointed out kH/kD = 3.1. This indicated that the dehydrogenation of the alcohol took place at or before the ratedetermining step. The treatment of 4-((phenylsulfonyl)methyl-D2)-1,1′-biphenyl 1g-d2 (82% D) with 2d, on the other hand, gave 3p in 74% yield without deuterium incorporation (Scheme 4b), and KIE = 1 is in agreement with the operation of fast base-mediated Aldol condensation reaction. The intermediacy of a nickel hydride species was then probed. An attempt to synthesize the hydride from NiBr2/L2

could quantitatively be deprotected in the presence of BBr3 to yield resveratrol,23 which displays antioxidant, anticancer, antiinflammatory, and other biological activities (Scheme 2).24 Next, to get an insight into the mechanistic pathway involved in the olefination reaction, a series of controlled experiments were carried out. At the onset, we hypothesized that a multistep catalytic process is involved in this nickelcatalyzed olefination reaction (Scheme 3). More specifically, the involvement of a dehydrogenation/hydrogenation cascade process (hydrogen autotransfer) and a base-mediated elimination of sulfinate for the generation of the olefins is proposed. During the control experiments, we observed that the product did not form in the absence of a base. We envisioned that the base not only involves in the elimination step, but also activates the nickel precatalyst via dehalogenation. The formation of the byproduct PhSO2K was confirmed by its isolation (72%) and characterization by 1H NMR and HRMS analyses (see the SI), 7084

DOI: 10.1021/acsomega.9b00567 ACS Omega 2019, 4, 7082−7087

ACS Omega

Article

Then, 1 (0.15 mmol) followed by 2 (0.1 mmol) was added and the tube was closed. It was aged at 140 °C under Ar for 24 h. After quenching the reaction with water (2 mL) and ethyl acetate (3 × 10 mL) extraction, it was dried over Na2SO4 and purified by column chromatography. 4.5. General Procedure of Substituted Styrene Synthesis. KOH (5.6 mg, 0.1 mmol), NiBr2 (1.1 mg, 0.005 mmol), and neocuproine (1.3 mg, 0.006 mmol) were taken along with 1,4-dioxane (0.5 mL) in a Schlenk tube (15 mL). Then, methyl phenyl sulfone 1i (0.15 mmol) followed by alcohol (0.1 mmol) was added and the tube was closed. It was aged at 140 °C under Ar for 24 h. Mesitylene (0.1 mmol) and CDCl3 (0.5 mL) were then added, and the mixture was filtered through a small plug of silica. NMR spectra were measured directly from the clear solution of the filtrate to record the yield. 4.5.1. (E)-1,2-Diphenylethene (3a). Yield 16 mg (0.089 mmol, 89%). Spectra referenced with the previous report.21 4.5.2. (E)-1-Methoxy-4-styrylbenzene (3b). Yield 19.3 mg (0.092 mmol, 92%). Spectra referenced with the previous report.21 4.5.3. (E)-4-Styryl-1,1′-biphenyl (3c). Yield 17.4 mg (0.068 mmol, 68%). Spectra referenced with the previous report.21 4.5.4. (E)-1-Methyl-4-styrylbenzene (3d). Yield 10.3 mg (0.053 mmol, 53%). Spectra referenced with the previous report.21 4.5.5. (E)-1-Chloro-4-styrylbenzene (3e). Yield 14.2 mg (0.066 mmol, 66%). Spectra referenced with the previous report.21 4.5.6. (E)-1-Chloro-3-styrylbenzene (3f). Yield 15 mg (0.070 mmol, 70%). Spectra referenced with the previous report.27 4.5.7. (E)-1-Bromo-4-styrylbenzene (3g). Yield 14.8 mg (0.057 mmol, 57%). Spectra referenced with the previous report.28 4.5.8. (E)-1-Iodo-3-styrylbenzene (3h). Yield 22 mg (0.072 mmol, 72%). Spectra referenced with the previous report.29 4.5.9. (E)-1-Styryl-4-(trifluoromethyl)benzene (3i). Yield 12.2 mg (0.049 mmol, 49%). Spectra referenced with the previous report.30 4.5.10. (E)-2-Styrylnaphthalene (3j). Yield 14.3 mg (0.062 mmol, 62%). Spectra referenced with the previous report.21 4.5.11. (E)-3-Styrylpyridine (3k). Yield 15 mg (0.083 mmol, 83%). Spectra referenced with the previous report.21 4.5.12. (E)-2-Styrylthiophene (3l). Yield 10 mg (0.054 mmol, 54%). Spectra referenced with the previous report.21 4.5.13. (E)-1-Chloro-4-(4-methoxystyryl)benzene (3m). Yield 15 mg (0.062 mmol, 62%). Spectra referenced with the previous report.31 4.5.14. (E)-4-(4-Chlorostyryl)-1,1′-biphenyl (3n). Yield 19 mg (0.066 mmol, 66%). Spectra referenced with the previous report.32 4.5.15. (E)-1-Bromo-4-(4-methoxystyryl)benzene (3o). Yield 16 mg (0.055 mmol, 55%). Spectra referenced with the previous report.33 4.5.16. (E)-4-(4-Methylstyryl)-1,1′-biphenyl (3p). Yield 17.6 mg (0.065 mmol, 65%). Spectra referenced with the previous report.34 4.5.17. 2-Vinylnaphthalene (3u). Yield 10 mg (0.065 mmol, 65%). Spectra referenced with the previous report.21 4.5.18. (E)-1,3-Dimethoxy-5-(4-methoxystyryl)benzene (3x). Yield 76% (21 mg, 0.076 mmol). Spectra referenced with the previous report.23

was unsuccessful using several hydride donors. A similar nickel hydride was found to be unstable and could not be detected at −75 °C.18c Conveniently, a Ni−H species L6−Ni−H was independently prepared using an electron-rich phosphine (L6).25 Importantly, the catalytic reaction using L6−Ni−H (5 mol %) proceeded at a similar efficiency to the NiBr2/L6 system (Scheme 4c). Further, the intermediate vinyl sulfone 4a was independently prepared and reacted with 2a under standard reaction condition using L6−Ni−H as a catalyst to prove the intermediacy of Ni−H species, a vinyl sulfone 4, and the involvement of alcohol as the hydride source (Scheme 4d). This delivered 3a in 85% yield. Again, a similar reaction of the vinyl sulfone 4b with 2c-d2 (94% D) was performed under standard reaction condition (Scheme 4e). 1H NMR analysis of the product (63% yield) revealed 62 and 14% D deuterium incorporation at the α- and β-positions, respectively. These results clearly indicated the intermediacy of a nickel hydride and vinyl sulfone 4 species and hydrogenation of 4 during the reaction, where the alcohol acts as the generic hydride source.

3. CONCLUSIONS In conclusion, direct olefination of primary alcohols with Julia olefinating agent was performed under mild reaction condition without the addition of an external reducing agent as the alcohol acts as the source of both the aldehyde and the hydride. A complex of a base metal nickel with an inexpensive commercially available nitrogen ligand catalyzed the reaction. A wide range of olefins were successfully synthesized with perfect (E) selectivity. The use of earth-abundant metals as catalyst can provide sustainability and potentially replace rare noble metals in homogeneous catalysis. We hope that this report provides a step further toward the evolution of sustainable and environmentally benign chemistry. 4. EXPERIMENTAL SECTION 4.1. General Considerations. The reactions have been conducted under argon atmosphere in oven-dried glassware. Temperatures for the reactions have been reported as the temperature of the oil bath that surrounds the reaction tube. For drying of solvents, standard procedures were followed. For recording 1H, 13C NMR spectra, Bruker (1H: 500 MHz, 13C {1H}1: 126 MHz) and JEOL (1H: 400 MHz, 13C {1H}: 101 MHz) were used, and the referencing is according to the solvent residual resonance. Multiplicities have been indicated as: br for broad, s for singlet, d for doublet, t for triplet, or m for multiplet. Coupling constants (J) are reported in hertz (Hz). Mass spectra have been recorded on Bruker micrOTOFQ II Spectrometer. 4.2. General Procedure of Alkyl Sulfone (1) Preparation. Alkyl sulfones were prepared according to the previous report.26 4.3. Optimization Studies (Table 1, Tables S1−S7). Base (0−100 mol %) and catalyst (0−5 mol %) were taken in a Schlenk tube (15 mL) and solvent (0.5 mL) was added. Then, 2a (0.1 mmol) followed by 1a (0.15 mmol) was added, and the tube was closed. It was aged at 140 °C under Ar for the specified time. After that, GC yield was measured by using mesitylene as the standard. 4.4. General Procedure of Substituted Stilbene Synthesis. KOH (5.6 mg, 0.1 mmol), NiBr2 (1.1 mg, 0.005 mmol), and neocuproine (1.3 mg, 0.006 mmol) were taken along with 1,4-dioxane (0.5 mL) in a Schlenk tube (15 mL). 7085

DOI: 10.1021/acsomega.9b00567 ACS Omega 2019, 4, 7082−7087

ACS Omega



Article

Molecular Clips for the Coordination-Driven Cofacial Assembly of πConjugated Systems. J. Am. Chem. Soc. 2006, 128, 3520−3521. (c) Shigemitsu, Y.; Komiya, K.; Mizuyama, N.; Tominaga, Y. Reaction of functionalized maleimides with versatile nucleophiles. Synthesis, electronic spectra and molecular orbital study. Dyes Pigm. 2007, 72, 271−284. (5) Dumeunier, R.; Marko, I. E. Modern Carbonyl Olefination: Methods and Applications; Wiley-VCH: Weinheim, 2004. (6) (a) Wittig, G.; Geissler, G. Zur Reaktionsweise des Pentaphenylphosphors und einiger Derivate. Justus Liebigs Ann. Chem. 1953, 580, 44−57. (b) Maryanoff, B. E.; Reitz, A. B. The Wittig olefination reaction and modifications involving phosphoryl-stabilized carbanions. Stereochemistry, mechanism, and selected synthetic aspects. Chem. Rev. 1989, 89, 863−927. (7) (a) Horner, L.; Hoffmann, H.; Wippel, H. G.; Klahre, G. Phosphororganische Verbindungen, XX. Phosphinoxyde als Olefinierungsreagenzien. Chem. Ber. 1959, 92, 2499−2505. (b) Clayden, J.; Warren, S. Stereocontrol in Organic Synthesis Using the Diphenylphosphoryl Group. Angew. Chem., Int. Ed. 1996, 35, 241−270. (8) (a) Peterson, D. J. Carbonyl olefination reaction using silylsubstituted organometallic compounds. J. Org. Chem. 1968, 33, 780− 784. (b) Staden, L. F. v; Gravestock, D.; Ager, D. J. New developments in the Peterson olefination reaction. Chem. Soc. Rev. 2002, 31, 195−200. (9) Julia, M.; Paris, J.-M. Syntheses a l′aide de sulfones v(+)methode de synthese generale de doubles liaisons. Tetrahedron Lett. 1973, 14, 4833−4836. (10) (a) Heck, R. F.; Nolley, J. P. Palladium-catalyzed vinylic hydrogen substitution reactions with aryl, benzyl, and styryl halides. J. Org. Chem. 1972, 37, 2320−2322. (b) Suzuki, A. Recent advances in the cross-coupling reactions of organoboron derivatives with organic electrophiles, 1995−1998. J. Organomet. Chem. 1999, 576, 147−168. (11) Tebbe, F. N.; Parshall, G. W.; Reddy, G. S. Olefin homologation with titanium methylene compounds. J. Am. Chem. Soc. 1978, 100, 3611−3613. (12) Handbook of Metathesis; Grubbs, R. H., Ed.; Wiley-VCH: Weinheim, 2003; Vol. 1. (13) Burns, N.; Baran, P.; Hoffmann, R. Redox Economy in Organic Synthesis. Angew. Chem., Int. Ed. 2009, 48, 2854−2867. (14) (a) Choi, J.; MacArthur, A. H. R.; Brookhart, M.; Goldman, A. S. Dehydrogenation and Related Reactions Catalyzed by Iridium Pincer Complexes. Chem. Rev. 2011, 111, 1761−1779. (b) Selander, N.; Szabó, K. J. Catalysis by Palladium Pincer Complexes. Chem. Rev. 2011, 111, 2048−2076. (c) Gunanathan, C.; Milstein, D. Applications of Acceptorless Dehydrogenation and Related Transformations in Chemical Synthesis. Science 2013, 341, No. 1229712. (d) Crabtree, R. H. Homogeneous Transition Metal Catalysis of Acceptorless Dehydrogenative Alcohol Oxidation: Applications in Hydrogen Storage and to Heterocycle Synthesis. Chem. Rev. 2017, 117, 9228− 9246. (e) Younus, H. A.; Su, W.; Ahmad, N.; Chen, S.; Verpoort, F. Ruthenium Pincer Complexes: Synthesis and Catalytic Applications. Adv. Synth. Catal. 2015, 357, 283−330. (f) Albrecht, M.; van Koten, G. Platinum Group Organometallics Based on “Pincer” Complexes: Sensors, Switches, and Catalysts. Angew. Chem., Int. Ed. 2001, 40, 3750−3781. (15) (a) Bolm, C.; Legros, J.; Le Paih, J.; Zani, L. Iron-Catalyzed Reactions in Organic Synthesis. Chem. Rev. 2004, 104, 6217−6254. (b) Enthaler, S.; Junge, K.; Beller, M. Sustainable Metal Catalysis with Iron: From Rust to a Rising Star? Angew. Chem., Int. Ed. 2008, 47, 3317−3321. (c) Sherry, B. D.; Fürstner, A. The Promise and Challenge of Iron-Catalyzed Cross Coupling. Acc. Chem. Res. 2008, 41, 1500−1511. (d) Mukherjee, A.; Milstein, D. Homogeneous Catalysis by Cobalt and Manganese Pincer Complexes. ACS Catal. 2018, 11435−11469. (e) Filonenko, G. A.; van Putten, R.; Hensen, E. J. M.; Pidko, E. A. Catalytic (de)hydrogenation promoted by nonprecious metals − Co, Fe and Mn: recent advances in an emerging field. Chem. Soc. Rev. 2018, 47, 1459−1483. (f) Irrgang, T.; Kempe, R. 3d-Metal Catalyzed N- and C-Alkylation Reactions via Borrowing Hydrogen or Hydrogen Autotransfer. Chem. Rev. 2019, 2524−2549.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b00567. Optimization of reaction conditions (Tables S1−S7); mechanistic experiments; and copies of NMR spectra (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Biplab Maji: 0000-0001-5034-423X Funding

This research was funded by IISER Kolkata (start-up grant). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The current research was supported by IISER Kolkata (startup grant). S.W., A.D., and M.K.B. acknowledge CSIR; INSPIRE, and SERB (PDF/2016/001952), respectively, for providing fellowship.



REFERENCES

(1) Matar, S.; Hatch, L. F. Chemistry of Petrochemical Processes; Gulf Professional Publishing: Houston, 2001. (2) (a) Simoni, D.; Roberti, M.; Invidiata, F. P.; Aiello, E.; Aiello, S.; Marchetti, P.; Baruchello, R.; Eleopra, M.; Di Cristina, A.; Grimaudo, S.; Gebbia, N.; Crosta, L.; Dieli, F.; Tolomeo, M. Stilbene-based anticancer agents: Resveratrol analogues active toward HL60 leukemic cells with a non-specific phase mechanism. Bioorg. Med. Chem. Lett. 2006, 16, 3245−3248. (b) Jung, J.-C.; Lim, E.; Lee, Y.; Kang, J.-M.; Kim, H.; Jang, S.; Oh, S.; Jung, M. Synthesis of novel trans-stilbene derivatives and evaluation of their potent antioxidant and neuroprotective effects. Eur. J. Med. Chem. 2009, 44, 3166−3174. (c) Hagiwara, K.; Kosaka, N.; Yoshioka, Y.; Takahashi, R.-u.; Takeshita, F.; Ochiya, T. Stilbene derivatives promote Ago2dependent tumour-suppressive microRNA activity. Sci. Rep. 2012, 2, No. 314. (d) Dai, J.; Liu, Z.-Q.; Wang, X.-Q.; Lin, J.; Yao, P.-F.; Huang, S.-L.; Ou, T.-M.; Tan, J.-H.; Li, D.; Gu, L.-Q.; Huang, Z.-S. Discovery of Small Molecules for Up-Regulating the Translation of Antiamyloidogenic Secretase, a Disintegrin and Metalloproteinase 10 (ADAM10), by Binding to the G-Quadruplex-Forming Sequence in the 5′ Untranslated Region (UTR) of Its mRNA. J. Med. Chem. 2015, 58, 3875−3891. (3) (a) Kang, N.-Y.; Ha, H.-H.; Yun, S.-W.; Yu, Y. H.; Chang, Y.-T. Diversity-driven chemical probe development for biomolecules: beyond hypothesis-driven approach. Chem. Soc. Rev. 2011, 40, 3613−3626. (b) Li, Q.; Min, J.; Ahn, Y.-H.; Namm, J.; Kim, E. M.; Lui, R.; Kim, H. Y.; Ji, Y.; Wu, H.; Wisniewski, T.; Chang, Y.-T. Styryl-Based Compounds as Potential in vivo Imaging Agents for βAmyloid Plaques. ChemBioChem 2007, 8, 1679−1687. (c) Li, Q.; Lee, J.-S.; Ha, C.; Park, C. B.; Yang, G.; Gan, W. B.; Chang, Y.-T. SolidPhase Synthesis of Styryl Dyes and their Application as Amyloid Sensors. Angew. Chem., Int. Ed. 2004, 43, 6331−6335. (d) Wang, S.; Chang, Y.-T. Discovery of heparin chemosensors through diversity oriented fluorescence library approach. Chem. Commun. 2008, 1173− 1175. (4) (a) Kraft, A.; Grimsdale, A. C.; Holmes, A. B. Electroluminescent Conjugated PolymersSeeing Polymers in a New Light. Angew. Chem., Int. Ed. 1998, 37, 402−428. (b) Nohra, B.; Graule, S.; Lescop, C.; Réau, R. Mimicking [2,2]Paracyclophane Topology: 7086

DOI: 10.1021/acsomega.9b00567 ACS Omega 2019, 4, 7082−7087

ACS Omega

Article

(g) Liu, T.; Wang, L.; Wu, K.; Yu, Z. Manganese-Catalyzed βAlkylation of Secondary Alcohols with Primary Alcohols under Phosphine-Free Conditions. ACS Catal. 2018, 8, 7201−7207. (h) Mukherjee, A.; Milstein, D. Homogeneous Catalysis by Cobalt and Manganese Pincer Complexes. ACS Catal. 2018, 8, 11435− 11469. (i) Liu, W.; Sahoo, B.; Junge, K.; Beller, M. Cobalt Complexes as an Emerging Class of Catalysts for Homogeneous Hydrogenations. Acc. Chem. Res. 2018, 51, 1858−1869. (j) Bhattacharya, P.; Krause, J. A.; Guan, H. Mechanistic Studies of Ammonia Borane Dehydrogenation Catalyzed by Iron Pincer Complexes. J. Am. Chem. Soc. 2014, 136, 11153−11161. (16) (a) Rosen, B. M.; Quasdorf, K. W.; Wilson, D. A.; Zhang, N.; Resmerita, A.-M.; Garg, N. K.; Percec, V. Nickel-Catalyzed CrossCouplings Involving Carbon−Oxygen Bonds. Chem. Rev. 2011, 111, 1346−1416. (b) Mesganaw, T.; Garg, N. K. Ni- and Fe-Catalyzed Cross-Coupling Reactions of Phenol Derivatives. Org. Process Res. Dev. 2013, 17, 29−39. (17) (a) Terrett, J. A.; Cuthbertson, J. D.; Shurtleff, V. W.; MacMillan, D. W. C. Switching on elusive organometallic mechanisms with photoredox catalysis. Nature 2015, 524, 330. (b) Hwang, S. J.; Powers, D. C.; Maher, A. G.; Anderson, B. L.; Hadt, R. G.; Zheng, S.L.; Chen, Y.-S.; Nocera, D. G. Trap-Free Halogen Photoelimination from Mononuclear Ni(III) Complexes. J. Am. Chem. Soc. 2015, 137, 6472−6475. (c) Shields, B. J.; Doyle, A. G. Direct C(sp3)−H Cross Coupling Enabled by Catalytic Generation of Chlorine Radicals. J. Am. Chem. Soc. 2016, 138, 12719−12722. (d) Welin, E. R.; Le, C.; Arias-Rotondo, D. M.; McCusker, J. K.; MacMillan, D. W. C. Photosensitized, energy transfer-mediated organometallic catalysis through electronically excited nickel(II). Science 2017, 355, 380−385. (e) Qi, Z.-H.; Ma, J. Dual Role of a Photocatalyst: Generation of Ni(0) Catalyst and Promotion of Catalytic C−N Bond Formation. ACS Catal. 2018, 8, 1456−1463. (18) (a) Shimizu, K.-i.; Kon, K.; Onodera, W.; Yamazaki, H.; Kondo, J. N. Heterogeneous Ni Catalyst for Direct Synthesis of Primary Amines from Alcohols and Ammonia. ACS Catal. 2013, 3, 112−117. (b) Shimizu, K.-i.; Imaiida, N.; Kon, K.; Hakim Siddiki, S. M. A.; Satsuma, A. Heterogeneous Ni Catalysts for N-Alkylation of Amines with Alcohols. ACS Catal. 2013, 3, 998−1005. (c) Vellakkaran, M.; Singh, K.; Banerjee, D. An Efficient and Selective Nickel-Catalyzed Direct N-Alkylation of Anilines with Alcohols. ACS Catal. 2017, 7, 8152−8158. (d) Das, S.; Maiti, D.; De Sarkar, S. Synthesis of Polysubstituted Quinolines from α-2-Aminoaryl Alcohols Via NickelCatalyzed Dehydrogenative Coupling. J. Org. Chem. 2018, 83, 2309− 2316. (e) Midya, S. P.; Rana, J.; Pitchaimani, J.; Nandakumar, A.; Madhu, V.; Balaraman, E. Ni-Catalyzed α-Alkylation of Unactivated Amides and Esters with Alcohols by Hydrogen Auto-Transfer Strategy. ChemSusChem 2018, 11, 3911−3916. (f) Yang, P.; Zhang, C.; Ma, Y.; Zhang, C.; Li, A.; Tang, B.; Zhou, J. S. Nickel-Catalyzed N-Alkylation of Acylhydrazines and Arylamines Using Alcohols and Enantioselective Examples. Angew. Chem., Int. Ed. 2017, 56, 14702− 14706. (g) Khan, I. A.; Saxena, A. K. Metal-Free, Mild, Nonepimerizing, Chemo- and Enantio- or Diastereoselective N-Alkylation of Amines by Alcohols via Oxidation/Imine−Iminium Formation/ Reductive Amination: A Pragmatic Synthesis of Octahydropyrazinopyridoindoles and Higher Ring Analogues. J. Org. Chem. 2013, 78, 11656−11669. (h) Parua, S.; Das, S.; Sikari, R.; Sinha, S.; Paul, N. D. One-Pot Cascade Synthesis of Quinazolin-4(3H)-ones via NickelCatalyzed Dehydrogenative Coupling of o-Aminobenzamides with Alcohols. J. Org. Chem. 2017, 82, 7165−7175. (i) Parua, S.; Sikari, R.; Sinha, S.; Chakraborty, G.; Mondal, R.; Paul, N. D. Accessing Polysubstituted Quinazolines via Nickel Catalyzed Acceptorless Dehydrogenative Coupling. J. Org. Chem. 2018, 83, 11154−11166. (19) (a) Barman, M. K.; Waiba, S.; Maji, B. Manganese-Catalyzed Direct Olefination of Methyl-Substituted Heteroarenes with Primary Alcohols. Angew. Chem., Int. Ed. 2018, 57, 9126−9130. (b) Barman, M. K.; Jana, A.; Maji, B. Phosphine-Free NNN-Manganese Complex Catalyzed α-Alkylation of Ketones with Primary Alcohols and Friedländer Quinoline Synthesis. Adv. Synth. Catal. 2018, 360, 3233−3238. (c) Jana, A.; Reddy, C. B.; Maji, B. Manganese Catalyzed

α-Alkylation of Nitriles with Primary Alcohols. ACS Catal. 2018, 8, 9226−9231. (d) Maji, B.; Barman, M. K. Recent Developments of Manganese Complexes for Catalytic Hydrogenation and Dehydrogenation Reactions. Synthesis 2017, 49, 3377−3393. (20) Srimani, D.; Leitus, G.; Ben-David, Y.; Milstein, D. Direct catalytic olefination of alcohols with sulfones. Angew. Chem., Int. Ed. 2014, 53, 11092−11095. (21) (a) Siddiki, S. M. A. H.; Touchy, A. S.; Kon, K.; Shimizu, K.-i. Direct Olefination of Alcohols with Sulfones by Using Heterogeneous Platinum Catalysts. Chem. - Eur. J. 2016, 22, 6111−6119. (b) Yao, C.Z.; Li, Q.-Q.; Wang, M.-M.; Ning, X.-S.; Kang, Y.-B. (E)-Specific direct Julia-olefination of aryl alcohols without extra reducing agents promoted by bases. Chem. Commun. 2015, 51, 7729−7732. (22) (a) Piotrowska, H.; Myszkowski, K.; Abraszek, J.; KwiatkowskaBorowczyk, E.; Amarowicz, R.; Murias, M.; Wierzchowski, M.; Jodynis-Liebert, J. DMU-212 inhibits tumor growth in xenograft model of human ovarian cancer. Biomed. Pharmacother. 2014, 68, 397−400. (b) Piotrowska, H.; Kujawska, M.; Nowicki, M.; Petzke, E.; Ignatowicz, E.; Krajka-Kuźniak, V.; Zawierucha, P.; Wierzchowski, M.; Murias, M.; Jodynis-Liebert, J. Effect of resveratrol analogue, DMU212, on antioxidant status and apoptosis-related genes in rat model of hepatocarcinogenesis. Hum. Exp. Toxicol. 2017, 36, 160−175. (23) Li, Q.-Q.; Shah, Z.; Qu, J.-P.; Kang, Y.-B. Direct Wittig Olefination of Alcohols. J. Org. Chem. 2018, 83, 296−302. (24) Santos, J. A.; de Gustavo, S. G.; Virgínia, O.; Nadia, R. B. R.; da Silva, A. D. Resveratrol and Analogues: A Review of Antioxidant Activity and Applications to Human Health. Recent Pat. Food, Nutr. Agric. 2013, 5, 144−153. (25) (a) Green, M. L. H.; Saito, T.; Tanfield, P. J. Stable nickel hydride complexes of tricyclohexylphosphine and triisopropylphosphine. J. Chem. Soc. A 1971, 152−154. (b) Lindner, M. M.; Beckmann, U.; Frank, W.; Klaui, W. Influence of the Steric Demand of Coligands on the Catalytic Activity of Nickel(II) Complexes in the Copolymerization of Ethene and Carbon Monoxide. ISRN Inorg. Chem. 2013, 2013, 13. (26) Waiba, S.; Barman, M. K.; Maji, B. Manganese-Catalyzed Acceptorless Dehydrogenative Coupling of Alcohols With Sulfones: A Tool To Access Highly Substituted Vinyl Sulfones. J. Org. Chem. 2019, 84, 973−982. (27) Etemadi-Davan, E.; Iranpoor, N. Efficient Ni-catalyzed conversion of phenols protected with 2,4,6-trichloro-1,3,5-triazine (TCT) to olefins. Chem. Commun. 2017, 53, 12794−12797. (28) Movassagh, B.; Ranjbari, S. Kryptofix 5 as an inexpensive and efficient ligand for the palladium-catalyzed Mizoroki-Heck reaction. Appl. Organomet. Chem. 2018, 32, No. e4224. (29) Lautens, M.; Roy, A.; Fukuoka, K.; Fagnou, K.; Martín-Matute, B. Rhodium-Catalyzed Coupling Reactions of Arylboronic Acids to Olefins in Aqueous Media. J. Am. Chem. Soc. 2001, 123, 5358−5359. (30) Das, U. K.; Chakraborty, S.; Diskin-Posner, Y.; Milstein, D. Direct Conversion of Alcohols into Alkenes by Dehydrogenative Coupling with Hydrazine/Hydrazone Catalyzed by Manganese. Angew. Chem., Int. Ed. 2018, 57, 13444−13448. (31) Takahashi, K.; Morishita, H.; Ogiwara, Y.; Sakai, N. Group 4 Metallocene Difluoride/Palladium Bimetallic Catalysts for the Reductive Cross-Coupling of Alkynes with Aryl Iodides and Bromides. J. Org. Chem. 2018, 83, 13734−13742. (32) Liu, Y.; Liu, P.; Liu, Y.; Wei, Y. Pd(0)-Catalyzed Tandem OnePot Reaction of Biphenyl Ketones/Aldehydes to the Corresponding Di-substituted Aryl Olefins. Chin. J. Chem. 2017, 35, 1141−1148. (33) Aukland, M. H.; Talbot, F. J. T.; Fernández-Salas, J. A.; Ball, M.; Pulis, A. P.; Procter, D. J. An Interrupted Pummerer/NickelCatalysed Cross-Coupling Sequence. Angew. Chem., Int. Ed. 2018, 57, 9785−9789. (34) Iwasaki, T.; Akimoto, R.; Kuniyasu, H.; Kambe, N. FeCatalyzed Cross-Coupling Reaction of Vinylic Ethers with Aryl Grignard Reagents. Chem. - Asian J. 2016, 11, 2834−2837.

7087

DOI: 10.1021/acsomega.9b00567 ACS Omega 2019, 4, 7082−7087