Acessibilidade / Reportar erro

EMERGING TRENDS OF ORGANIC ELECTROSYNTHESIS IN BRAZIL

Resumo

This review provides an overview of the recent progress in organic electrosynthesis in Brazil, ranging from 2013 to 2023. By highlighting the principles, applications, and limitations of various electrosynthetic methodologies, this article aims to contribute to the understanding of basic principles and stimulate research in this field in Brazil. Moreover, the limitations and challenges associated with organic electrosynthesis are covered. Factors such as electrode material selection, solvent effects, and reaction optimization are discussed, highlighting the need for further research and development to overcome these obstacles. Finally, future prospects are outlined.


INTRODUCTION

Electrosynthesis research and industrial developments have been established over the years, dating back almost two centuries.11 Vogt, H.; J. Electrochem. Soc. 1981, 128, 29C. [Crossref]
Crossref...
Research in organic electrosynthesis has produced a vast literature, and several researchers have stood out in this area.22 Yan, M.; Kawamata, Y.; Baran, P. S.; Chem. Rev. 2017, 117, 13230. [Crossref]
Crossref...
,33 Shatskiy, A.; Lundberg, H.; Kärkäs, M. D.; Chemelectrochem 2019, 6, 4067. [Crossref]
Crossref...
,44 Hilt, G.; ChemElectroChem 2020, 7, 395. [Crossref]
Crossref...
,55 Pollok, D.; Waldvogel, S. R.; Chem. Sci. 2020, 11, 12386. [Crossref]
Crossref...
,66 Ali, W.; Benedetti, R.; Handzlik, J.; Zwergel, C.; Battistelli, C.; Drug Discovery Today 2021, 26, 256. [Crossref]
Crossref...
,77 Li, Y.; Dana, S.; Ackermann, L.; Curr. Opin. Electrochem. 2023, 40, 101312. [Crossref]
Crossref...

In the beginning, electrochemical studies in Brazil focused on fundamental areas such as electrochemical kinetics and reaction mechanisms. Over time, the studies have turned to materials science, such as corrosion and protection, electrocatalysis, water treatment, electrolysis, and organic electrosynthesis.88 Pernaut, J. M.; Matencio, T.; Quim. Nova 1999, 22, 899. [Crossref]
Crossref...
,99 Avaca, L. A.; Tokoro, R.; Quim. Nova 2002, 25, 25. [Crossref]
Crossref...
,1010 de Luca, M. A.; de Luca, S. J.; Santana, M. A.; Quim. Nova 2003, 26, 420. [Crossref]
Crossref...
,1111 Gandra, P. G.; Alves, A. A.; de Macedo, D. V.; Kubota, L. T.; Quim. Nova 2004, 27, 980. [Crossref]
Crossref...
,1212 Batista, E. C.; de Oliveira, R. T. S.; Ferreira, R. Q.; Miwa, D.; dos Santos, M. C.; Quim. Nova 2011, 34, 1517. [Crossref]
Crossref...
,1313 Zanoni, M.; Borges, A.; Benedetti, A.; Yamanaka, H.; Sotomayor, M. P.; Bessegato, G.; Stradiotto, N.; Zanta, C. L.; Andrade, A.; Quim. Nova 2017, 40, 663. [Crossref]
Crossref...
We highlight that nowadays significant advances in the field of organic electrosynthesis have been observed in Brazil, concurrently with the increasing of this scientific pursuit worldwide (Figure 1).

Figure 1
Worldwide publications of scientific articles on organic electrosynthesis between the 1980s and 2023 (source: Scopus)

As a matter of fact, we are facing an era of scarce resources, excessive use of fossil energy, and a high level of pollution, which encourages researchers to seek the development and use of clean energy sources, including in the area of synthetic organic chemistry.1414 Wiebe, A.; Gieshoff, T.; Möhle, S.; Rodrigo, E.; Zirbes, M.; Waldvogel, S. R.; Angew. Chem., Int. Ed. 2018, 57, 5594. [Crossref]
Crossref...
,1515 Beil, S. B.; Pollok, D.; Waldvogel, S. R.; Angew. Chem., Int. Ed. 2021, 60, 14750. [Crossref]
Crossref...
As a consequence, there was a growing worldwide interest for the development of milder synthetic methods, such as the use of organic electrosynthesis to obtain organic molecules.

Organic electrosynthesis is a powerful and sustainable method of conducting organic reactions, which has gained significant attention in recent years for its environmental benefits and synthetic efficiency.1616 Frontana-Uribe, B. A.; Little, R. D.; Ibanez, J. G.; Palma, A.; Vasquez-Medrano, R.; Green Chem. 2010, 12, 2099. [Crossref]
Crossref...
,1717 Meyer, T. H.; Choi, I.; Tian, C.; Ackermann, L.; Chem 2020, 6, 2484. [Crossref]
Crossref...
,1818 Yuan, Y.; Lei, A.; Nat. Commun. 2020, 11, 802. [Crossref]
Crossref...
This innovative technique involves the use of electrical energy to drive chemical transformations of organic compounds. The mild reaction conditions are beneficial, allowing for easier handling of delicate functional groups and a broader range of substrates. By harnessing the principles of electrochemistry, it enables the construction of complex molecules with high selectivity and atom economy, paving the way for green and sustainable synthesis. Electrosynthesis allowed precise control of the reactions by adjusting the current and electrode potential, leading to the desired product with high accuracy. Moreover, it reduces energy requirements, making it environmentally friendly and cost-effective. Organic electrosynthesis can generate redox reagents on the spot from readily available materials, eliminating the need for complex pre-modifications. Unlike traditional methods that depend on costly pre-formed reagents, this approach provides enhanced reaction flexibility, facilitating chemical transformations in an efficient manner, enabling chemical transformations in an efficient pathway, often eliminating complex pre-modifications. Another advantages are the minimal waste produced, the selectivity, reproducibility, scalability under continuous flow conditions,1919 Noël, T.; Cao, Y.; Laudadio, G.; Acc. Chem. Res. 2019, 52, 2858. [Crossref]
Crossref...
,2020 Capaldo, L.; Wen, Z.; Noël, T.; Chem. Sci. 2023, 14, 4230. [Crossref]
Crossref...
and much more, aligning with many green chemistry principles.2121 Cembellín, S.; Batanero, B.; Chem. Rec. 2021, 21, 2453. [Crossref]
Crossref...

Organic reactions often involve the transfer of electrons between different species, which are referred to as redox processes. In a redox reaction, one species undergoes oxidation, losing electrons, while another species undergoes reduction, receiving electrons. Electrons are transferred from the species being oxidized (the reducing agent) to the species being reduced (the oxidizing agent). Reactions performed under electrochemical conditions are enabled by the application of an electric potential which allows a non-spontaneous electron-flow always using an electrochemical cell. Reaction mechanisms can occur in an electrochemical cell similar to reactions in conventional systems. However, in an electrochemical cell, electricity is used to boost reactions by applying a potential difference (electrical voltage) between two electrodes, an anode and a cathode. These reactions are by definition voltage-controlled and very specifically promoted. After the formation of oxidized or reduced intermediates via single electron transfer (SET), which can encompass a variety of species including radicals or formally charged molecules, different reaction pathways can be favored, promoting different transformations, such as functionalization, substitution, elimination, rearrangement, additions, or coupling reactions, among others (Scheme 1a). In general, there are two main types of reactions: direct reactions and mediated reactions (Scheme 1b). In direct reactions, the electron transfer occurs directly between the electrode and the organic substrate. The organic reagent itself accepts or donates electrons directly to the electrode. This is common in electrosynthesis when the organic substrate is easily oxidable or reduced. On the other hand, in mediated reactions, the electron transfer occurs through a redox mediator, which is a compound that accepts or donates electrons between the electrode and the organic reagent. The mediator acts as an intermediary in the electron transfer, allowing the reaction to proceed more efficiently. This is useful when the organic substrate is not directly electroactive.2222 Fuchigami, T.; Atobe, M.; Inagi, S.; Fundamentals and Applications of Organic Electrochemistry, 1st ed.; John Wiley & Sons: Chichester, 2014.,2323 Hammerich, B. S. O.; Organic Electrochemistry, 5th ed.; CRC Press: Boca Raton, 2015. [Crossref]
Crossref...
,2424 Lodh, J.; Paul, S.; Sun, H.; Song, L.; Schöfberger, W.; Roy, S.; Front. Chem. 2023, 10, 1. [Crossref]
Crossref...

Scheme 1
General possible mechanisms for the electron transfer (a) and principle of redox mediation (b)

Experimentally, in an electrochemical cell, two electrodes are immersed in an electrolyte solution containing the reactants. The electrode in which the oxidation occurs is known as the anode, and the electrode at which reduction occurs is the cathode. When an electric potential is applied across the electrodes, electrons flow from the anode to the cathode through the external circuit (potential application), while ions migrate through the electrolyte to maintain overall charge balance.1515 Beil, S. B.; Pollok, D.; Waldvogel, S. R.; Angew. Chem., Int. Ed. 2021, 60, 14750. [Crossref]
Crossref...
The electrolysis setup can be assembled using either a potentiostatic or galvanostatic cell. In potentiostatic cells, a constant potential (voltage) is applied between the working electrode and the reference electrode. The potential is maintained at a fixed value throughout the reaction, and as the reaction proceeds, the current flows between the working and counter electrodes. On the other hand, in galvanostatic cells, a constant current (ampere) is applied between the working and counter electrodes. As the reaction progresses, the potential between the working electrode and the reference electrode is continuously measured to be monitored and stabilized by the variation of the potential. The reference electrode provides a stable potential, essential for accurately measuring the potential of the working electrode. This is valuable in controlling the potential of the electrochemical cell and studying the kinetics of electrochemical reactions. On the other hand, when the goal is to observe just a reaction transformation, electrical current can be applied to the electrochemical cell without the reference electrode, and the products generated can be analyzed using techniques such as spectroscopy or chromatography without the need for rigorous potential control.2525 Yosida, J. In Encyclopedia of Applied Electrochemistry; Kreysa, G.; Ota, K.; Savinell, R. F., eds.; Springer: New York, 2014, p. 386-392.,2626 Roscher, J.; Holze, R.; Encyclopedia 2023, 3, 478. [Crossref]
Crossref...
Both have their unique advantages and applications, and researchers should choose the most suitable configuration based on the objectives of their electrochemical reactions.

Two main types of electrochemical cells are commonly employed: undivided cells or divided cells (Figure 2).2424 Lodh, J.; Paul, S.; Sun, H.; Song, L.; Schöfberger, W.; Roy, S.; Front. Chem. 2023, 10, 1. [Crossref]
Crossref...
Undivided cells do not have a physical barrier between the anode and cathode compartments, allowing the electrolyte solution and reactants to freely mix between the two electrodes. These cells offer simplicity and high current efficiency due to the unrestricted movement of reactants, enabling a variety of paths for electron transfer. However, are susceptible to the unwanted mixture of the electrolyte solution with the reactants between the two electrodes, and may result in non-desirable chemical reactions. On the other hand, divided cells incorporate a physical barrier (a glass frit, a porous ceramic, a porous polymer sheet, or an ion-selective membrane), which separates the anode and cathode compartments. Each compartment contains its own electrode and electrolyte solution, ensuring that the electrochemically formed intermediate at each electrode remains isolated and cannot be mixed. This enhances product selectivity and control over reactions, as it reduces cross-contamination and allows better results. However, divided cells generally require a more complex setup and may experience reduced current efficiency due to increased resistance caused by the presence of the barrier. The choice between undivided and divided cells depends on the specific requirements of the organic electrosynthesis reaction. For simpler setups and higher current efficiency, undivided cells may be the best choice. Conversely, when enhanced selectivity and controlled reactions are essential, divided cells become more advantageous.2222 Fuchigami, T.; Atobe, M.; Inagi, S.; Fundamentals and Applications of Organic Electrochemistry, 1st ed.; John Wiley & Sons: Chichester, 2014.,2323 Hammerich, B. S. O.; Organic Electrochemistry, 5th ed.; CRC Press: Boca Raton, 2015. [Crossref]
Crossref...

Figure 2
Undivided cell (top). Divided cell (bottom). Reprinted with permission from Lodh et al.2424 Lodh, J.; Paul, S.; Sun, H.; Song, L.; Schöfberger, W.; Roy, S.; Front. Chem. 2023, 10, 1. [Crossref]
Crossref...
(copyright (2023) permission from Frontiers Media SA, all rights reserved)

In organic electrosynthesis, the choice of electrode materials plays a crucial role in determining the success and efficiency of the electrochemical reactions. The two primary types of electrodes applied are inert electrodes or sacrificial electrodes. Inert electrodes, such as graphite, platinum, and glassy carbon, are widely employed due to their chemical stability. These electrodes do not participate in the redox reactions themselves and act as mere electron transfer mediators. Their inherent stability ensures minimal contamination, making them suitable for a broad range of electrosynthetic applications. Moreover, inert electrodes offer excellent reproducibility and can endure extended operation periods, making them desirable choices for long-term processes. On the other hand, the sacrifice electrodes are typically composed of metals with low oxidation potentials, and are utilized to release metal ions into the solution during the electrochemical process. These metal ions serve as active species, playing crucial roles as catalysts and coordinators in synthetic organic reactions, and examples of this electrodes are made of zinc, magnesium, aluminum, silver, etc. Furthermore, the choice of electrode material can directly impact the selectivity and yield of the desired products. Certain materials may promote side reactions or generate undesired by-products, while others may offer enhanced control over the chemical reaction.

The choice of an electrode material plays a significant role in the redox potential window for electrosynthesis reactions. The redox potential window refers to the range of electrical potentials within which a specific electrochemical reaction can occur. This choice influences the conditions under which the reaction can take place and can impact the efficiency, selectivity, and yield of the process. These potentials determine at which electrical potentials a species can be reduced or oxidized at the electrode’s surface. Besides the redox potential window, the electrode material can also affect the reaction kinetics, including electron transfer rates, polarization, among other factors. The material should not only exhibit suitable electrical conductivity but also possess chemical stability and compatibility with the reaction medium. Additionally, the electrode’s surface area and porosity can significantly influence the reaction kinetics and mass transfer, thereby affecting the overall efficiency of the electrosynthetic process.2727 Couper, A. M.; Pletcher, D.; Walsh, F. C.; Chem. Rev. 1990, 90, 837. [Crossref]
Crossref...
,2828 Heard, D. M.; Lennox, A. J. J.; Angew. Chem. 2020, 132, 19026. [Crossref]
Crossref...

In electrochemical synthesis, both batch reactions and flow reactions are widely applied techniques. Each approach has its own set of characteristics, advantages, and disadvantages that should be considered when planning and executing electrochemical transformations. Batch reactions involve conducting the entire electrochemical synthesis in a single vessel or cell. The reactants are combined in the cell, and the reaction proceeds until completion. Characteristics of batch reactions include simplicity and ease of setup, making them accessible for both beginners and expert chemists. On the other hand, a continuous flow electrochemistry involves the continuous flow of reactants through an electrochemical cell, wherein the reactions occur as the reagents pass through the cell (Figure 3).2929 Pletcher, D.; Green, R. A.; Brown, R. C. D.; Chem. Rev. 2018, 118, 4573. [Crossref]
Crossref...
Flow reactors in electrochemistry offers several unique advantages over batch reactors conditions, such as improving mixing, fast heat exchange, multistep reaction sequences, smaller amount of electrolyte, reliable scale-up, among others.1919 Noël, T.; Cao, Y.; Laudadio, G.; Acc. Chem. Res. 2019, 52, 2858. [Crossref]
Crossref...
Flow reactions also facilitate the optimization of reaction conditions through rapid screening of different parameters. This accelerates the identification of the optimal reaction conditions and saves valuable time and resources. However, flow reactions may require specialized equipment and expertise in flow chemistry techniques, which can be a barrier for some researchers. Recently, Cantillo and co-workers3030 Maljuric, S.; Jud, W.; Kappe, C. O.; Cantillo, D.; J. Flow Chem. 2020, 10, 181. [Crossref]
Crossref...
reported an organic chemist’s guide for translating batch electrochemistry to single-pass continuous flow conditions, and can be useful for those who want to go deeper into the subject. The choice between the two approaches depends on the specific requirements of the synthesis, the available resources, and the expertise of the researcher.

Considering the current advancements and the significant importance of electrosynthesis in organic chemistry, this short-review article highlights the recent advances in the field of organic electrosynthesis in Brazil, focusing on the period from 2013 to 2023. We discuss the main concepts, practical applications, and also the limitations of electrosynthesis methods, aiming to contribute to a better understanding and progress in this field. Factors such as electrode material selection, solvent effects, and reaction scope will be discussed. Additionally, we evaluate the challenges that organic electrosynthesis faces in Brazil. Finally, we outline some ideas for the future and potential advances of organic electrosynthesis in Brazil.

Figure 3
Syrris® electroflow microflow cell and its components (top) and a generic flow cell in electrosynthesis (bottom). Reprinted with permission from Green et al.3131 Green, R.; Brown, R.; Pletcher, D.; J. Flow Chem. 2015, 5, 31. [Crossref]
Crossref...
(copyright (2015) permission from Springer, all rights reserved)

ORGANIC ELECTROSYNTHESIS IN BRAZIL

Inert electrodes

Over the years, numerous review articles have highlighted the importance of electrosynthesis, discussing important transformations such as electrohalogenation,3232 Scheide, M. R.; Nicoleti, C. R.; Martins, G. M.; Braga, A. L.; Org. Biomol. Chem. 2021, 19, 2578. [Crossref]
Crossref...
electrochemical annulation,3333 Martins, G. M.; Zimmer, G. C.; Mendes, S. R.; Ahmed, N.; Green Chem. 2020, 22, 4849. [Crossref]
Crossref...
phosphorylation,3434 Sbei, N.; Martins, G. M.; Shirinfar, B.; Ahmed, N.; Chem. Rec. 2020, 20, 1530. [Crossref]
Crossref...
alkene difunctionalization,3535 Martins, G. M.; Shirinfar, B.; Hardwick, T.; Ahmed, N.; ChemElectroChem 2019, 6, 1300. [Crossref]
Crossref...
alkynes functionalization,3636 Martins, G. M.; Shirinfar, B.; Hardwick, T.; Murtaza, A.; Ahmed, N.; Catal. Sci. Technol. 2019, 9, 5868. [Crossref]
Crossref...
electrochemical chalcogenation,3737 Martins, G. M.; Meirinho, A. G.; Ahmed, N.; Braga, A. L.; Mendes, S. R.; ChemElectroChem 2019, 6, 5928. [Crossref]
Crossref...
mono-, di-, and trifluoromethylation,3838 Kisukuri, C. M.; Fernandes, V. A.; Delgado, J. A. C.; Häring, A. P.; Paixão, M. W.; Waldvogel, S. R.; Chem. Rec. 2021, 21, 2502. [Crossref]
Crossref...
and metalla-electrocatalysis,3939 Martins, G. M.; Sbei, N.; Zimmer, G. C.; Ahmed, N.; Electrocatalysis and Electrocatalysts for a Cleaner Environment - Fundamentals and Applications, 1st ed.; IntechOpen: London, 2022, ch. 6. [Crossref]
Crossref...
among others.

Significant advances in the synthesis of organochalcogenides have been observed,4040 Rotruck, J. T.; Pope, A. L.; Ganther, H. E.; Swanson, A. B.; Hafeman, D. G.; Hoekstra, W. G.; Science 1973, 179, 588. [Crossref]
Crossref...
,4141 Alberto, E. E.; do Nascimento, V.; Braga, A. L.; J. Braz. Chem. Soc. 2010, 21, 2032. [Crossref]
Crossref...
and they have been increasingly recognized by their diverse biological and medicinal properties, encompassing antiviral, antimicrobial, antitumor, and antioxidant activities.4242 Martins, G. M.; Mendes, S. R. In Organoselenium Chemistry; Hanu, B. C.; Banerjee, B., eds.; De Gruyter: Boston, 2020, ch. 7. [Link] accessed in April 2024
Link...
Considering this, the investigation of environmentally friendly methodologies for the synthesis of chalcogenides compounds has gained significant scientific interest. In 2016, Navarro and co-workers4343 Ribeiro Neto, P. B.; Santana, S. O.; Levitre, G.; Galdino, D.; Oliveira, J. L.; Ribeiro, R. T.; Barros, M. E. S. B.; Bieber, L. W.; Menezes, P. H.; Navarro, M.; Green Chem. 2016, 18, 657. [Crossref]
Crossref...
developed an electrochemical synthesis of organochalcogenides using NaOH aqueous solution employing a divided-cell (Scheme 2, Figure 4). Employing a two-step approach, monochalcogenides were obtained as the major products. Selenide (Se2-), telluride (Te2-), and sulfide (S2-) anions were produced employing a two-compartment electrochemical cell with a Nafion® membrane separation, equipped with stainless steel electrodes. The corresponding chalcogenides (Se, Te or S, 0.5 mmol) were electrochemically reduced in a 0.2 M NaOH solution (pH 13), applying a constant current of 70 mA under an inert atmosphere. Subsequently, the respective organic halide (R–X, 1.0 mmol) in tetrahydrofuran (THF) (5 mL) was added and the mixture was stirred at 25 °C for 12 h, delivering products in yields of up to 96%. Significant yields were obtained when employing benzyl and primary alkyl halides as reaction substrates. Whereas lower yields were obtained with secondary halides and no products were observed applying adamantyl or aryl halides. This suggests that the reaction follows an SN2 mechanism,4444 Sonawane, A. D.; Sonawane, R. A.; Ninomiya, M.; Koketsu, M.; Dalton Trans. 2021, 50, 12764. [Crossref]
Crossref...
as proposed by the authors.

Scheme 2
Electrochemical synthesis of organochalcogenides

Figure 4
Two compartment electrochemical cell separated by a Nafion® membrane applied for the electrochemical synthesis of organochalcogenides. Reprinted with permission from Navarro and co-workers4343 Ribeiro Neto, P. B.; Santana, S. O.; Levitre, G.; Galdino, D.; Oliveira, J. L.; Ribeiro, R. T.; Barros, M. E. S. B.; Bieber, L. W.; Menezes, P. H.; Navarro, M.; Green Chem. 2016, 18, 657. [Crossref]
Crossref...
(copyright (2016) permission from Royal Society of Chemistry (RSC), all rights reserved)

Introducing fluorine atoms into molecules holds great importance due to the exceptional properties. Performing fluorinations in organic molecules is quite challenging due to fluorine’s high reactivity and the difficulty in achieving selectivity. The reaction conditions frequently involve harsh conditions, and certain fluorinated reagents can be costly. Additionally, environmental concerns are significant, as the traditional synthetic procedure usually need the use of excessive fluorine reagents. However, with technological advances, we are finding better ways to address these challenges. In the last years, Waldvogel and co-workers,4545 Haupt, J. D.; Berger, M.; Waldvogel, S. R.; Org. Lett. 2019, 21, 242. [Crossref]
Crossref...
Fuchigami and co-workers,4646 Sawamura, T.; Kuribayashi, S.; Inagi, S.; Fuchigami, T.; Adv. Synth. Catal. 2010, 352, 2757. [Crossref]
Crossref...
,4747 Sawamura, T.; Kuribayashi, S.; Inagi, S.; Fuchigami, T.; Org. Lett. 2010, 12, 644. [Crossref]
Crossref...
and Wirth and co-workers4848 Winterson, B.; Rennigholtz, T.; Wirth, T.; Chem. Sci. 2021, 12, 9053. [Crossref]
Crossref...
have conducted an electrochemical fluorocyclization of N-allylcarboxamides, fluoro-desulfurization of dithioacetals, and fluorination of α-dicarbonyl compounds utilizing (difluoroiodo)arene (a hypervalent iodine) as fluorine source. Additionally, Lennox and co-workers4949 Doobary, S.; Sedikides, A. T.; Caldora, H. P.; Poole, D. L.; Lennox, A. J. J.; Angew. Chem., Int. Ed. 2020, 59, 1155. [Crossref]
Crossref...
conducted an electrochemical vicinal difluorination of alkenes.

The use of iodine salts as mediators has been widely studied to promote the synthesis of new compounds. Miller and Watkins5050 Miller, L. L.; Watkins, B. F.; J. Am. Chem. Soc. 1976, 98, 1515. [Crossref]
Crossref...
pioneered the utilization of electrolytically generated I+ for iodination reactions in aromatic species, and an extensive mechanistic investigation was carried out. Their findings revealed that the electrochemical oxidation of iodine (I2) in acetonitrile resulted in the production of MeCNI+ as an equivalent species to I+, facilitating subsequent electrochemical reactions with aromatic compounds. In 2006, Yoshida and co-workers5151 Midorikawa, K.; Suga, S.; Yoshida, J.; Chem. Commun. 2006, 36, 3794. [Crossref]
Crossref...
performed a comprehensive study employing CSI-MS, which demonstrated that (MeCN)2I+ serves as the primary species generated through the anodic oxidation of I2 in MeCN, effectively identifying the precise equivalent species of I+. The application of iodine salts as mediators has been widely studied to promote the synthesis of new compounds.5252 Möckel, R.; Babaoglu, E.; Hilt, G.; Chem. - Eur. J. 2018, 24, 15781. [Crossref]
Crossref...
,5353 Liu, K.; Song, C.; Lei, A.; Org. Biomol. Chem. 2018, 16, 2375. [Crossref]
Crossref...
Considering this, in 2019, Mendes and co-workers5454 Meirinho, A. G.; Pereira, V. F.; Martins, G. M.; Saba, S.; Rafique, J.; Braga, A. L.; Mendes, S. R.; Eur. J. Org. Chem. 2019, 2019, 6465. [Crossref]
Crossref...
reported a synthesis of selenylated arenes through electrochemically oxidative C(sp2)–H bond selenylation mediated by iodine salt (Scheme 3). This regioselective reaction offers good yields using Pt electrodes and KI as an electrolyte and mediator. The method is scalable, operates under atmospheric conditions, and utilizes non-toxic and easily handleable reagents. The electrochemical selenylation of 2-naphthol, N,N-dimethylaniline, and 1H-indole derivatives resulted in yields of up to 97%. However, experiments involving 1,3,5-trimethoxybenzene and 5,6,7,8-tetrahydronaphthalen-2-ol derivatives were less efficient, yielding products in 41 and 44%, respectively. Additionally, naphthylamine derivatives were completely suppressed.

Scheme 3
Electrochemical oxidative C(sp2)–H bond selenylation of activated arenes

Two reaction pathways were considered (Scheme 4). In pathway A, the reaction initiates with the anodic oxidation of the iodide ion, resulting in the formation of the iodonium ion. Subsequently, electrophilic substitution occurs, which is quickly captured by the diselenide, leading to the formation of the final product and iodine. In pathway B, the reaction begins with the anodic oxidation of the iodide ion, giving rise to the iodine mediator. The diselenide undergoes oxidation, forming RSeI, which then attacks the activated arenes to generate the final product followed by the elimination of a proton and iodide ion.

Scheme 4
Proposed mechanism for selenation of (hetero)arenes

In 2019, Noël and co-workers5555 Laudadio, G.; Bartolomeu, A. A.; Verwijlen, L. M. H. M.; Cao, Y.; de Oliveira, K. T.; Noël, T.; J. Am. Chem. Soc. 2019, 141, 11832. [Crossref]
Crossref...
developed an electrochemical method for synthesizing sulfonyl fluorides. The method involved the use of thiols or disulfides and utilized KF as the source of fluoride (Scheme 5). This protocol eliminates the need for external oxidants or catalysts, and the reaction was suitable in a wide range of substrates, encompassing alkyl, benzyl, aryl, and heteroaryl thiols or disulfides. Kinetic experiments were conducted to investigate the anodic oxidation of 4-(trifluoromethyl)thiophenol and its conversion to the corresponding disulfide. Upon the formation of the disulfide, it was immediately consumed, leading to the subsequent formation of the corresponding sulfonyl fluoride. The use of KF as a fluoride source ensures a safe and cost-effective transformation, thereby increasing the economic viability of the process and reducing potential hazards.

Scheme 5
Synthesis of sulfonyl fluoride through electrochemical oxidative coupling of thiols and potassium fluoride

In the following year, Noël and co-workers5656 Cao, Y.; Adriaenssens, B.; Bartolomeu, A. A.; Laudadio, G.; de Oliveira, K. T.; Noël, T.; J. Flow Chem. 2020, 10, 191. [Crossref]
Crossref...
made a further progress with the previously discussed methodology by proposing a continuous-flow procedure for achieving the electrochemical oxidative coupling of thiols and KF to synthesize sulfonyl fluorides (Scheme 6). This flow method significantly reduced reaction times to 5 min, compared to 6-36 h in batch, and demonstrated scalability for this approach (2.0 to 10 mmol scale). They achieved yields of up to 85% by applying 0.1 M of thiol, 0.6 M of pyridine, 0.5 M of KF, 3.30 V, 1 M of HCl/MeCN (1:1 v/v), at a total flow rate of 150 μL min-1 and a residence time of 5 min, using graphite and stainless steel as electrodes. Furthermore, the flow-based sulfonyl fluoride synthesis was applied for telescoping without the need for interim purification.

Scheme 6
Continuous-flow procedure for the electrochemical oxidative coupling of thiols and KF to synthesize sulfonyl fluorides

The authors emphasize that the experiments were conducted in a homemade electrochemical flow reactor, as shown in Figure 5. Furthermore, the synthesis of sulfonyl fluoride in a continuous flow system can easily be telescoped and integrated into a subsequent SuFEx reaction (sulfur fluoride exchange),5757 Chrominski, M.; Ziemkiewicz, K.; Kowalska, J.; Jemielity, J.; Org. Lett. 2022, 24, 4977. [Crossref]
Crossref...
eliminating the necessity for intermediate purification.

Figure 5
Homemade electrochemical flow reactor used for the sulfonyl fluoride synthesis through oxidative coupling of thiols and potassium fluoride. Reprinted with permission from Noël and co-workers5858 Laudadio, G.; de Smet, W.; Struik, L.; Cao, Y.; Noël, T.; J. Flow Chem. 2018, 8, 157. [Crossref]
Crossref...
(copyright (2018) permission from Springer, all rights reserved)

In 2020, da Silva Júnior and co-workers5959 Kharma, A.; Jacob, C.; Bozzi, I. A. O.; Jardim, G. A. M.; Braga, A. L.; Salomão, K.; Gatto, C. C.; Silva, M. F. S.; Pessoa, C.; Stangier, M.; Ackermann, L.; da Silva Júnior, E. N.; Eur. J. Org. Chem. 2020, 2020, 4474. [Crossref]
Crossref...
described an electrochemical approach to introduce selenium into naphthoquinone compounds through an anodic oxidative selenation/cyclization process (Scheme 7). Several quinone-hybrid molecules were obtained using an undivided cell, platinum electrodes, nBu4NPF6 as electrolite and MeCN as solvent, at room temperature. These quinone-hybrid molecules exhibited interesting biological activity against five different cancer cell lines and Trypanosoma cruzi, the causative agent of Chagas disease. For the investigation of the reaction mechanism, the authors relied on the literature and analyses of cyclic voltammetry. The results suggested that the [(PhSe)2]+ radical cation generated (probably a dicationic tetramer)6060 Kunai, A.; Harada, J.; Izumi, J.; Tachihara, H.; Sasaki, K.; Electrochim. Acta 1983, 28, 1361. [Crossref]
Crossref...
,6161 Ludvík, J.; Nygård, B.; J. Electroanal. Chem. 1997, 423, 1. [Crossref]
Crossref...
,6262 Mueller, B.; Poleschner, H.; Seppelt, K.; Dalton Trans. 2008, 33, 4424. [Crossref]
Crossref...
undergoes a carbophilic reaction of the selenium dication with lapachol, forming a cationic intermediate, which undergoes nucleophilic cyclization, delivering products. In 2023, da Silva Júnior and co-workers6363 Diogo, E. B. T.; Delolo, F. G.; Graça, G. A. P.; Paz, E. R. S.; Bozzi, I. A. O.; Diniz, R.; Passos, J. P.; Matencio, T.; Soares, L. K.; Alves, D.; Costa, P. M. S.; Pessoa, C.; Pereira, C. L. M.; Ackermann, L.; da Silva Júnior, E. N.; Eur. J. Org. Chem. 2023, 26, e2023005. [Crossref]
Crossref...
reported a variation of the previous method, applying disulfides to sulphenylation/cyclization of quinones, obtaining yields up to 93% under the same reaction conditions.

Scheme 7
Synthesis of selenil-naphthoquinone compounds through an anodic oxidative selenation/cyclization process

Recently, Menezes and co-workers6464 Vicente, D. A.; Galdino, D.; Navarro, M.; Menezes, P. H.; Green Chem. 2020, 22, 5262. [Crossref]
Crossref...
proposed an intriguing methodology for a solvent-free electrochemical synthesis of sulfonamides using a graphite powder macroelectrode (Scheme 8). An electrolytic cell was developed (Figure 6), and the electro-oxidation was conducted on the graphite powder macroelectrode utilizing an aqueous electrolyte within a cavity cell. The active electrode (300 mg of graphite powder) was meticulously measured, combined with sodium p-toluenesulfinate (1.0 mmol), and introduced into the cavity created by the graphite stick and Teflon® base. A correspondingly-sized aluminum rod was then positioned atop the mixture, and a weight of 2.5 kg was applied for 10 min. Subsequently, benzylamine (0.5 mmol) was incrementally added (with an automatic pipette) to the compressed graphite powder and p-toluenesulfinate. To prevent any potential dispersion of the compacted substance into the electrolytic solution, a filter paper with small holes was added above the compressed material. The electrolysis was carried out in a constant current of 20 mA for 1 h. Aqueous solution of LiClO4 (0.1 M) was used as the supporting electrolyte. As a mechanistic pathway, electro-synthetic conditions favor the oxidation of sulfinate (with an oxidation potential of 1.04 V). The resulting radical reacts with the amino group, forming an intermediate that undergoes hydrogen loss, delivering the final product.

Scheme 8
Solvent-free electrochemical synthesis of sulfonamides

Figure 6
Representative illustration of an electrochemical cavity cell. Reprinted with permission from Menezes and co-workers6464 Vicente, D. A.; Galdino, D.; Navarro, M.; Menezes, P. H.; Green Chem. 2020, 22, 5262. [Crossref]
Crossref...
(copyright (2020) permission from Royal Society of Chemistry (RSC), all rights reserved)

Although existing literature for synthesizing selenide compounds through oxidative C–Se bond formation via C(sp2)–H bond activation in arenes, many of these methods involve the use of external oxidants or catalysts.6565 Silveira, C. C.; Mendes, S. R.; Wolf, L.; Martins, G. M.; Tetrahedron Lett. 2010, 51, 2014. [Crossref]
Crossref...
,6666 Silveira, C. C.; Mendes, S. R.; Wolf, L.; Martins, G. M.; von Mühlen, L.; Tetrahedron 2012, 68, 10464. [Crossref]
Crossref...
,6767 Azeredo, J. B.; Godoi, M.; Martins, G. M.; Silveira, C. C.; Braga, A. L.; J. Org. Chem. 2014, 79, 4125. [Crossref]
Crossref...
,6868 Ricordi, V. G.; Thurow, S.; Penteado, F.; Schumacher, R. F.; Perin, G.; Lenardão, E. J.; Alves, D.; Adv. Synth. Catal. 2015, 357, 933. [Crossref]
Crossref...
,6969 Silva, L. T.; Azeredo, J. B.; Saba, S.; Rafique, J.; Bortoluzzi, A. J.; Braga, A. L.; Eur. J. Org. Chem. 2017, 2017, 4740. [Crossref]
Crossref...
,7070 Rafique, J.; Saba, S.; Franco, M. S.; Bettanin, L.; Schneider, A. R.; Silva, L. T.; Braga, A. L.; Chem. - Eur. J. 2018, 24, 4173. [Crossref]
Crossref...
,7171 Saba, S.; Rafique, J.; Franco, M. S.; Schneider, A. R.; Espíndola, L.; Silva, D. O.; Braga, A. L.; Org. Biomol. Chem. 2018, 16, 880. [Crossref]
Crossref...
,7272 Belladona, A. L.; Cervo, R.; Alves, D.; Barcellos, T.; Cargnelutti, R.; Schumacher, R. F.; Tetrahedron Lett. 2020, 61, 152035. [Crossref]
Crossref...
In 2020, Braga and co-workers7373 Scheide, M. R.; Schneider, A. R.; Jardim, G. A. M.; Martins, G. M.; Durigon, D. C.; Saba, S.; Rafique, J.; Braga, A. L.; Org. Biomol. Chem. 2020, 18, 4916. [Crossref]
Crossref...
proposed an intramolecular electrochemical oxidative oxyseleno-cyclization of allylnaphthol and allylphenol derivatives without external oxidants (Scheme 9). This method demonstrates remarkable compatibility with diverse functional groups, resulting in the synthesis of a range of dihydrofurans containing organoselenium moieties. The successful conduction of a large-scale experiment was achieved, and a scope with total of 20 selenyl-dihydrofurans derivatives were obtained with yields of up to 99% using 0.2 equiv. of nBu4NClO4 as an electrolyte, MeCN as a solvent, and Pt working electrodes in an undivided cell. Interestingly, the reaction failed to deliver products when performed with KI salts. Considering the results of the control reactions, such as the complete inhibition of transformation with a radical scavenger 2,2,6,6-tetramethylpiperidine-N-oxyl (TEMPO), the authors believe that the reaction mechanism follows a radical pathway, albeit not ruling out a route involving the phenyl selenium cation. Additionally, several selenylated products have demonstrated comparable inhibitory activity against the enzyme acetylcholinesterase (AChE) using galanthamine as model, a commercially available drug applied as Alzheimer’s disease treatment.

Scheme 9
Selenofunctionalization of allylnaphthol/phenol derivatives

In 2021, Mendes and co-workers7474 Lazzaris, M. J.; Martins, G. M.; Xavier, F. R.; Braga, A. L.; Mendes, S. R.; Eur. J. Org. Chem. 2021, 2021, 4411. [Crossref]
Crossref...
reported a protocol to achieve a regioselective electrochemical oxidative C(sp2)–H selenylation of resveratrol (Scheme 10), leading to the formation of unprecedented mono- and bisselenylated compounds involving a galvanostatic electrolysis utilizing platinum and graphite electrodes in an undivided cell. The method exhibited high efficiency under oxidant-free, base-free, and transition metal-free conditions, all performed in an open system at ambient temperature. Notably, this strategy required a sub-stoichiometric quantity of NaI, serving as both an electrolyte and mediator. The method demonstrated good functional group compatibility and proved effective for allylic diselenide derivatives, usually problematic substrates in this type of transformation, delivering products in yields of up to 73%.

Scheme 10
Electrochemical oxidative C(sp2)–H bond selenylation of resveratrol

Aziridines find a versatile application in pharmaceutical, agrochemical, and material sciences due to their unique three-membered ring structure.7575 Singh, G. S.; Mini-Rev. Med. Chem. 2016, 16, 892. [Crossref]
Crossref...
,7676 Chaudhari, P. J.; Bari, S. B.; Surana, S. J.; Shirkhedkar, A. A.; Bonde, C. G.; Khadse, S. C.; Ugale, V. G.; Nagar, A. A.; Cheke, R. S.; ACS Omega 2022, 7, 17270. [Crossref]
Crossref...
The most common approach involves the ring expansion of epoxides using nucleophilic attack by amines under the influence of Lewis acids or metal catalysts.7777 Dequina, H. J.; Jones, C. L.; Schomaker, J. M.; Chem 2023, 9, 1658. [Crossref]
Crossref...
Noteworthy advances have been made in electrochemical synthesis methodologies. For instance, Siu and Yudin7878 Siu, T.; Yudin, A. K.; J. Am. Chem. Soc. 2002, 124, 530. [Crossref]
Crossref...
developed an electrochemical aziridination using N-amino-phthalimide as the source of electrophilic nitrogen, applying a divided-cell. A variant of this protocol was achieved by Little and co-workers7979 Chen, J.; Yan, W. Q.; LamCrossref]
Crossref...
through an iodide-mediated electrocatalytic approach. Additionally, Cheng and co-workers8080 Li, J.; Huang, W.; Chen, J.; He, L.; Cheng, X.; Li, G.; Angew. Chem., Int. Ed. 2018, 57, 5695. [Crossref]
Crossref...
presented an electrochemical strategy utilizing trifluoromethylated sulfamates as coupling partners. Recently, Wickens and co-workers8181 Holst, D. E.; Wang, D. J.; Kim, M. J.; Guzei, I. A.; Wickens, Z. K.; Nature 2021, 596, 74. [Crossref]
Crossref...
proposed an aziridine synthesis by coupling amines and alkenes via an electrogenerated dication. In 2021, Noël and co-workers8282 Ošeka, M.; Laudadio, G.; van Leest, N. P.; Dyga, M.; Bartolomeu, A. A.; Gooßen, L. J.; de Bruin, B.; de Oliveira, K. T.; Noël, T.; Chem 2021, 7, 255. [Crossref]
Crossref...
developed an electrochemical aziridination of internal alkenes with primary amines (Scheme 11). This reaction was conducted in a homemade electrochemical flow reactor (Figure 5), and a wide reaction scope was evaluated (28 examples), affording products in yields of up to 93%, in a residence time of 5 min. The reaction was conducted applying alkene (1.0 mmol), amine (5.0 equiv.), 1,1,1,3,3,3-hexafluoro-2-propanol (HFIP, 5.0 equiv), γ-terpinene (0.5 equiv.), MeCN as solvent (0.1 M), a C anode and Fe cathode, and a current density between 2.5 and 5.0 mA. The cathode played a crucial role in generating hydrogen gas, which could be employed in a subsequent reactor to facilitate the reduction of aziridine, leading to the desired hydroaminated product. Through mechanistic investigations and theoretical calculations using density functional theory (DFT), a reaction mechanism was proposed, suggesting that the alkene undergoes anodic oxidation before to follow for a coupling reaction with the amine. This approach demonstrated the possibility of directly converting olefins and primary alkyl amines into aziridines, as evidenced by the broad applicability demonstrated in 28 different examples.

Scheme 11
Continuous-flow procedure for the electrochemical aziridination of internal alkenes with primary amines

Braga and co-workers8383 Doerner, C. V.; Scheide, M. R.; Nicoleti, C. R.; Durigon, D. C.; Idiarte, V. D.; Sousa, M. J. A.; Mendes, S. R.; Saba, S.; Neto, J. S. S.; Martins, G. M.; Rafique, J.; Braga, A. L.; Front. Chem. 2022, 10, 880099. [Crossref]
Crossref...
also developed a regioselective electrochemical synthesis of selenylbenzo[b]furan derivatives via cyclization of 2-alkynylphenols (Scheme 12). By a galvanostatic electrolysis, platinum electrodes in an undivided cell under oxidant-free, base-free, and transition metal-free conditions in an open system at ambient temperature delivered products in up to 98% yield. A gram-scale synthesis was achieved through cyclization of 2-(phenylethynyl)phenol with diphenyl diselenide. The reaction was conducted for 20 h, resulting in a 45% yield of the desired product. Interestingly, the use of 2-[(trimethylsilyl)ethynyl]phenol afforded the bis-selenylated product. According to control reactions, the reaction mechanism suggests a radical pathway, similarly as proposed above.

Scheme 12
Synthesis of selenylbenzo[b]furan derivatives through the cyclization of 2-alkynylphenols

In 2022, de Oliveira and co-workers8484 de Souza, A. A. N.; Bartolomeu, A. A.; Brocksom, T. J.; Noël, T.; de Oliveira, K. T.; J. Org. Chem. 2022, 87, 5856. [Crossref]
Crossref...
investigated an electrochemical synthesis of α-sulfenylated ketones using both batch and continuous flow systems (Schemes 13 and 14). By employing thiophenols/thiols and enol-acetates, and nBu4NBF4 as electrolyte, the authors successfully obtained α-sulfenylated ketone derivatives without requiring additional oxidants or catalysts. The authors emphasized the remarkable tolerance of the Boc-cysteine substrate, significantly enhancing its practical applicability. The potential scalability of the methodology was successfully demonstrated in both batch and continuous flow conditions. Mechanistic insights were also reported while the TEMPO radical scavenger trapped the thiyl radical, and the transformation was completely inhibited when subjected to standard conditions. Similarly, through gas chromatography–mass spectrometry (GC-MS) analysis, the presence of the corresponding disulfide formed via cathodic reduction was observed. The process begins with the formation of the thiyl radical through the electrochemical anodic oxidation of thiol. This radical is then captured by the enol-acetate. Following this initial step, the desired α-sulfenylated ketone is obtained through the cathodic reduction and hydrolysis of the intermediate.

Scheme 13
Electrochemical synthesis of α-sulfenylated ketones

Scheme 14
Electrochemical synthesis of α-sulfenylated ketones in flow conditions

The transformation under continuous flow conditions was carried out using the commercially available electrochemical system from Syrris® (Flux Asia) with a microflow electrocell (9 cm2 surface area, 225 μL volume reactor). The optimal reaction condition was achieved at a residence time of 7.5 min, delivering 1-(phenylthio) propan-2-one in a 83% yield and a productivity of 0.36 g per day (Scheme 14). Overall, the authors described a direct protocol to obtain α-sulfenylated ketones by electrosynthesis using thiols or thiophenols and enol-acetates. With this approach it is possible to obtain significant yields (up to 94%), a broad scope, and scalability.

Thiazolidine-2,4-diones are a relevant class of heterocycles due to their diverse range of biological activities, particularly in the context of developing novel anti-diabetic agents, such as pioglitazone, ciglitazone, lobeglitazone, rosiglitazone, and troglitazone.8585 Jain, V. S.; Vora, D. K.; Ramaa, C. S.; Bioorg. Med. Chem. 2013, 21, 1599. [Crossref]
Crossref...
,8686 Giglio, R. V.; Papanas, N.; Rizvi, A. A.; Ciaccio, M.; Patti, A. M.; Ilias, I.; Pantea Stoian, A.; Sahebkar, A.; Janez, A.; Rizzo, M.; Medicina 2022, 58, 1475. [Crossref]
Crossref...
The synthesis of these derivatives typically involves a condensation reaction between thiazolidine-2,4-dione and aldehydes, followed by a hydrogenation step using either H2/Pd under pressure or an excess of NaBH4.8787 Sharma, V. K.; Barde, A.; Rattan, S.; Synth. Commun. 2021, 51, 57. [Crossref]
Crossref...
However, a significant limitation in many of these synthetic routes is the need of transition metal catalysis, which can represent a problematic issue in the context of the synthesis of active pharmaceutical ingredients (API). Therefore, there is a growing interest in the development of alternative methodologies that are more sustainable, efficient, and environmentally friendly for the synthesis of these crucial compounds.8888 Becker, J.; Manske, C.; Randl, S.; Curr. Opin. Green Sustainaible Chem. 2022, 33, 100562. [Crossref]
Crossref...
The pursuit of transition-metal free protocols, applying greener hydrogen sources and innovative reaction conditions, holds great promise in addressing these challenges and advancing the field of organic synthesis in the context of drug development. In 2023, de Oliveira and co-workers8989 de Castro, P. P.; Martins, G. M.; Gomes, R. B.; Simoso, G. B.; Amarante, G. W.; Brocksom, T. J.; de Oliveira, K. T.; Chem. Commun. 2023, 59, 9404. [Crossref]
Crossref...
have introduced a novel transition-metal free methodology designed for the chemoselective reduction of benzylidene thiazolidine-2,4-diones and analogous heterocycles (Scheme 15). The approach enables the synthesis of a diverse range of reduced derivatives, affording yields of up to 90%. Additionally, the scaled-up reaction (0.5 to 6 mmol) displayed comparable reactivity, yielding 78% for the synthesis of 5-(4-fluorobenzyl)thiazolidine-2,4-dione. One of the highlights of this methodology is its simplicity and safety, as it relies on water as the hydrogen source, making it an attractive and practical option for chemical transformations. In order to demonstrate the practical application and potential of this method, the authors successfully synthesized the antidiabetic API pioglitazone with a notable 81% yield. This is the first hydride and transition-metal free protocol for the synthesis of pioglitazone. The proposed mechanism was supported by control experiments and cyclic voltammetry, providing a basis for the elucidation of the reaction pathway. The reaction begins with the deprotonation of thiazolidine-2,4-dione by sodium carbonate, affording the corresponding anion. This anion undergoes anodic oxidation, generating a radical cation that subsequently reacts with a water molecule, abstracting a hydrogen atom. Next, this intermediate receives an electron at the cathode electrode and, following to second hydrogen abstraction, deliver the desired product.

Scheme 15
Electrochemical reduction of 5-benzylidene thiazolidine-2,4-diones derivatives

Recently, Noël and co-workers9090 Costa, R.; Vega, C.; Regnier, M.; Capaldo, L.; Wesenberg, L.; Lowe, G.; de Oliveira, K.; Noel, T.; Adv. Synth. Catal. 2023, 365, 955. [Crossref]
Crossref...
developed a novel electrochemical technique for the synthesis of aryliminophosphoranes. The key innovation of this method is its utilization in a continuous-flow system, which enhances efficiency and convenience (Scheme 16). A mixture of trifluoroacetic acid (TFA) (50 mol%), PPh3 (3.5 equiv.), nitro(hetero)arenes (1.0 equiv.), and MeCN (0.05 M) as the solvent provided products with yields of up to 92%. The method allowed the synthesis of aryliminophosphoranes without the need for additional supporting electrolytes. The transformation was conducted in a flow electrochemical reactor equipped with two graphite electrodes operated in a galvanostatic cell in a residence time of 5 min, and exhibited tolerance towards functional groups, especially electron-deficient nitroarenes. The experiment was scaled-up over 15.2 h, resulting in 4.1 g of product (79% yield - productivity rate of 6.45 g per day). In addition, the versatility of the aryliminophosphoranes is demonstrated as intermediaries to synthesize anilines, amines and amides. From respective aryliminophosphorane, a telescoped-flow method has been developed applying a tube-in-tube reactor charged with CO2 for in situ production of isocyanate. The isocyanate was then transformed into amide through the coupling with TFA. In contrast to literature, this methodology notably eliminates the need for potentially hazardous reagents, such as azides and molecular bromine, even when conducted at room temperature. Furthermore, it enabled a scale-up reaction without the need for additional electrolyte.

Scheme 16
Continuous-flow procedure for the synthesis of aryliminophosphoranes

In 2023, Brondani and co-workers9191 Voigt, M. A.; Pinheiro, E. B.; Zapp, E.; de Jesus, P. C.; Meier, L.; Curbani, L.; Bulegon Brondani, P.; ChemistrySelect 2023, 8, e202301996. [Crossref]
Crossref...
combined electrochemistry and biocatalysis for the electroreduction of acetophenone by applying lipase stabilization to obtain 1-phenylethanol (Scheme 17). A Sn/ Pb electrode (63 and 37% m/m) was polished and covered with a Nafion film modified using lipase from Thermomyces lanuginosus, an applied potential of –2.0 V using an acetate buffer at pH 5.0 for 4 h, delivering 1-phenylethanol in 87% without forming the corresponding pinacol dimer. The 1-phenylethanol was favored in the presence of lipases from Candida antarctica B (Novozym 435), Rhizomucor miehei (Lipozima RM-IM), and Burkholderia cepacia (PS-D Amano I), but the best results were obtained using Thermomyces lanuginosus (immobilized on methylcyclopentadienyl manganese tricarbonyl (MMT), silica gel and Lipolase 100T). The method stood out for enzyme stability, reproducibility, and the possibility of storing and reusing the modified electrode.

Scheme 17
Electroreduction of acetophenone by applying lipases

Sacrificial electrodes

Among various electrochemical techniques, the use of sacrificial electrodes in organic electrosynthesis have gained increasing interest. These electrodes, typically composed of metals with low oxidation potentials, are employed to release metal ions into the solution during the electrochemical process. These metal ions can act as active species, playing fundamental roles as catalysts and coordinators in synthetic organic reactions. Magnesium, copper, zinc, aluminum, iron, and silver are among the frequently metals employed as sacrificial electrodes in electro-organic transformations. One notable disadvantage of employing sacrificial metals is that the electrode becomes non-recyclable as a result of metal salt formation. However, the benefits of this approach can overcome these disadvantages, as these metals in solution could suppress the overoxidation of substrates, active intermediates, or final products, and can replace a catalyst.

In the early 1990s, review articles already discussed the use of sacrificial electrodes. For example, Chaussard et al.9292 Chaussard, J.; Folest, J. C.; Nedelec, J. Y.; Perichon, J.; Sibille, S.; Troupel, M.; Synthesis 1990, 1990, 369. [Crossref]
Crossref...
described a paper focusing on “The Use of Sacrificial Anodes in Electrochemical Functionalization of Organic Halides.” More recently, several studies have reported different electrosynthesis methods utilizing sacrificial electrodes. These include obtaining silver acetylides from acetylenes using Ag0 as the sacrificial electrode,9393 Mitsudo, K.; Shiraga, T.; Mizukawa, J.; Suga, S.; Tanaka, H.; Chem. Commun. 2010, 46, 9256. [Crossref]
Crossref...
electrochemically synthesizing copper(I) acetylides with Cu0 as the electrode,9494 Seavill, P. W.; Holt, K. B.; Wilden, J. D.; Green Chem. 2018, 20, 5474. [Crossref]
Crossref...
,9595 Seavill, P. W.; Holt, K. B.; Wilden, J. D.; RSC Adv. 2019, 9, 29300. [Crossref]
Crossref...
employing a sacrificial Cu0 electrode to synthesize copper-N-heterocyclic carbene complexes using a flow reactor,9696 Chapman, M. R.; Shafi, Y. M.; Kapur, N.; Nguyen, B. N.; Willans, C. E.; Chem. Commun. 2015, 51, 1282. [Crossref]
Crossref...
the using of Mg0 or Al0 anodes for an electrochemical Birch reduction,9797 Peters, B. K.; Rodriguez, K. X.; Reisberg, S. H.; Beil, S. B.; Hickey, D. P.; Kawamata, Y.; Collins, M.; Starr, J.; Chen, L.; Udyavara, S.; Klunder, K.; Gorey, T. J.; Anderson, S. L.; Neurock, M.; Minteer, S. D.; Baran, P. S.; Science 2019, 363, 838. [Crossref]
Crossref...
and the electroreducing of triphenylphosphine oxide (TPPO) to triphenylphosphine (TPP).9898 Manabe, S.; Wong, C. M.; Sevov, C. S.; J. Am. Chem. Soc. 2020, 142, 3024. [Crossref]
Crossref...
In 2023, Guo and co-workers9999 Li, Y.; Wen, L.; Guo, W.; Chem. Soc. Rev. 2023, 52, 1168. [Crossref]
Crossref...
have highlighted this growing interest of the scientific community in a review article entitled “A Guide to Organic Electroreduction Using sacrificial Anodes.”

Several methods for obtaining organoboron compounds are described in the literature. However, expensive and toxic catalysts are often employed, such as palladium, copper, nickel, iron, iridium salts, etc.100100 Presset, M.; Fleury-Brégeot, N.; Oehlrich, D.; Rombouts, F.; Molander, G. A.; J. Org. Chem. 2013, 78, 4615. [Crossref]
Crossref...
,101101 Molander, G. A.; Trice, S. L. J.; Kennedy, S. M.; Org. Lett. 2012, 14, 4814. [Crossref]
Crossref...
,102102 Molander, G. A.; Cavalcanti, L. N.; García-García, C.; J. Org. Chem. 2013, 78, 6427. [Crossref]
Crossref...
,103103 Obligacion, J. V.; Chirik, P. J.; Org. Lett. 2013, 15, 2680. [Crossref]
Crossref...
,104104 Joliton, A.; Carreira, E. M.; Org. Lett. 2013, 15, 5147. [Crossref]
Crossref...
In 2014, Menezes and co-workers105105 Nascimento, W.; Oliveira, J.; Freitas, J.; Navarro, M.; Menezes, P.; Synthesis 2014, 46, 2579. [Crossref]
Crossref...
proposed an electrochemical synthesis of potassium aryltrifluoroborates from the corresponding aryl halides (1.0 equiv.), triisopropyl borate (2.0 equiv.), using THF as the solvent within an inert atmosphere at temperatures ranging from 0 to 25 °C, applying Mg0 as the sacrificial anodic electrode (Scheme 18). The methodology was developed with demonstration of a broad group tolerance, while the electronic effects of substituents in the aromatic bromide minimally affect the transformation, delivering products in yields of up to 70%. An undivided cell was employed, utilizing sacrificial Mg0 as the anode and nickel foam as the cathode. A 30 mA constant current was applied until the full consumption of the starting material (1.5 to 6 h).

Scheme 18
Electrochemical synthesis of potassium aryltrifluoroborates using Mg0 as sacrificial electrode

In 2023, Mendes and co-workers106106 Santos, W. A. B.; de Castro, P. P.; Xavier, F. R.; Braga, A. L.; Martins, G. M.; Mendes, S. R.; Synthesis 2023, 55, 2985. [Crossref]
Crossref...
developed an electrochemical synthesis of flavanones through the intramolecular oxa-Michael addition of hydroxychalcones, applying a sacrificial silver electrode (Scheme 19). The corresponding flavanones were obtained in up to 95% yield by using silver electrodes at the anode, graphite at the cathode, nBu4NPF6 (1.0 equiv.) as the electrolyte, and methanol as the solvent. Cyclic voltammetry and theoretical studies guided the authors to propose a plausible mechanism. The reaction involves the cathodic reduction of methanol, providing a methoxide anion and hydrogen gas. Concurrently, the anodic formation of silver(I) played a vital role in maintaining reaction charge balance. After a 6-endo-trig cyclization, the solvent-mediated proton transfers lead to the desired product. The method demonstrated promising prospects for silver-catalyzed transformations. Additionally, this methodology revealed new synthetic possibilities, raising intriguing questions about the possibility of replacing expensive metal catalysts with sacrifice electrodes. This could make chemical transformations more convenient and environmentally friendly due to the reduction of toxicity and elimination of by-products.

Scheme 19
Electrosynthesis of flavanones via oxa-Michael addition using Ag0 as sacrificial electrode

PERSPECTIVES OF ORGANIC ELECTROSYNTHESIS IN BRAZIL

Organic electrosynthesis has gained worldwide prominence as an alternative to conventional synthetic methods. In recent years, there has been a growing interest and adoption of this innovative approach by researchers and the chemical industry. In this context, various future perspectives emerge, pointing towards a promising and transformative landscape in the field of organic synthesis.

One of the most notable perspectives is the growing emphasis on environmental sustainability. Organic electrosynthesis offers numerous advantages in terms of toxic waste reduction and lower environmental impact. When operating in softer reaction conditions and using electrodes such as catalysts, this approach aligns perfectly with the worldwide sustainable development goals. With the increasing pressure for more environmentally friendly and responsible practices, it is expected that electrosynthesis finds wide application in various sectors of the national chemical industry. Brazil has a robust scientific and technological community, capable of exploiting and improving efficiency in organic reactions. The search for new reaction conditions as well as the development of different electrodes, reactors and reaction cells promises to open new horizons in terms of selectivity and reaction paths in electrochemical reactions, enabling the synthesis of valuable products.

A critical factor for the consolidation of electrosynthesis here in Brazil is the investment in the formation of specialized human resources. Training scientists and students in organic electrochemistry is crucial to promote research and innovation in this field. The creation of postgraduate programs and the update of courses focused on this subject is fundamental for the development of specific expertise and the strengthening of this research line in Brazilian universities and research institutions.

In addition, it is important to highlight the fundamental role of international collaborations. Knowledge exchange with researchers from other countries can provide access to advanced technologies, resources and innovative perspectives. The establishment of partnerships with world renowned research groups can catalyze the fast advance of electrosynthesis in Brazil, providing technical knowledge and accelerating the incorporation of new technologies.

CONCLUSIONS

It has become evident that exploring the field of organic electrosynthesis in Brazil is essential to follow the global advances in this domain. The utilization of electrochemical methods for organic synthesis presents several advantages over traditional synthetic approaches. Notably, the eco-friendly and sustainable nature of electrosynthesis offers a promising alternative to conventional methods, reducing the reliance on toxic reagents and minimizing the generation of harmful waste. Furthermore, the ability to conduct reactions under milder conditions enhances the overall green and sustainable character of electrosynthesis, making it an attractive prospect for modern chemical synthesis.

Given the highlighted advantages and potential benefits, it is imperative for Brazil to continue exploring and investing in the field of organic electrosynthesis to remain at the forefront of scientific development. By addressing the challenges and limitations identified in this review, researchers can pave the way for further improvements and breakthroughs in this area. Additionally, identifying future prospects and potential directions outlined in this review will provide a valuable tutorial for guiding research efforts and driving innovation in organic electrosynthesis within the Brazilian scientific community. Ultimately, promoting advances in this field will not only contribute to the nation’s scientific progress but also contribute to sustainable and environmentally conscious chemical synthesis practices on a global scale.

At the end, this review has shown that we still have a relatively small number of scientific publications in this area compared to other countries around the world. Given the promising potential and advantages of electrosynthesis in relation to traditional synthetic methods, it is crucial to invest in this research field to strengthen our presence in the global scientific scenario. By expanding our efforts and resources, we can boost not only Brazilian science and technology, but also contribute significantly to the search for more sustainable and efficient chemistry solutions. We are convinced that by encouraging and supporting our researchers, we can accelerate the evolution of electrosynthesis in our country and consolidate our role as protagonists in the advancement of this important scientific area.

ACKNOWLEDGMENTS

The authors would like to thank the Fundação de Amparo à Pesquisa do Estado de São Paulo - FAPESP (grants 2022/00074-3 (G. M. M.) and 2023/04020-8 (K. T. O.)) as well as the Conselho Nacional de Pesquisa - CNPq (K. T. O. research fellowship 303333/2022-7).

REFERÊNCIAS

  • 1
    Vogt, H.; J. Electrochem. Soc. 1981, 128, 29C. [Crossref]
    » Crossref
  • 2
    Yan, M.; Kawamata, Y.; Baran, P. S.; Chem. Rev. 2017, 117, 13230. [Crossref]
    » Crossref
  • 3
    Shatskiy, A.; Lundberg, H.; Kärkäs, M. D.; Chemelectrochem 2019, 6, 4067. [Crossref]
    » Crossref
  • 4
    Hilt, G.; ChemElectroChem 2020, 7, 395. [Crossref]
    » Crossref
  • 5
    Pollok, D.; Waldvogel, S. R.; Chem. Sci. 2020, 11, 12386. [Crossref]
    » Crossref
  • 6
    Ali, W.; Benedetti, R.; Handzlik, J.; Zwergel, C.; Battistelli, C.; Drug Discovery Today 2021, 26, 256. [Crossref]
    » Crossref
  • 7
    Li, Y.; Dana, S.; Ackermann, L.; Curr. Opin. Electrochem. 2023, 40, 101312. [Crossref]
    » Crossref
  • 8
    Pernaut, J. M.; Matencio, T.; Quim. Nova 1999, 22, 899. [Crossref]
    » Crossref
  • 9
    Avaca, L. A.; Tokoro, R.; Quim. Nova 2002, 25, 25. [Crossref]
    » Crossref
  • 10
    de Luca, M. A.; de Luca, S. J.; Santana, M. A.; Quim. Nova 2003, 26, 420. [Crossref]
    » Crossref
  • 11
    Gandra, P. G.; Alves, A. A.; de Macedo, D. V.; Kubota, L. T.; Quim. Nova 2004, 27, 980. [Crossref]
    » Crossref
  • 12
    Batista, E. C.; de Oliveira, R. T. S.; Ferreira, R. Q.; Miwa, D.; dos Santos, M. C.; Quim. Nova 2011, 34, 1517. [Crossref]
    » Crossref
  • 13
    Zanoni, M.; Borges, A.; Benedetti, A.; Yamanaka, H.; Sotomayor, M. P.; Bessegato, G.; Stradiotto, N.; Zanta, C. L.; Andrade, A.; Quim. Nova 2017, 40, 663. [Crossref]
    » Crossref
  • 14
    Wiebe, A.; Gieshoff, T.; Möhle, S.; Rodrigo, E.; Zirbes, M.; Waldvogel, S. R.; Angew. Chem., Int. Ed. 2018, 57, 5594. [Crossref]
    » Crossref
  • 15
    Beil, S. B.; Pollok, D.; Waldvogel, S. R.; Angew. Chem., Int. Ed. 2021, 60, 14750. [Crossref]
    » Crossref
  • 16
    Frontana-Uribe, B. A.; Little, R. D.; Ibanez, J. G.; Palma, A.; Vasquez-Medrano, R.; Green Chem. 2010, 12, 2099. [Crossref]
    » Crossref
  • 17
    Meyer, T. H.; Choi, I.; Tian, C.; Ackermann, L.; Chem 2020, 6, 2484. [Crossref]
    » Crossref
  • 18
    Yuan, Y.; Lei, A.; Nat. Commun. 2020, 11, 802. [Crossref]
    » Crossref
  • 19
    Noël, T.; Cao, Y.; Laudadio, G.; Acc. Chem. Res. 2019, 52, 2858. [Crossref]
    » Crossref
  • 20
    Capaldo, L.; Wen, Z.; Noël, T.; Chem. Sci. 2023, 14, 4230. [Crossref]
    » Crossref
  • 21
    Cembellín, S.; Batanero, B.; Chem. Rec. 2021, 21, 2453. [Crossref]
    » Crossref
  • 22
    Fuchigami, T.; Atobe, M.; Inagi, S.; Fundamentals and Applications of Organic Electrochemistry, 1st ed.; John Wiley & Sons: Chichester, 2014.
  • 23
    Hammerich, B. S. O.; Organic Electrochemistry, 5th ed.; CRC Press: Boca Raton, 2015. [Crossref]
    » Crossref
  • 24
    Lodh, J.; Paul, S.; Sun, H.; Song, L.; Schöfberger, W.; Roy, S.; Front. Chem. 2023, 10, 1. [Crossref]
    » Crossref
  • 25
    Yosida, J. In Encyclopedia of Applied Electrochemistry; Kreysa, G.; Ota, K.; Savinell, R. F., eds.; Springer: New York, 2014, p. 386-392.
  • 26
    Roscher, J.; Holze, R.; Encyclopedia 2023, 3, 478. [Crossref]
    » Crossref
  • 27
    Couper, A. M.; Pletcher, D.; Walsh, F. C.; Chem. Rev. 1990, 90, 837. [Crossref]
    » Crossref
  • 28
    Heard, D. M.; Lennox, A. J. J.; Angew. Chem. 2020, 132, 19026. [Crossref]
    » Crossref
  • 29
    Pletcher, D.; Green, R. A.; Brown, R. C. D.; Chem. Rev. 2018, 118, 4573. [Crossref]
    » Crossref
  • 30
    Maljuric, S.; Jud, W.; Kappe, C. O.; Cantillo, D.; J. Flow Chem. 2020, 10, 181. [Crossref]
    » Crossref
  • 31
    Green, R.; Brown, R.; Pletcher, D.; J. Flow Chem. 2015, 5, 31. [Crossref]
    » Crossref
  • 32
    Scheide, M. R.; Nicoleti, C. R.; Martins, G. M.; Braga, A. L.; Org. Biomol. Chem. 2021, 19, 2578. [Crossref]
    » Crossref
  • 33
    Martins, G. M.; Zimmer, G. C.; Mendes, S. R.; Ahmed, N.; Green Chem. 2020, 22, 4849. [Crossref]
    » Crossref
  • 34
    Sbei, N.; Martins, G. M.; Shirinfar, B.; Ahmed, N.; Chem. Rec. 2020, 20, 1530. [Crossref]
    » Crossref
  • 35
    Martins, G. M.; Shirinfar, B.; Hardwick, T.; Ahmed, N.; ChemElectroChem 2019, 6, 1300. [Crossref]
    » Crossref
  • 36
    Martins, G. M.; Shirinfar, B.; Hardwick, T.; Murtaza, A.; Ahmed, N.; Catal. Sci. Technol. 2019, 9, 5868. [Crossref]
    » Crossref
  • 37
    Martins, G. M.; Meirinho, A. G.; Ahmed, N.; Braga, A. L.; Mendes, S. R.; ChemElectroChem 2019, 6, 5928. [Crossref]
    » Crossref
  • 38
    Kisukuri, C. M.; Fernandes, V. A.; Delgado, J. A. C.; Häring, A. P.; Paixão, M. W.; Waldvogel, S. R.; Chem. Rec. 2021, 21, 2502. [Crossref]
    » Crossref
  • 39
    Martins, G. M.; Sbei, N.; Zimmer, G. C.; Ahmed, N.; Electrocatalysis and Electrocatalysts for a Cleaner Environment - Fundamentals and Applications, 1st ed.; IntechOpen: London, 2022, ch. 6. [Crossref]
    » Crossref
  • 40
    Rotruck, J. T.; Pope, A. L.; Ganther, H. E.; Swanson, A. B.; Hafeman, D. G.; Hoekstra, W. G.; Science 1973, 179, 588. [Crossref]
    » Crossref
  • 41
    Alberto, E. E.; do Nascimento, V.; Braga, A. L.; J. Braz. Chem. Soc. 2010, 21, 2032. [Crossref]
    » Crossref
  • 42
    Martins, G. M.; Mendes, S. R. In Organoselenium Chemistry; Hanu, B. C.; Banerjee, B., eds.; De Gruyter: Boston, 2020, ch. 7. [Link] accessed in April 2024
    » Link
  • 43
    Ribeiro Neto, P. B.; Santana, S. O.; Levitre, G.; Galdino, D.; Oliveira, J. L.; Ribeiro, R. T.; Barros, M. E. S. B.; Bieber, L. W.; Menezes, P. H.; Navarro, M.; Green Chem. 2016, 18, 657. [Crossref]
    » Crossref
  • 44
    Sonawane, A. D.; Sonawane, R. A.; Ninomiya, M.; Koketsu, M.; Dalton Trans. 2021, 50, 12764. [Crossref]
    » Crossref
  • 45
    Haupt, J. D.; Berger, M.; Waldvogel, S. R.; Org. Lett. 2019, 21, 242. [Crossref]
    » Crossref
  • 46
    Sawamura, T.; Kuribayashi, S.; Inagi, S.; Fuchigami, T.; Adv. Synth. Catal. 2010, 352, 2757. [Crossref]
    » Crossref
  • 47
    Sawamura, T.; Kuribayashi, S.; Inagi, S.; Fuchigami, T.; Org. Lett. 2010, 12, 644. [Crossref]
    » Crossref
  • 48
    Winterson, B.; Rennigholtz, T.; Wirth, T.; Chem. Sci. 2021, 12, 9053. [Crossref]
    » Crossref
  • 49
    Doobary, S.; Sedikides, A. T.; Caldora, H. P.; Poole, D. L.; Lennox, A. J. J.; Angew. Chem., Int. Ed. 2020, 59, 1155. [Crossref]
    » Crossref
  • 50
    Miller, L. L.; Watkins, B. F.; J. Am. Chem. Soc. 1976, 98, 1515. [Crossref]
    » Crossref
  • 51
    Midorikawa, K.; Suga, S.; Yoshida, J.; Chem. Commun. 2006, 36, 3794. [Crossref]
    » Crossref
  • 52
    Möckel, R.; Babaoglu, E.; Hilt, G.; Chem. - Eur. J. 2018, 24, 15781. [Crossref]
    » Crossref
  • 53
    Liu, K.; Song, C.; Lei, A.; Org. Biomol. Chem. 2018, 16, 2375. [Crossref]
    » Crossref
  • 54
    Meirinho, A. G.; Pereira, V. F.; Martins, G. M.; Saba, S.; Rafique, J.; Braga, A. L.; Mendes, S. R.; Eur. J. Org. Chem. 2019, 2019, 6465. [Crossref]
    » Crossref
  • 55
    Laudadio, G.; Bartolomeu, A. A.; Verwijlen, L. M. H. M.; Cao, Y.; de Oliveira, K. T.; Noël, T.; J. Am. Chem. Soc. 2019, 141, 11832. [Crossref]
    » Crossref
  • 56
    Cao, Y.; Adriaenssens, B.; Bartolomeu, A. A.; Laudadio, G.; de Oliveira, K. T.; Noël, T.; J. Flow Chem. 2020, 10, 191. [Crossref]
    » Crossref
  • 57
    Chrominski, M.; Ziemkiewicz, K.; Kowalska, J.; Jemielity, J.; Org. Lett. 2022, 24, 4977. [Crossref]
    » Crossref
  • 58
    Laudadio, G.; de Smet, W.; Struik, L.; Cao, Y.; Noël, T.; J. Flow Chem. 2018, 8, 157. [Crossref]
    » Crossref
  • 59
    Kharma, A.; Jacob, C.; Bozzi, I. A. O.; Jardim, G. A. M.; Braga, A. L.; Salomão, K.; Gatto, C. C.; Silva, M. F. S.; Pessoa, C.; Stangier, M.; Ackermann, L.; da Silva Júnior, E. N.; Eur. J. Org. Chem. 2020, 2020, 4474. [Crossref]
    » Crossref
  • 60
    Kunai, A.; Harada, J.; Izumi, J.; Tachihara, H.; Sasaki, K.; Electrochim. Acta 1983, 28, 1361. [Crossref]
    » Crossref
  • 61
    Ludvík, J.; Nygård, B.; J. Electroanal. Chem. 1997, 423, 1. [Crossref]
    » Crossref
  • 62
    Mueller, B.; Poleschner, H.; Seppelt, K.; Dalton Trans. 2008, 33, 4424. [Crossref]
    » Crossref
  • 63
    Diogo, E. B. T.; Delolo, F. G.; Graça, G. A. P.; Paz, E. R. S.; Bozzi, I. A. O.; Diniz, R.; Passos, J. P.; Matencio, T.; Soares, L. K.; Alves, D.; Costa, P. M. S.; Pessoa, C.; Pereira, C. L. M.; Ackermann, L.; da Silva Júnior, E. N.; Eur. J. Org. Chem. 2023, 26, e2023005. [Crossref]
    » Crossref
  • 64
    Vicente, D. A.; Galdino, D.; Navarro, M.; Menezes, P. H.; Green Chem. 2020, 22, 5262. [Crossref]
    » Crossref
  • 65
    Silveira, C. C.; Mendes, S. R.; Wolf, L.; Martins, G. M.; Tetrahedron Lett. 2010, 51, 2014. [Crossref]
    » Crossref
  • 66
    Silveira, C. C.; Mendes, S. R.; Wolf, L.; Martins, G. M.; von Mühlen, L.; Tetrahedron 2012, 68, 10464. [Crossref]
    » Crossref
  • 67
    Azeredo, J. B.; Godoi, M.; Martins, G. M.; Silveira, C. C.; Braga, A. L.; J. Org. Chem. 2014, 79, 4125. [Crossref]
    » Crossref
  • 68
    Ricordi, V. G.; Thurow, S.; Penteado, F.; Schumacher, R. F.; Perin, G.; Lenardão, E. J.; Alves, D.; Adv. Synth. Catal. 2015, 357, 933. [Crossref]
    » Crossref
  • 69
    Silva, L. T.; Azeredo, J. B.; Saba, S.; Rafique, J.; Bortoluzzi, A. J.; Braga, A. L.; Eur. J. Org. Chem. 2017, 2017, 4740. [Crossref]
    » Crossref
  • 70
    Rafique, J.; Saba, S.; Franco, M. S.; Bettanin, L.; Schneider, A. R.; Silva, L. T.; Braga, A. L.; Chem. - Eur. J. 2018, 24, 4173. [Crossref]
    » Crossref
  • 71
    Saba, S.; Rafique, J.; Franco, M. S.; Schneider, A. R.; Espíndola, L.; Silva, D. O.; Braga, A. L.; Org. Biomol. Chem. 2018, 16, 880. [Crossref]
    » Crossref
  • 72
    Belladona, A. L.; Cervo, R.; Alves, D.; Barcellos, T.; Cargnelutti, R.; Schumacher, R. F.; Tetrahedron Lett. 2020, 61, 152035. [Crossref]
    » Crossref
  • 73
    Scheide, M. R.; Schneider, A. R.; Jardim, G. A. M.; Martins, G. M.; Durigon, D. C.; Saba, S.; Rafique, J.; Braga, A. L.; Org. Biomol. Chem. 2020, 18, 4916. [Crossref]
    » Crossref
  • 74
    Lazzaris, M. J.; Martins, G. M.; Xavier, F. R.; Braga, A. L.; Mendes, S. R.; Eur. J. Org. Chem. 2021, 2021, 4411. [Crossref]
    » Crossref
  • 75
    Singh, G. S.; Mini-Rev. Med. Chem. 2016, 16, 892. [Crossref]
    » Crossref
  • 76
    Chaudhari, P. J.; Bari, S. B.; Surana, S. J.; Shirkhedkar, A. A.; Bonde, C. G.; Khadse, S. C.; Ugale, V. G.; Nagar, A. A.; Cheke, R. S.; ACS Omega 2022, 7, 17270. [Crossref]
    » Crossref
  • 77
    Dequina, H. J.; Jones, C. L.; Schomaker, J. M.; Chem 2023, 9, 1658. [Crossref]
    » Crossref
  • 78
    Siu, T.; Yudin, A. K.; J. Am. Chem. Soc. 2002, 124, 530. [Crossref]
    » Crossref
  • 79
    Chen, J.; Yan, W. Q.; LamCrossref]
    » Crossref
  • 80
    Li, J.; Huang, W.; Chen, J.; He, L.; Cheng, X.; Li, G.; Angew. Chem., Int. Ed. 2018, 57, 5695. [Crossref]
    » Crossref
  • 81
    Holst, D. E.; Wang, D. J.; Kim, M. J.; Guzei, I. A.; Wickens, Z. K.; Nature 2021, 596, 74. [Crossref]
    » Crossref
  • 82
    Ošeka, M.; Laudadio, G.; van Leest, N. P.; Dyga, M.; Bartolomeu, A. A.; Gooßen, L. J.; de Bruin, B.; de Oliveira, K. T.; Noël, T.; Chem 2021, 7, 255. [Crossref]
    » Crossref
  • 83
    Doerner, C. V.; Scheide, M. R.; Nicoleti, C. R.; Durigon, D. C.; Idiarte, V. D.; Sousa, M. J. A.; Mendes, S. R.; Saba, S.; Neto, J. S. S.; Martins, G. M.; Rafique, J.; Braga, A. L.; Front. Chem. 2022, 10, 880099. [Crossref]
    » Crossref
  • 84
    de Souza, A. A. N.; Bartolomeu, A. A.; Brocksom, T. J.; Noël, T.; de Oliveira, K. T.; J. Org. Chem. 2022, 87, 5856. [Crossref]
    » Crossref
  • 85
    Jain, V. S.; Vora, D. K.; Ramaa, C. S.; Bioorg. Med. Chem. 2013, 21, 1599. [Crossref]
    » Crossref
  • 86
    Giglio, R. V.; Papanas, N.; Rizvi, A. A.; Ciaccio, M.; Patti, A. M.; Ilias, I.; Pantea Stoian, A.; Sahebkar, A.; Janez, A.; Rizzo, M.; Medicina 2022, 58, 1475. [Crossref]
    » Crossref
  • 87
    Sharma, V. K.; Barde, A.; Rattan, S.; Synth. Commun. 2021, 51, 57. [Crossref]
    » Crossref
  • 88
    Becker, J.; Manske, C.; Randl, S.; Curr. Opin. Green Sustainaible Chem. 2022, 33, 100562. [Crossref]
    » Crossref
  • 89
    de Castro, P. P.; Martins, G. M.; Gomes, R. B.; Simoso, G. B.; Amarante, G. W.; Brocksom, T. J.; de Oliveira, K. T.; Chem. Commun. 2023, 59, 9404. [Crossref]
    » Crossref
  • 90
    Costa, R.; Vega, C.; Regnier, M.; Capaldo, L.; Wesenberg, L.; Lowe, G.; de Oliveira, K.; Noel, T.; Adv. Synth. Catal. 2023, 365, 955. [Crossref]
    » Crossref
  • 91
    Voigt, M. A.; Pinheiro, E. B.; Zapp, E.; de Jesus, P. C.; Meier, L.; Curbani, L.; Bulegon Brondani, P.; ChemistrySelect 2023, 8, e202301996. [Crossref]
    » Crossref
  • 92
    Chaussard, J.; Folest, J. C.; Nedelec, J. Y.; Perichon, J.; Sibille, S.; Troupel, M.; Synthesis 1990, 1990, 369. [Crossref]
    » Crossref
  • 93
    Mitsudo, K.; Shiraga, T.; Mizukawa, J.; Suga, S.; Tanaka, H.; Chem. Commun. 2010, 46, 9256. [Crossref]
    » Crossref
  • 94
    Seavill, P. W.; Holt, K. B.; Wilden, J. D.; Green Chem. 2018, 20, 5474. [Crossref]
    » Crossref
  • 95
    Seavill, P. W.; Holt, K. B.; Wilden, J. D.; RSC Adv. 2019, 9, 29300. [Crossref]
    » Crossref
  • 96
    Chapman, M. R.; Shafi, Y. M.; Kapur, N.; Nguyen, B. N.; Willans, C. E.; Chem. Commun. 2015, 51, 1282. [Crossref]
    » Crossref
  • 97
    Peters, B. K.; Rodriguez, K. X.; Reisberg, S. H.; Beil, S. B.; Hickey, D. P.; Kawamata, Y.; Collins, M.; Starr, J.; Chen, L.; Udyavara, S.; Klunder, K.; Gorey, T. J.; Anderson, S. L.; Neurock, M.; Minteer, S. D.; Baran, P. S.; Science 2019, 363, 838. [Crossref]
    » Crossref
  • 98
    Manabe, S.; Wong, C. M.; Sevov, C. S.; J. Am. Chem. Soc. 2020, 142, 3024. [Crossref]
    » Crossref
  • 99
    Li, Y.; Wen, L.; Guo, W.; Chem. Soc. Rev. 2023, 52, 1168. [Crossref]
    » Crossref
  • 100
    Presset, M.; Fleury-Brégeot, N.; Oehlrich, D.; Rombouts, F.; Molander, G. A.; J. Org. Chem. 2013, 78, 4615. [Crossref]
    » Crossref
  • 101
    Molander, G. A.; Trice, S. L. J.; Kennedy, S. M.; Org. Lett. 2012, 14, 4814. [Crossref]
    » Crossref
  • 102
    Molander, G. A.; Cavalcanti, L. N.; García-García, C.; J. Org. Chem. 2013, 78, 6427. [Crossref]
    » Crossref
  • 103
    Obligacion, J. V.; Chirik, P. J.; Org. Lett. 2013, 15, 2680. [Crossref]
    » Crossref
  • 104
    Joliton, A.; Carreira, E. M.; Org. Lett. 2013, 15, 5147. [Crossref]
    » Crossref
  • 105
    Nascimento, W.; Oliveira, J.; Freitas, J.; Navarro, M.; Menezes, P.; Synthesis 2014, 46, 2579. [Crossref]
    » Crossref
  • 106
    Santos, W. A. B.; de Castro, P. P.; Xavier, F. R.; Braga, A. L.; Martins, G. M.; Mendes, S. R.; Synthesis 2023, 55, 2985. [Crossref]
    » Crossref

Datas de Publicação

  • Publicação nesta coleção
    07 Jun 2024
  • Data do Fascículo
    2025

Histórico

  • Recebido
    02 Dez 2023
  • Aceito
    23 Fev 2024
  • Publicado
    26 Abr 2024
Sociedade Brasileira de Química Secretaria Executiva, Av. Prof. Lineu Prestes, 748 - bloco 3 - Superior, 05508-000 São Paulo SP - Brazil, C.P. 26.037 - 05599-970, Tel.: +55 11 3032.2299, Fax: +55 11 3814.3602 - São Paulo - SP - Brazil
E-mail: quimicanova@sbq.org.br