Next Article in Journal
Hybrid Chlorides with Methylhydrazinium Cation: [CH3NH2NH2]CdCl3 and Jahn-Teller Distorted [CH3NH2NH2]CuCl3
Next Article in Special Issue
Improving the Energy Storage of Supercapattery Devices through Electrolyte Optimization for Mg(NbAgS)x(SO4)y Electrode Materials
Previous Article in Journal
New Supramolecular Hypoxia-Sensitive Complexes Based on Azo-Thiacalixarene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanistic Aspects of the Electrochemical Oxidation of Aliphatic Amines and Aniline Derivatives

by
Ashwin K. V. Mruthunjaya
and
Angel A. J. Torriero
*
School of Life and Environmental Sciences, Deakin University, Burwood VIC 3125, Australia
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(2), 471; https://doi.org/10.3390/molecules28020471
Submission received: 22 December 2022 / Revised: 30 December 2022 / Accepted: 1 January 2023 / Published: 4 January 2023
(This article belongs to the Special Issue Exclusive Feature Papers in Electrochemistry)

Abstract

:
The electrochemical oxidation of amines is an essential alternative to the conventional chemical transformation that provides critical routes for synthesising and modifying a wide range of chemically useful molecules, including pharmaceuticals and agrochemicals. As a result, the anodic reactivity of these compounds has been extensively researched over the past seven decades. However, the different mechanistic aspects of the electrochemical oxidation of amines have never been discussed from a comprehensive and general point of view. This review examines the oxidation mechanism of aliphatic amines, amides, aniline and aniline derivatives, carbamates, and lactams, either directly oxidised at different electrode surfaces or indirectly oxidised by a reversible redox molecule, in which the reactive form was generated in situ. The mechanisms are compared and simplified to understand all possible pathways for the oxidation of amines using only a few general mechanisms. Examples of the application of these oxidation reactions are also provided.

Graphical Abstract

1. Introduction

A normal synthetic reaction implies the attack of a nucleophile on an electrophilic centre, with the reaction between molecules of similar polarity being considered inviable. In this scenario, the inversion of the polarity of one of those molecules is required, which is not an easy task in conventional organic synthesis. However, it is commonly realised in electrochemical organic synthesis, making possible a large variety of reactions [1,2].
Amines are a family of chemical compounds that share as a common feature the presence of at least one nitrogen atom whose hybridisation depends on the structure of the molecule. For example, aliphatic amines contain sp3-hybridised nitrogen atoms. This hybridisation lies between sp3 and sp2 when the amine is part of a resonance structure. Alternatively, it shows sp2-hybridisation when forming part of a heterocycle. In all cases, the amines contain a lone pair of electrons in the unbounded orbital, which is responsible for their relatively easy electrochemical oxidation.
Due to the large variety of amine-containing molecules available and extensive research on them over the past seven decades, it is impossible to discuss and cite all work performed in this area in a single paper. Nevertheless, several review articles and book chapters have been published summarising the impressive advances in this field over the years [3,4,5,6,7,8]. However, a detailed discussion of the different mechanistic aspects of the electrochemical oxidation of amines is still elusive. Therefore, this manuscript focuses on the comprehensive discussion of the oxidation mechanism of aliphatic amines, amides, aniline and its derivatives, and carbamates and lactams, the respective similarities and differences between their mechanisms, and catalysed electrochemical oxidations. Examples of the application of these oxidation reactions are also provided.

2. Aliphatic Amines

The potential at which aliphatic amines can be electrochemically oxidised depends on their structure (Table 1), with secondary or tertiary amines being easier to oxidise than primary amines. Nevertheless, the general mechanism for the electrochemical oxidation of simple aliphatic amines is the same and independent of the number of organic substituents attached to the nitrogen atom [9,10,11,12,13,14,15]. Upon the oxidation of a tertiary amine, the overall reaction provides a secondary amine, an aldehyde, and protons. These protons protonate the starting amine or the secondary one (product from this reaction) to give an electrochemically inactive ammonium ion, and the reaction consumes one electron per starting molecule [10,13,16,17,18]. Similarly, the oxidation of a secondary amine produces a primary ammonium ion as the product, while the oxidation of primary amines forms ammonia [19].
The one-electron oxidation reaction starts with the amine oxidation to the respective radical cation (Equation (1)), which deprotonates to give a radical at the α-carbon connected to the nitrogen atom (Equation (2)).
Molecules 28 00471 i001
Molecules 28 00471 i002
This radical can be involved in a disproportionation process to yield the starting amine and an enamine (1, Equation (3)) or be involved in a second oxidation step to produce an iminium cation (2, Equation (4)). Independently of this, it is expected that 1 and 2 be in equilibrium (Equation (5)), as the enamine would be a stronger base than the starting saturated amine [13]. Intermediate 2 is also formed during the oxidation of alicyclic amines, such as piperidine, piperazine, and their derivatives [29].
Molecules 28 00471 i003
Molecules 28 00471 i004
Molecules 28 00471 i005
It is important to highlight that although the mechanism postulates the consumption of two electrons, coulometry generally shows the consumption of one electron per molecule of starting amine because an extra molecule is inactivated by protonation.
Because of its positive charge, 2 is a better electron acceptor than a ketone carbonyl. Thus, any weak or strong nucleophile can react with 2 to form various products. For example, Equations (6) and (7) show the reaction of 2 in the presence of water. It begins with the nucleophilic addition of water to the iminium group (Equation (6)), followed by the transfer of a proton from oxygen to nitrogen to yield the protonated amino alcohol 3 (or carbinolamine), which converts the amine into a better leaving group. Next, the E1-like loss of amine produces a protonated aldehyde (Equation (7)). Finally, the loss of a proton from oxygen gives the final aldehyde and a quaternary ammonium product (Equation (7)).
Molecules 28 00471 i006
Molecules 28 00471 i007
Ross considered that the hydrolysis reaction follows a concerted or two-step base-catalysed mechanism (Equation (8)) [30], which produces the aldehyde in the enolic form and the amine. Then, keto-enol tautomerisation produces the final aldehyde product. However, this is not ubiquitous, and the following evidence confirmed that Equations (6) and (7) are more appropriate in specific situations [17]:
(1) Equation (8) cannot explain the demethylation of trimethylamine or other methylated amines.
(2) Experimental evidence confirmed that the reaction produces an ammonium ion and not an amine.
Molecules 28 00471 i008
Alternatively, if the anodic oxidation of amines is performed in the presence of methanol, the methoxylation of the amine-containing molecule is observed. For example, the methoxylation reaction of N,N-dimethylbenzylamine was reported [14,31,32,33]. Based on the previous mechanism, the equivalent molecule of intermediate 2 (Equation (4)) reacts with methanol. However, the molecule offers two possible intermediate 2 (2a and 2b):
  • Molecules 28 00471 i009
It was observed that 2a is of greatest preference (methyl attack/benzyl attack = 10) as the methyl hydrogens are more reactive than the methylene hydrogens [31]. Equation (9) shows this reaction and the final product obtained in a larger yield (4). The demethylation product N-methylbenzylamine (5) and 1-methoxy-N,N-dimethyl-1-phenylmethanamine (6) were also obtained as secondary products in the same reaction [31].
Molecules 28 00471 i010
The electrochemical oxidation of benzylamine in acetonitrile at a stainless-steel mesh anode in the presence of 0.1 M tetrabutylammonium perchlorate as the supporting electrolyte was studied [34]. Due to the absence of methyl groups, the formation of an intermediate equivalent to 2b occurs, which reacts with starting material (Equation (10)), resulting in the formation of N-benzylidenebenzylamine (7) and a small percentage of benzonitrile [34].
Molecules 28 00471 i011
A similar situation to that described above for N,N-dimethylbenzylamine was observed in our group during the electrochemical oxidation of N,N-dicyclohexylmethylamine, N,N-dimethylcyclohexylamine, and N,N-dicyclohexylamine [17]. Using N,N-dicyclohexylmethylamine as an example, the initial oxidation affords a radical cation (Equation (1)), which can deprotonate following two possible paths to give a radical. One pathway is the formation of cyclohexyl radical 8 (Equation (11)), and the second option is the formation of methylene radical 9 (Equation (12)). From the electrolysis results of N,N-dicyclohexylmethylamine and N,N-dimethylcyclohexylamine, it was possible to observe that the formation of radical 9 is preferential over that of radical 8.
Molecules 28 00471 i012
Molecules 28 00471 i013
However, in the absence of methyl groups, the formation of radical 8 occurs, as was observed with N,N-dicyclohexylamine. It seems worthwhile to point out that in the case of radical 9, an enamine intermediate cannot be formed. Nevertheless, the demethylation process still takes place. Water molecules in the organic solvent reacted with the N,N-dicyclohexylmethylamine and N,N-dimethylcyclohexylamine iminium products to yield formaldehyde and protonated N,N-dicyclohexylamine and N-methylcyclohexylamine, respectively. N-cyclohexylamine was obtained as the oxidation product of N,N-dicyclohexylamine [17].
The experimental results were rationalised by considering that planarity at the iminium intermediate (sp2-hybridised carbon) is required (Equation (5)), which will be favoured in the methyl group rather than in the cyclohexyl functional group. Moreover, steric effects contribute to the difficulty in accessing the α-carbon hydrogens in highly substituted tertiary amines (e.g., tri-isopropylamine and 9-t-butylazabicyclo[3.3.1]nonane) showing electrochemical reversible oxidation processes [19,35,36]. Similar reversibility is also observed with amines containing no hydrogens on the α-carbons [35,36]. Interestingly, the reversibility of 9-t-butylazabicyclo[3.3.1]nonane is partially lost when the t-butyl group is replaced by i-propyl [36].
Despite the well-known mechanism, research on adding nucleophiles to 2 is limited. Table 2 summarises the anodic oxidation of aliphatic and alicyclic amines in the presence of different nucleophiles.
Table 2. Oxidation of tertiary amines in the presence of different nucleophiles.
Table 2. Oxidation of tertiary amines in the presence of different nucleophiles.
Entry #Starting AmineNucleophileProduct% YieldRef.
1Molecules 28 00471 i014CH3OHMolecules 28 00471 i01548[37]
Molecules 28 00471 i01612
2Molecules 28 00471 i017Intramolecular OHMolecules 28 00471 i01825[37]
Molecules 28 00471 i01945
3Molecules 28 00471 i020Intramolecular OHMolecules 28 00471 i02125
Molecules 28 00471 i02245
4Molecules 28 00471 i023–CNMolecules 28 00471 i02453[38]
DEM–Molecules 28 00471 i02580[39]
DEP–Molecules 28 00471 i02670
5Molecules 28 00471 i027–CNMolecules 28 00471 i02836[38]
6Molecules 28 00471 i029–CNMolecules 28 00471 i03031[38]
Molecules 28 00471 i031
7Molecules 28 00471 i032–CNMolecules 28 00471 i03332[38]
Molecules 28 00471 i034
8Molecules 28 00471 i035–CNMolecules 28 00471 i03640[38]
Molecules 28 00471 i037
9Molecules 28 00471 i038–CNMolecules 28 00471 i03940[38]
10Molecules 28 00471 i040–CNMolecules 28 00471 i04143[38]
11Molecules 28 00471 i042–CNMolecules 28 00471 i04346[38]
Molecules 28 00471 i044
12Molecules 28 00471 i045–CNMolecules 28 00471 i04657[38]
Molecules 28 00471 i047
13Molecules 28 00471 i048–CNMolecules 28 00471 i04957[38]
Molecules 28 00471 i050
14Molecules 28 00471 i051–CNMolecules 28 00471 i05261[38]
Molecules 28 00471 i053
15Molecules 28 00471 i054–CNMolecules 28 00471 i05559[38]
16Molecules 28 00471 i056–CNMolecules 28 00471 i05757[38]
17Molecules 28 00471 i058–CNMolecules 28 00471 i05962[38]
18Molecules 28 00471 i060–CNMolecules 28 00471 i06163[38]
19Molecules 28 00471 i062–CNMolecules 28 00471 i063[40]
20Molecules 28 00471 i064Intramolecular OHMolecules 28 00471 i06572[41,42]
–CH2CNMolecules 28 00471 i06622[42]
–CNMolecules 28 00471 i06723
21Molecules 28 00471 i068DEM–Molecules 28 00471 i06976[39]
DEP–Molecules 28 00471 i07060
22Molecules 28 00471 i071Molecules 28 00471 i072Molecules 28 00471 i07374[43]
23Molecules 28 00471 i074HCOOMolecules 28 00471 i07579[44]
24Molecules 28 00471 i076Molecules 28 00471 i077Molecules 28 00471 i07890[44]
Molecules 28 00471 i079Molecules 28 00471 i08082[44]
25Molecules 28 00471 i081H2OMolecules 28 00471 i08255[45]
26Molecules 28 00471 i083H2OMolecules 28 00471 i08464[45]
27Molecules 28 00471 i085–CNMolecules 28 00471 i08670[46]
28Molecules 28 00471 i087–CNMolecules 28 00471 i08870[46]
29Molecules 28 00471 i089Intramolecular OHMolecules 28 00471 i09048[47]
30Molecules 28 00471 i091CH3OHMolecules 28 00471 i09276[48]
31Molecules 28 00471 i093Intramolecular C=CMolecules 28 00471 i09489[49,50]
32Molecules 28 00471 i095Intramolecular NMolecules 28 00471 i09679[51]
33Molecules 28 00471 i097CH3OHMolecules 28 00471 i09888[52]
Molecules 28 00471 i099Molecules 28 00471 i10087
CH3COOHMolecules 28 00471 i10154
34Molecules 28 00471 i102TsN3Molecules 28 00471 i10368[53]
35Molecules 28 00471 i104TsN3Molecules 28 00471 i10576[53]
36Molecules 28 00471 i106TsN3Molecules 28 00471 i10752[53]
Abbreviations: –CN = cyanide ion; DEP– = diethyl phosphonate ion, DEM– = diethyl malonate ion, TsN3 = tosyl azide.
Aziridines are extremely reactive cyclic secondary amines that undergo ring-opening reactions in the presence of nucleophiles. They are used in textile chemicals, adhesives, binders, petroleum refining chemicals, fuels, lubricants, hardeners, etc. [54]. Aziridines undergo ring cleavage by anodic oxidation. For example, the electrochemical oxidation of 2-phenyl-2-ethylaziridine was studied in anhydrous methanol at 0 °C under a nitrogen atmosphere using a platinum working electrode [55]. The electrolysis showed a four-electron process with the production of (1,1-dimethoxypropyl)benzene in 50% yield. The reaction was postulated to proceed via initial two-electron oxidation to form an azaallyl cation intermediate (10, Equation (13)), which reacts with methanol to produce the imine 11. This imine can suffer further two-electron oxidation to make 12 (12% yield, Equation (14)) and react with water to form propiophenone (observed in 6% yield) or react again with methanol to produce (1,1-dimethoxypropyl)benzene [55].
Molecules 28 00471 i108
Molecules 28 00471 i109

2.1. Catalysed Oxidation of Amines

To decrease the overpotential needed for the direct oxidation of amines, as well as minimise the electrode surface fouling effect, increase the sensitivity, and enhance the reliability and reproducibility of the data, the catalytic oxidation of aliphatic amines with reversible redox couples acting as a mediator has been introduced. As discussed by Torriero et al., an ideal redox catalyst (or mediator) needs to have a standard reversible potential less positive than the oxidation potential of the substrate, exhibit fast electron-transfer kinetics, and be stable in both the oxidised and reduced form toward the species present in the reaction media [17,18,56,57]. Effective mediators that meet these requirements are based on ruthenium complexes, quinone, ferrocene (Fc), and their derivatives, either homogeneously dispersed in the solution or immobilised in a monolayer or multilayer configuration onto the electrode surface [17,18,58,59,60,61]. However, other options were equally reported. For example, the anodic oxidation of primary amines at nickel hydroxide electrodes in alkaline solutions forms nickel oxide hydroxide at a potential of 0.39 V vs. SCE, which reacts with propylamine and butylamine, forming propionitrile (84% yield) and butyronitrile (85% yield), respectively [62]. Nevertheless, when i-propylamine was used, acetone was formed in an 80% yield (Equation (15)).
Molecules 28 00471 i110
Chloride, bromide, and iodide have been used as redox catalysts (Table 3). For example, sodium chloride and sodium bromide were used as redox catalysts in a CH3CN-saturated NaCl aqueous (pH 4) solvent mixture using platinum as the working electrode for the electrochemical oxidation of cyclic aziridines to form keto nitriles in an 80% yield [63]. Keto nitriles are valuable intermediates for various synthetic transformations [64]. This reaction was explained by the generation of an azaallenyl intermediate cation, which is hydrated to form an α-hydroxyimine. Second oxidation by Cl+ followed by a rearrangement leads to the final keto nitrile (13, Equation (16)) [63].
Molecules 28 00471 i111
Table 3. Oxidation of amines using halides as redox catalysts.
Table 3. Oxidation of amines using halides as redox catalysts.
Entry #Starting AmineSecond ReactantRedox CatalystProduct% YieldRef.
1Molecules 28 00471 i112H2OIMolecules 28 00471 i11375[65]
2Molecules 28 00471 i114Molecules 28 00471 i115IMolecules 28 00471 i11685[65]
3Molecules 28 00471 i117Molecules 28 00471 i118IMolecules 28 00471 i11983[65]
4Molecules 28 00471 i120Molecules 28 00471 i121IMolecules 28 00471 i12275[65]
Molecules 28 00471 i123IMolecules 28 00471 i12417[65]
Molecules 28 00471 i125IMolecules 28 00471 i12679[66]
5Molecules 28 00471 i127H2OIMolecules 28 00471 i12853[65]
6Molecules 28 00471 i129Molecules 28 00471 i130IMolecules 28 00471 i13181[66]
7Molecules 28 00471 i132IMolecules 28 00471 i13374[67]
8Molecules 28 00471 i134H2OBrMolecules 28 00471 i13550[68]
9Molecules 28 00471 i136H2OBrMolecules 28 00471 i13785[68]
10Molecules 28 00471 i138PhS-SPhBrMolecules 28 00471 i13972[69]
11Molecules 28 00471 i140CH3OHClMolecules 28 00471 i14183[70]
12Molecules 28 00471 i142ClMolecules 28 00471 i14366[67]
13Molecules 28 00471 i144CH3OHClMolecules 28 00471 i14538[51]
14Molecules 28 00471 i146CH3OHClMolecules 28 00471 i14765[71]
The previously postulated aziridine oxidation mechanism (Equations (13) and (14)) was also reported using different catalysts. For example, the acid-catalysed oxidation of N-benzylaziridine in methanol was reported to form 1,4,7,10-tetraphenyl-1,4,7,10-tetraazacyclododecane in 60% yield [72,73]. The same reaction catalysed by tris(4-bromophenyl)amine was also reported [4]. However, 1,4-dibenzylpiperazine was described as the product when the reaction was catalysed by iron porphyrin in a 5:1 CH3CN/H2O deaerated solvent system [74].
The two-electron oxidation of (3-amino-2,4-dihydroxyphenyl)(phenyl)methanone (14) to the corresponding benzoquinoneimine in the presence of different amines was studied in methanol containing either LiClO4, [Et4N][PF6], or [Et4N]ClO4 as the supporting electrolyte and using platinum as the working electrode (Equation (17)) [59,75]. The oxidation peak of the catalyst was observed at a potential of 0 V vs. SCE under these conditions.
Molecules 28 00471 i148
The amine transfers one electron per molecule to the benzoquinoneimine to regenerate the catalyst and afford the imine dimeric product. It was reported that only primary alkyl amines react with this mediator, with linear alkyl groups more reactive than branched ones, such as cyclohexyl or t-butyl groups. Meanwhile, phenylethylamine deactivates the catalyst following the reaction shown in Equation (18). It involves the reaction of the enamine form of the product (see Equation (5)) with the benzoquinoneimine form of the mediator following a Diels–Alder-type mechanism. The product 15 is unstable. Nevertheless, its two-electron oxidation product was isolated in a 65% yield [59].
Molecules 28 00471 i149
The reversible one-electron oxidation of para-substituted triphenylamines was also used as redox catalysts in acetonitrile and sometimes even in nucleophilic media for several amine oxidation reactions. The formal potential for the oxidation of the p-substituted triphenylamines depends on the electron-donative or electron-withdrawing nature of the p-substituent, covering the potential range of 0.6 to 1.8 V vs. SCE [25,76,77,78]. For example, the electrochemical oxidation of N-ethyl-S-(2-nitrophenyl)thiohydroxylamine was performed at 0.69 V vs. SCE in dichloromethane containing 0.1 M [Bu4N][ClO4] as the supporting electrolyte in the presence of N1,N1,N4,N4-tetrakis(4-bromophenyl)benzene-1,4-diamine as the catalyst (Equation (19)) [79]. The respective sulfenimine product was isolated in a 72% yield. Similarly, the electrochemical oxidation of benzylamine was performed at 0.99 V vs. SCE in acetonitrile containing 0.1 M [Bu4N][BF4] as the supporting electrolyte in the presence of tris(4-bromophenyl)amine as the catalyst (Equation (20)). Under these conditions, the respective iminium product was identified. Nevertheless, when the reaction was performed in a 50:50 dichloromethane:methanol solvent mixture, N-benzyl-1-phenylmethanimine was isolated in a 78% yield [80].
Molecules 28 00471 i150
Molecules 28 00471 i151
Pentaammineaquaruthenium(II) trifluoroacetate was used to catalyse the formation of amide functional groups from nitrile [81,82]. The reaction starts with the nitrile substituting the aqua ligand in the ruthenium complex (Equation (21)). Oxidation of this ruthenium(II) complex to ruthenium(III) permits the hydrolysis of nitrile to the amide. The reduction of this complex ensures the formation of the amide in a 70% yield and the recovery of the catalyst (Equation (22)) [82].
Molecules 28 00471 i152
Molecules 28 00471 i153
When a primary amine is used instead of a nitrile, the oxidation reaction forms the imine product or even the nitrile. For example, in the case of benzylamine, the final product was benzonitrile, with an 85% yield [82].
The oxidation of Fc to the corresponding ferrocenium cation, in the presence of different cyclohexylamines and n-alkylamines, was also studied in acetonitrile and dichloromethane containing 0.1 M [Bu4N][PF6] as the supporting electrolyte [17,18]. The reaction follows the mechanism described in Equations (1)–(7), obtaining the dealkylated amine as the main product (see above).

2.2. Examples of Applications

The electrochemical oxidation of cardiovascular-active kopsingine alkaloid in a CH2Cl2/CH3CN solvent system and in the presence of the non-nucleophilic base 2,6-lutidine results in an intramolecular cyclisation, generated by the attack of a hydroxy group to the electrochemically generated iminium ion, forming kopsidine A in 72% yield (Table 2, entry 20) [41,42].
Catharanthus alkaloids are valuable reagents for treating several cancers [83,84]. The electrochemical oxidation of the catharanthus roseus alkaloids catharanthine produces the iminium ion, which fragmented to give the highly cytostatic anhydrovinblastine. In the presence of methanol, the methoxy group is introduced in C16, making 16-methoxycleavamine in a 95% yield [85].
The 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) catalyst was used in the electrochemical N-demethylation of opiates resulting in noropiates, a critical intermediate in the opiate medicine partial chemical synthesis with a good yield of up to 83% [86]. This N-demethylation process follows through the anodic oxidation of TEMPO, resulting in an oxoammonium species, which then oxidises opiate to an iminium cation. Finally, this intermediate hydrolyses to yield noropiate.
The microsomal cytochrome P-450 monooxygenase system catalyses the dealkylation of secondary and tertiary amines and amides via an oxidation pathway. Therefore, the product of the enzymatic processes was compared with those discussed above to gain insight into the cytochrome P-450 catalysed N-dealkylation [87]. An agreement on the selectivity of dealkylation between microsomal and anodic dealkylations was observed [87].
Our group used the ferrocene-mediated oxidation of alkylamines to construct a selective and sensitive electrochemical biosensor to detect DNA hybridisation by employing the electrocatalytic activity of Fc-bearing Zn-cyclen complexes [60]. A sandwich-type approach was created, which involves hybridising a target probe with the immobilised thiolated capture probe attached to a gold electrode. Electrochemical signals are generated by voltammetric interrogation of Fc complexes that selectively and quantitatively bind to the duplex layers through strong chelation between the Zn-cyclen complexes and thymine bases within the DNA sequence. Coupling the redox chemistry of the surface-bound Fc-bearing Zn-cyclen complex and dimethylamine provides an electrocatalytic pathway that increases the sensitivity of the assay and allows the target DNA sequence to be detected at a 100 fM concentration level [60].
The selective electrochemical oxidation of tropane alkaloids to their nortropane derivatives using GC as the working electrode in a 2:1 ethanol:water or 2:1 methanol:water solvent system (0.1 M NaClO4) was described [40]. These oxidation products are important intermediates in the production of anticholinergics ipratropium and oxitropium bromide drugs. The reaction proceeds at room temperature following a mechanism similar to that described above to form the iminium intermediate (see Equations (1)–(5)), which reacts in the presence of water to form the respective nortropane (16) and formaldehyde (Equation (23)). When the oxidation reaction was performed in the presence of cyanide or 16 as nucleophiles, the respective addition reactions were observed (Table 2, entry 19 and Equation (24)) [7,40,88].
Molecules 28 00471 i154
Molecules 28 00471 i155

2.3. Alkanolamines, Amides, Carbamates and Lactams

The term alkanolamines describes molecules that simultaneously contain amino and hydroxyl functional groups. The oxidation of these molecules in alkaline or neutral aqueous systems undergoes the previously mentioned dealkylation mechanism. For example, the oxidation of 2-(dimethylamino)-1-phenylethan-1-ol (17) yields benzaldehyde and the respective radical (Equation (25)), which disproportionate or loses a second electron to form an iminium intermediate (Equation (26)). The formation of formaldehyde and dimethylamine is observed after the interaction of this reactive intermediate with water molecules [89]. The electrochemical oxidation of ephedrine (2-(methylamino)-1-phenylpropan-1-ol) was evaluated using GCE as the working electrode in a pH 10 aqueous solution. The oxidation products benzaldehyde, acetaldehyde, and N-methylamine were obtained in yields of 65%, 68%, and 87%, respectively [90]. A similar mechanism and product distribution were observed for related alkanolamines, confirming the proposed mechanism [79,91,92].
Molecules 28 00471 i156
Molecules 28 00471 i157
The electrochemical oxidation of 2,2′-(benzylazanediyl)bis(ethan-1-ol) was studied in alkaline methanol using a platinum working electrode. The major electrolysis products were 2-(2-phenyloxazolidin-3-yl)ethan-1-ol and 3-benzyloxazolidine in a 25% and 45% yield, respectively (Table 2, entry 2) [37]. Similarly, the anodic oxidation of 2-(benzyl(methyl)amino)ethan-1-ol under the same conditions leads to the formation of a mixture of 3-methyl-2-phenyl-oxazolidine and 3-benzyloxazolidine (Table 2, entry 3) [37].
The anodic oxidation of amides and carbamates is of considerable value from a synthetic point of view due to the stability of the intermediates [7]. Examples of synthetic applications can be observed in the alkaloids, enantiopure amino acids, chiral α-hydroxyl amide metabolites, and peptidomimetics areas [7,43,44,46,47,48,49,50,51,52,88,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109]. The reaction proceeds via the initial one-electron oxidation of the nitrogen atom to the respective radical cation, which follows a similar pathway to that described before (Equations (1)–(5)) to produce the respective iminium intermediate (18, Equation (27)) [11,12]. The subsequent reactions of the iminium cations with nucleophiles (Equation (28)) have been extensively reported as amidoalkylation (examples in Table 2, entries 22–33). Typical nucleophiles include hydroxyl, enamines, isocyanides, enol esters, electron-rich olefins and aromatics, enol ethers, trimethylsilyl cyanide, vinyl and allyl silanes, and trialkylphosphites [44,46,47,48,49,50,51,52,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,110,111]. Meanwhile, as the iminium is in equilibrium with its enamine form (Equation (5)), this last intermediate can react with electrophilic groups (Equation (29)). Typical electrophiles include acyl chlorides and alkyl halides [112,113,114,115,116,117].
Molecules 28 00471 i158
Molecules 28 00471 i159
Molecules 28 00471 i160
The conversion of an α-aminomalonic half-ester or an N-acylated amino acid via the Hofer–Moest reaction also allows the formation of 18 (Equation (30)). The reaction mechanism is similar to that shown above for alkanolamines (Equations (25) and (26)) and comprises an electrolytic decarboxylation reaction in neutral or alkaline solutions [118]. Because of the similarity in the mechanism with alkanolamines, the electrochemical oxidation of N-acylated β-amino alcohols will also generate the same N-acyliminium ion 18 (Equation (31)).
Molecules 28 00471 i161
Molecules 28 00471 i162
Secondary amide anions can be oxidised at potentials about 1 V more cathodic to the related amine (Table 1). The electrochemical irreversible one-electron oxidation produces the aminyl radical, which dimerises to make the respective hydrazine (Equation (32)). A similar result was obtained when secondary amines were electrochemically oxidised in an alkaline media [119,120,121]. Furthermore, it was observed that the reaction is effective if the intramolecular coupling occurs, producing cyclic hydrazines [119,120]. When di-n-butylamide is electrochemically oxidised in THF in the presence of di-n-butylamine, N,N-dibutyltetrahydrofuran-2-amine was formed in 30% yields [122].
Molecules 28 00471 i163
The electrochemical oxidation of lactams follows the same mechanism postulated for amides. Nevertheless, the oxidation position depends on the lactam ring size, structure of the N-alkyl substituent and electrolysis conditions. In the case of five- and six-member N-alkyl lactam rings, the oxidation and nucleophilic addition occur selectively at the lactam α-carbon to nitrogen (Table 2, entries 25, 26). Meanwhile, seven-member lactam rings show the reaction at the N-alkyl α-carbon [123,124,125,126]. Methoxylation at both positions was observed after the anodic oxidation of N-alkyl-β-lactam and N-benzyl-β-lactams in methanol using a platinum working electrode [45,127].
The anodic oxidation of 4-carboxy-2-azetidinone (19) in acetonitrile follows the decarboxylation mechanism described in Equation (30), which in the presence of sodium acetate produces 4-acetoxy-2-azetidinones (20) in 76% yield (Equation (33)) [128].
Molecules 28 00471 i164

3. Aniline and Derivatives

The electrochemical anodic oxidation of aromatic amines has been studied extensively. The prime focus of all those investigations was to establish the oxidation mechanism under multiple electrolysis setups due to their wide variety of applications in synthetic and polymer chemistry and pharmaceutical and dye industries. In the literature, several reviews of the electrochemical oxidation of aromatic amines are available [4,7,129]. Hence, the anodic oxidation mechanism of aniline and its derivatives, which are not comprehensively discussed in the available literature, is addressed in this work.
Molecules 28 00471 i165
The electrochemical potential at which aniline derivatives oxidise depends on the nature of the substituents present both at the nitrogen and the aromatic ring. Electron-withdrawing substituents shift the oxidation redox potential towards more positive values, while the electron-donating substituents change the oxidation potential in the opposite direction (Table 1). Nevertheless, a common feature in all cases is that the anodic oxidation of aniline and its derivatives starts with the nitrogen loss of one electron to yield the radical cation 21 (Equation (34)). After this point, the oxidation mechanism depends on the substitution level at the nitrogen atom and the basicity of the reaction media. For N,N-disubstituted anilines under basic conditions, the oxidation mechanism is similar to that previously described for aliphatic amines with the formation of the respective iminium product 23 (Equation (34)). However, nitrogen deprotonation reaction is predominantly observed in mono- or no N-substituted anilines, producing the respective radical intermediate 24. Alternatively, the subsequent oxidation produces 22, which may deprotonate to generate the respective nitrenium cation (Equation (34)) [130].
Writing the resonance structures for 24 may help to understand all the reaction pathways discussed below. Based on this resonance, it is possible to see that the radical may reside on the nitrogen atom or delocalised at the ortho (25) and para (26) positions in the aromatic ring (Equation (35)).
Molecules 28 00471 i166
Therefore, it is evident that a diverse range of products can be formed, where two radical cations, two radicals, a radical cation and a radical, or a cation and a starting molecule can couple together via the formation of a new C-C, C-N, or N-N bond [131]. For example, the following general reactions can be postulated [132]:
(i).
Two resonance structures 26 can react through a tail-to-tail coupling to form dimer 27 (Equation (36)).
(ii).
The nitrenium cation can react with starting molecule through a head-to-tail coupling to form dimer 28. The same outcome is obtained by reacting 24 with 26 (Equation (37)).
(iii).
Two resonance structures 24 can react via a head-to-head coupling to form dimer 30 (Equation (38)).
In the case of N,N-disubstituted anilines under basic conditions, iminium product 23 can react with nucleophiles present in the solution to produce the respective addition (or an α-substitution if we consider the starting molecule) products (Equation (39)).
Molecules 28 00471 i167
Molecules 28 00471 i168
Molecules 28 00471 i169
Molecules 28 00471 i170

3.1. Aniline and Para-Substituted Anilines

The electrochemical anodic oxidation of aniline may follow the different mechanisms outlined in Equations (36)–(38). The major product obtained depends on the aqueous or nonaqueous, acidic, basic, or neutral pH media conditions. Acidic conditions, which are usually used to facilitate the dissolution in aqueous media of otherwise insoluble aniline derivatives, may require the addition to the reaction mechanism of an initial deprotonation homogeneous chemical reaction before the first electron transfer, which most of the time is not taken into consideration when reporting the oxidation mechanisms under these conditions.
The three different anodic oxidation products of aniline are as follows. When there is no para substituent (or X=H), the intermediate 26 can undergo tail-to-tail coupling with itself or head-to-tail coupling with 24 in an acidic aqueous medium, producing benzidine (27) in a small yield and p-aminodiphenylamine (28), respectively (Equations (36) and (37)), both of which can be reversibly oxidised via a two-electron and two-proton process [133]. The head-to-head dimerisation reaction to form hydrazobenzene 30 was reported both in acidic and basic reaction conditions [23,134]. In the case of 28, the oxidation product may undergo slow acid-catalysed hydrolysis to produce p-benzoquinone and the parent amine, in this case, aniline (Equation (40)). The anodic oxidation of aniline in a nonaqueous medium (e.g., acetonitrile) generates 28 as the main product, which is stable under these conditions due to the non-availability of protons to catalyse the previously mentioned hydrolysis.
Molecules 28 00471 i171
When there is a substituent at para position (X = OCH3, OC2H5, Cl, CH3, COOH, NO2), intermediate 24 undergoes a head-to-head coupling predominantly to produce hydrazobenzene 30 derivatives [133].
When aniline and its derivatives are oxidised using an acetonitrile/pyridine solvent system, azobenzene 31 is produced in 39% yield, with 30 appearing as a by-product [23]. The head-to-head coupling of radical cations in the presence of pyridine can be explained using Equations (34) and (38), where the radical cation 21 reacts with pyridine in a Bronsted–Lowry acid–base reaction producing 24 (in this case, R1 = R2 = H), which undergoes head-to-head coupling to yield 30 (Equation (38)). A simple explanation for the generation of 29 could be the head-to-tail coupling of the neutral radical and the subsequent two-electron and two-proton oxidation process. Pyridine plays the crucial role of proton acceptor in this reaction. For example, electrochemical oxidation of p-nitroaniline and p-chloroaniline in acetonitrile/pyridine produced 4,4′-dinitroazobenzene (39.2% yield) and 4,4′-dichloroazobenzene (24.3% yield), respectively [23]. The anodic oxidation of 2,4-dinitroaniline produced 2,2′,4,4′-tetranitroazobenzene in a 38% and 31% yield when a 3:5 water:acetonitrile and 1:3 water:DMF solvent mixtures were used, respectively [135]. Aniline was oxidised in 1M KOH, forming hydrazobenzene, which produced 31 (Equation (38)) in a 30% yield after further oxidation [136].

3.2. N-Substituted Anilines

After the first oxidative electron transfer step, the mono N-alkyl or N-aryl substituted anilines show a relatively more stable radial cation due to the stabilisation effect of its electron-donating substituents. There are four possible ways this radical cation undergoes subsequent reactions depending on the reaction conditions. The first three pathways are identical to those explained for aniline (Equations (36)–(38)). The fourth one is in the presence of nucleophiles following the nucleophilic substitution at the α-carbon to the nitrogen through an iminium ion intermediate (Equation (39)). For example, if the electrolysis reaction is performed under high current density conditions (e.g., 4 mA cm2), the high concentration of radical cations produced and the lower concentration of parent molecules remaining at the electrode surface leads to a tail-to-tail coupling product. However, when the reactions are performed under strongly basic conditions and at lower current densities (e.g., ≤0.8 mA cm2), the higher concentration of parent molecules at the electrode surface favours the head-to-tail coupling product formation, which in this case may undergo oxidation and hydrolysis to yield p-benzoquinone and starting amine [7]. Moreover, the bulkiness of the N-alkyl group plays a crucial role in product formation. For example, in acidic conditions, electrolysis of N-methylaniline at a high current density produces p-benzoquinone (50%) and N,N′-dimethylbenzidine (50%); N-ethylaniline produces p-benzoquinone (40%) and N,N′-diethylbenzidine (60%); N-t-butylaniline in acetonitrile at a high current density yields N,N′-di-t-butylbenzidine (100%) [133].
The oxidation of diphenylamine in acetonitrile using platinum electrodes generates the electroactive N,N′-diphenylbenzidine product [137]. The same type of coupling can be observed even when the oxidation of diphenylamine is carried out under weakly basic conditions [138]. Nevertheless, under weakly basic conditions, if a methoxy group is present in the para position, as in the case of dianisylamine, 2,7-dimethoxy-9,10-dianisyl-9,10-dihydrophenazine is formed as the main product [138].
N-alkyl substituted anilines also undergo nucleophilic substitution at the α-carbon to the nitrogen (Equation (39)) in the presence of nucleophiles such as enol ethers [139]. For example, when N-methylaniline is oxidised under constant current in methanol containing LiClO4 as the supporting electrolyte and in the presence of 2,3-dihydrofuran, the tetrahydroquinoline 32 (12%) and the acetal 33 (6.3% trans-isomer and 5% cis-isomer) derivatives were obtained (Equation (41)) [139].
Molecules 28 00471 i172
The anodic oxidation of diphenylamine and its derivatives, o-methoxydiphenylamine, o-methyldiphenylamine, and m-methyldiphenylamine in methanol containing sodium cyanide as the supporting electrolyte and the source of nucleophile was performed, resulting in the formation of the respective cyanodiphenylamines in relatively good yield [140]. For example, the cyanation of diphenylamine happened at the para position of both phenyl rings to produce bis(4-cyanophenyl)amine in a 61% yield (Equation (42)). However, in the case of o-methoxydiphenylamine, o-methyldiphenylamine, and m-methyldiphenylamine, the nucleophilic aromatic substitution happened only at the para position of the substituted phenyl ring in a 40–50% yield.
Molecules 28 00471 i173

3.3. N,N-Disubstituted Anilines

The N,N-dialkylanilines and N,N-diarylanilines undergo a similar oxidation mechanism as discussed earlier for N-substituted anilines. For example, in the absence of para substituent, the tail-to-tail coupling producing derivatives of 27 was observed. The resulting dimer is more easily oxidised than the starting material, generating the respective quinoidal diimino cation [131].
The anodic oxidation of N,N-dimethylaniline derivatives has been extensively studied in different reaction media and it was found that the primary product formed through tail-to-tail coupling is N,N,N′,N′-tetramethylbenzidiene [24,141,142,143,144]. However, if the electrochemical oxidation of N,N-dimethylaniline is performed in the presence of phenothiazine, a para-selective head-to-tail coupling happens with the formation of N,N,N′,N′-tetramethylbenzidiene as a secondary product (Equation (43)) [145]. Similarly, anodic oxidation of 8-aminoquinolines and para-substituted N,N-dimethylaniline in the presence of sodium sulfinates generates radical–radical cross-coupling, resulting in the formation of sulfones with a new C-S bond [146,147].
Molecules 28 00471 i174
The oxidation of N,N-diphenylaniline in acetonitrile undertakes tail-to-tail coupling to yield tetraphenyl benzidine, and this dimer undergoes further oxidation, resulting in a quinoidal dication [25]. On the other hand, the oxidation of N,N-dimethyl-p-anisidine in acetonitrile containing traces of water results in the formation of 4-dimethylaminophenol [148]. The anodic oxidation of 4-dimethylaminophenol in aqueous media produces N,N-dimethylbenzoquinoneimine, which undergoes further hydrolysis to yield benzoquinone and dimethylamine in a reaction similar to that described in Equation (40) [149]. The oxidation of N,N-dimethyl-p-toluidine in acetonitrile leads to the tail-to-tail dimerisation product 4,4′-(ethane-1,2-diyl)bis(N,N-dimethylaniline) [148,150,151].
N,N-dialkylanilines can undergo nucleophilic substitution in α-position to nitrogen. For example, the anodic oxidation of N,N-dimethylaniline in methanol under basic conditions yields two products, α-methoxy-N-N-dimethylaniline and α-α′-dimethoxy-N-N-dimethylaniline in a ratio of 6:1 [37,152]. However, the methoxylation of N-ethyl-N-methylaniline predominately occurs at the methyl group, resulting in a highly regioselective reaction (Equation (44)) [153]. Similarly, the oxidation of N-ethyl-N-methylaniline in acetonitrile containing tetraethylammonium cyanide as the supporting electrolyte and the nucleophile source produced cyanation preferentially at the methyl position (64% yield) with the formation of 2-(methyl(phenyl)amino)propanenitrile by-product in a 34% yield [154].
Molecules 28 00471 i175
The anodic cyanation of 3,5-dimethyl-2,3,4,5-tetrahydro-1-benzazepine was performed at a carbon electrode in methanol containing sodium cyanide and lithium acetate as the supporting electrolyte. This reaction produced the trans-α-cyano derivative 34 and side product 35, where the cyanation occurs at the side chain (Equation (45)) [155].
Molecules 28 00471 i176

3.4. Aminophenols and N-Acylated Anilines

Like other anilines, the aminophenol radical cation formed after the initial electron transfer may undergo dimerisation reactions, resulting in a new N-N, C-C, or C-N bond depending on the reaction condition. These dimerisation pathways can be explained using Equations (36)–(38). For example, the anodic two-electron oxidation of p-aminophenol to quinoneimine is a well-documented reaction, which may undergo hydrolysis in acidic media, resulting in the formation of the respective p-benzoquinone (Equation (46)) [156,157]. The oxidation of o-aminophenol in basic or neutral media produces a dimer formed through N-N coupling of o-aminophenol cation radicals similar to that reported in Equation (38) [158]. However, if the same reaction was performed in acidic solutions, the C-N coupling of o-aminophenol results in the formation of 2-aminophenoxazin-3-one [158].
Molecules 28 00471 i177
When the anodic oxidation of o- or p-aminophenol derivatives results in a stable quinoneimine intermediate, they can then react with nucleophiles present in the solution. For example, the anodic oxidation of 1-(4-(4-hydroxyphenyl)piperazin-1-yl)ethan-1-one, 36, in a phosphate buffer/acetonitrile solvent mixture undergoes a two-electron and one proton transfer process to yield the respective quinoneimine, 37, which, in the presence of 2-mercaptobenzothiazole, produces the mono-thiolated product 38 (Equation (47)) [159]. This product can undergo a second two-electron oxidation process and add a second 2-mercaptobenzothiazole molecule to the remaining ortho position to the OH-group, generating the respective di-thiolated product in a 93% yield.
Molecules 28 00471 i178
The anodic oxidation of N-acylated aniline derivatives can yield various products depending on the nature of the reactants and reaction conditions. The reaction proceeds, as discussed in Equation (34), with the formation of the radical 24. Then, the radicals dimerise to create a new N-N or C-C bond. For example, the anodic oxidation of substituted anilides at graphite electrodes in acetonitrile containing potassium acetate and acetic acid produced N,N-diarylhydrazine derivatives, 39, in a ca. 63% yield (Equation (48)) [160]. Meanwhile, when the reaction is performed at a glassy carbon electrode in methanol containing tributylmethylammonium methylsulfate as the supporting electrolyte and 1,1,1,3,3,3-hexafluoro-2-propanol (HFP) as a stabiliser, the C-C coupling product 40 is obtained in a 51% yield (Equation (49)) [161,162,163]. It was postulated that HFP may help to prolong the amidyl radical intermediate lifetime. Alternatively, the HFP alkoxide generated in situ at the cathode electrode may help during the initial electron-transfer process [161,162,163].
Molecules 28 00471 i179
Molecules 28 00471 i180
Like o- or p-aminophenol derivatives, quinoneimine intermediates also are produced upon the oxidation of amino-protected o- or p-aminophenol derivatives. These quinoneimine intermediates are valuable electrophiles to perform different organic reactions. N-(4-hydroxyphenyl)-4-methylbenzenesulfonamide can be electrochemically oxidised at a graphite electrode in acetate buffer/acetonitrile solvent mixture to produce 41 via a two-electron two-proton process (Equation (50)) [164]. This quinoneimine can react with sodium benzenesulfinate to produce N-[4-hydroxy-5-(phenylsulfonyl)phenyl]benzenesulfonamide, 42, in a 55% yield. The presence of the amino-protecting group increases the regioselectivity of the reaction, happening exclusively at the ortho-position to the -OH group [164].
Molecules 28 00471 i181

3.5. Catalysed Oxidation of Aniline and Its Derivatives

The electrocatalytic oxidation of N,N-dialkylaniline in acetonitrile containing water, lutidine (as the base), NaClO4 as the supporting electrolyte, and TEMPO as the catalyst was reported to produce N-alkylformanilide and N-alkylaniline as a secondary product [165,166]. The electrochemically generated oxoammonium cation (43, Equation (51)) reacts with N,N-dialkylaniline, generating the iminium derivative 44 (Equation (52)). As previously mentioned, the iminium can hydrolyse in the presence of water, resulting in the formation of N-alkylaniline, 45 via an amino alcohol intermediate. Alternatively, it can be further oxidised for a second mol of 43 to form N-alkylformanilide derivative, 46 (Equation (53)), in a 75–92% yield, depending on the nature of the R group [165].
Molecules 28 00471 i182
Molecules 28 00471 i183
Molecules 28 00471 i184
Molecules 28 00471 i185

4. Conclusions

Over the past few decades, electrochemical anodic oxidation of amines has achieved remarkable advancement and shown great promise as a tool for organic transformations. The idea of creating radical precursors through electrode-initiated electron transfer provides a greener choice than conventional chemical reagents and can offer the possibility of using solvents in a sustainable form. In this work, we discussed the electrochemical oxidation mechanism of aliphatic amines, amides, aniline and aniline derivatives, carbamates, and lactams, either directly oxidised at different electrode surfaces or indirectly oxidised by a reversible redox molecule, in which the reactive form was generated in situ. A common feature in the oxidation mechanism of these amines is that the reaction starts with the nitrogen loss of one electron to yield a radical cation, which is stabilised by a following deprotonation step. This means that hydrogen at the α-carbon to the amine, directly connected to the amine, or at the aromatic ring becomes acidic in this process, facilitating α-substitution, radical–radical dimerisation, or nucleophilic aromatic substitution reaction, depending on the substituents present on the amine molecule under study and the experimental conditions. Meanwhile, if the formation of an iminium in equilibrium with its enamine form is possible, then this last intermediate can react with electrophilic groups, facilitating β-substitutions.
This work could help readers understand the comparative similarities and differences in the anodic oxidation mechanism of different amines.

Author Contributions

Conceptualization, A.K.V.M. and A.A.J.T.; methodology, A.K.V.M. and A.A.J.T.; software, A.K.V.M. and A.A.J.T.; validation, A.K.V.M. and A.A.J.T.; formal analysis, A.K.V.M. and A.A.J.T.; investigation, A.K.V.M. and A.A.J.T.; resources, A.K.V.M. and A.A.J.T.; data curation, A.K.V.M. and A.A.J.T.; writing—original draft preparation, A.K.V.M. and A.A.J.T.; writing—review and editing, A.K.V.M. and A.A.J.T.; visualization, A.K.V.M. and A.A.J.T.; supervision, A.K.V.M. and A.A.J.T.; project administration, A.K.V.M. and A.A.J.T.; funding acquisition, A.K.V.M. and A.A.J.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lund, H.; Hammerich, O. Organic Electrochemistry, 4th ed.; Marcel Dekker: New York, NY, USA, 2001. [Google Scholar]
  2. Bard, A.J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and Applications; Wiley: New York, NY, USA, 2001. [Google Scholar]
  3. Najmi, A.A.; Bischoff, R.; Permentier, H.P. N-dealkylation of amines. Molecules 2022, 27, 3293. [Google Scholar] [CrossRef]
  4. Steckhan, E. Anodic oxidation of nitrogen-containing compounds. In Organic Electrochemistry, 4th ed.; Lund, H., Hammerich, O., Eds.; Marcel Dekker, Inc.: New York, NY, USA, 2001. [Google Scholar]
  5. Chen, N.; Xu, H.-C. Electrochemical generation of nitrogen-centered radicals for organic synthesis. Green Synth. Catal. 2021, 2, 165–178. [Google Scholar] [CrossRef]
  6. Kärkäs, M.D. Electrochemical strategies for C–H functionalization and C–N bond formation. Chem. Soc. Rev. 2018, 47, 5786–5865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Shono, T. Electroorganic chemistry in organic synthesis. Tetrahedron 1984, 40, 811–850. [Google Scholar] [CrossRef]
  8. Jaworski, J.S. Organic Electrochemistry: Revised and Expanded; Taylor & Francis Group: Boca Raton, FL, USA, 2015. [Google Scholar]
  9. Barnes, K.K.; Mann, C.K. Electrochemical oxidation of primary aliphatic amines. J. Org. Chem. 1967, 32, 1474–1479. [Google Scholar] [CrossRef]
  10. Mann, C.K. Cyclic stationary electrode voltammetry of some aliphatic amines. Anal. Chem. 1964, 36, 2424–2426. [Google Scholar] [CrossRef]
  11. O’Donnell, J.F.; Mann, C.K. Determination of aliphatic amides by controlled-potential coulometry. J. Electroanal. Chem. Interfacial Electrochem. 1967, 13, 163–166. [Google Scholar] [CrossRef]
  12. O’Donnell, J.F.; Mann, C.K. Controlled-potential oxidation of aliphatic amides. J. Electroanal. Chem. Interfacial Electrochem. 1967, 13, 157–162. [Google Scholar] [CrossRef]
  13. Portis, L.C.; Bhat, V.V.; Mann, C.K. Electrochemical dealkylation of aliphatic tertiary and secondary amines. J. Org. Chem. 1970, 35, 2175–2178. [Google Scholar] [CrossRef]
  14. Smith, P.J.; Mann, C.K. Anodic methoxylation of tertiary amines. J. Org. Chem. 1968, 33, 316–317. [Google Scholar] [CrossRef]
  15. Smith, P.J.; Mann, C.K. Electrochemical dealkylation of aliphatic amines. J. Org. Chem. 1969, 34, 1821–1826. [Google Scholar] [CrossRef]
  16. Adenier, A.; Chehimi, M.M.; Gallardo, I.; Pinson, J.; Vila, N. Electrochemical oxidation of aliphatic amines and their attachment to carbon and metal surfaces. Langmuir 2004, 20, 8243–8253. [Google Scholar] [CrossRef]
  17. Torriero, A.A.J.; Morda, J.; Saw, J. Electrocatalytic dealkylation of amines mediated by ferrocene. Organometallics 2019, 38, 4280–4287. [Google Scholar] [CrossRef]
  18. Torriero, A.A.J.; Shiddiky, M.J.A.; Burgar, I.; Bond, A.M. Homogeneous electron-transfer reaction between electrochemically generated ferrocenium ions and amine containing compounds. Organometallics 2013, 32, 5731–5739. [Google Scholar] [CrossRef]
  19. Bock, H.; Goebel, I.; Havlas, Z.; Liedle, S.; Oberhammer, H. Triisopropylamine: A sterically overcrowded molecule with a flattened nc3 pyramid and a “p-type” nitrogen electron pair. Angew. Chem. Int. Ed. Engl. 1991, 30, 187–190. [Google Scholar] [CrossRef]
  20. Deinhammer, R.S.; Ho, M.; Anderegg, J.W.; Porter, M.D. Electrochemical oxidation of amine-containing compounds: A route to the surface modification of glassy carbon electrodes. Langmuir 1994, 10, 1306–1313. [Google Scholar] [CrossRef]
  21. Bauer, R.; Wendt, H. Anodic N-N Linkage by oxidation of anions of primary and secondary amines. Angew. Chem. Int. Ed. Engl. 1978, 17, 202–203. [Google Scholar] [CrossRef]
  22. Loveland, J.W.; Dimeler, G.R. Anodic voltammetry to +2.0 Volts. Application to hydrocarbons and oxidation stability studies. Anal. Chem. 1961, 33, 1196–1201. [Google Scholar] [CrossRef]
  23. Wawzonek, S.; McIntyre, T.W. Electrolytic oxidation of aromatic amines. J. Electrochem. Soc. 1967, 114, 1025–1029. [Google Scholar] [CrossRef]
  24. Dvořák, V.; Němec, I.; Zýka, J. Electrochemical oxidation of some aromatic amines in acetonitrile medium. I. N,N-dimethylaniline, triphenylamine, diphenylamine and di-4-tolylamine. Microchem. J. 1967, 12, 99–116. [Google Scholar] [CrossRef]
  25. Seo, E.T.; Nelson, R.F.; Fritsch, J.M.; Marcoux, L.S.; Leedy, D.W.; Adams, R.N. Anodic oxidation pathways of aromatic amines. Electrochemical and electron paramagnetic resonance studies. J. Am. Chem. Soc. 1966, 88, 3498–3503. [Google Scholar] [CrossRef]
  26. Yang, H.; Wipf, D.O.; Bard, A.J. Application of rapid scan cyclic voltammetry to a study of the oxidation and dimerization of N,N-dimethylaniline in acetonitrile. J. Electroanal. Chem. 1992, 331, 913–924. [Google Scholar] [CrossRef]
  27. Zweig, A.; Maurer, A.H.; Roberts, B.G. Oxidation, reduction, and electrochemiluminescence of donor-substituted polycyclic aromatic hydrocarbons. J. Org. Chem. 1967, 32, 1322–1329. [Google Scholar] [CrossRef]
  28. Zweig, A.; Lancaster, J.E.; Neglia, M.T.; Jura, W.H. Cumulative influence of dimethylamino groups on the π-system properties of aromatic hydrocarbons. J. Am. Chem. Soc. 1964, 86, 4130–4136. [Google Scholar] [CrossRef]
  29. Novak Jovanović, I.; Jadreško, D.; Miličević, A.; Hranjec, M.; Perin, N. An electrochemical study on the redox chemistry of cyclic benzimidazole derivatives with potent anticancer activity. Electrochim. Acta 2019, 297, 452–462. [Google Scholar] [CrossRef]
  30. Ross, S.D. The mechanism of anodic dealkylation of aliphatic amines in acetonitrile. Tetrahedron Lett. 1973, 14, 1237–1240. [Google Scholar] [CrossRef]
  31. Barry, J.E.; Finkelstein, M.; Mayeda, E.A.; Ross, S.D. Products and mechanisms in the anodic oxidation of N,N-dimethylbenzylamine in methanol. J. Org. Chem. 1974, 39, 2695–2699. [Google Scholar] [CrossRef]
  32. Weinberg, N.L. Anodic oxidation of organic compounds. IV. Mechanism of electrochemical methoxylation of N,N-dimethylbenzylamine. J. Org. Chem. 1968, 33, 4326–4329. [Google Scholar] [CrossRef]
  33. Becker, J.Y.; Miller, L.L.; Stermitz, F.R. Electrode processes in the oxidation of tetrahydroisoquinoline derivatives. J. Electroanal. Chem. Interfacial Electrochem. 1976, 68, 181–191. [Google Scholar] [CrossRef]
  34. Lee, K.; Choi, H.; An, J.; Kwon, K.-Y. Stainless steel promoted the electrochemical oxidation of amines into imines. Bull. Korean Chem. Soc. 2022, 43, 937–940. [Google Scholar] [CrossRef]
  35. Nelsen, S.F.; Ippoliti, J.T. The deprotonation of trialkylamine cation radicals by amines. J. Am. Chem. Soc. 1986, 108, 4879–4881. [Google Scholar] [CrossRef]
  36. Nelsen, S.F.; Kessel, C.R. 9-butylazabicyclo[3.3.1]nonane radical cation, the first long-lived saturated amine radical cation. J. Chem. Soc. Chem. Commun. 1977, 490–491. [Google Scholar] [CrossRef]
  37. Weinberg, N.L.; Brown, E.A. The anodic oxidation of organic compounds. II. The electrochemical alkoxylation of tertiary amines. J. Org. Chem. 1966, 31, 4058–4061. [Google Scholar] [CrossRef]
  38. Chiba, T.; Takata, Y. Anodic cyanation of tertiary aliphatic and heterocyclic amines. J. Org. Chem. 1977, 42, 2973–2977. [Google Scholar] [CrossRef]
  39. Bidan, G.; Genies, M. Utilisation en synthese du cation iminium genere in situ par oxydation electrochimique d’amines tertiaires. Tetrahedron 1981, 37, 2297–2301. [Google Scholar] [CrossRef]
  40. Alipour Najmi, A.; Xiao, Z.; Bischoff, R.; Dekker, F.J.; Permentier, H.P. Electrochemical N-demethylation of tropane alkaloids. Green Chem. 2020, 22, 6455–6463. [Google Scholar] [CrossRef]
  41. Tan, G.-H.; Lim, T.-M.; Kam, T.-S. Electrochemical oxidation of aspidofractinine-type indole alkaloids. A facile, electrochemically-mediated conversion of kopsingine to kopsidines A, B, C, and kopsinganol. Tetrahedron Lett. 1995, 36, 1327–1330. [Google Scholar] [CrossRef]
  42. Kam, T.-S.; Lim, T.-M.; Tan, G.-H. Electrochemical oxidation of aspidofractinine-type alkaloids: Formation of kopsine, kopsidine, kopsinitarine and bisindole derivatives. J. Chem. Soc. Perkin Trans. 2001, 1, 1594–1604. [Google Scholar] [CrossRef]
  43. Shono, T.; Matsumura, Y.; Tsubata, K. Electroorganic chemistry. 46. A new carbon-carbon bond forming reaction at the.alpha.-position of amines utilizing anodic oxidation as a key step. J. Am. Chem. Soc. 1981, 103, 1172–1176. [Google Scholar] [CrossRef]
  44. Ross, S.D.; Finkelstein, M.; Petersen, R.C. Anodic oxidations. III. The reaction mechanism in the electrochemical acetoxylation and alkoxylation of N,N-dimethylamides. J. Am. Chem. Soc. 1966, 88, 4657–4660. [Google Scholar] [CrossRef]
  45. Okita, M.; Wakamatsu, T.; Ban, Y. Anodic oxidation of N-alkyl-lactams. J. Chem. Soc. Chem. Commun. 1979, 749. [Google Scholar] [CrossRef]
  46. Asher, V.; Becu, C.; Anteunis, M.J.O.; Callens, R. New synthesis of pipecolic acid and analogs. Tetrahedron Lett. 1981, 22, 141–144. [Google Scholar] [CrossRef]
  47. Cornille, F.; Slomczynska, U.; Smythe, M.L.; Beusen, D.D.; Moeller, K.D.; Marshall, G.R. Electrochemical cyclization of dipeptides toward novel bicyclic, reverse-turn peptidomimetics. 1. Synthesis and conformational analysis of 7,5-bicyclic systems. J. Am. Chem. Soc. 1995, 117, 909–917. [Google Scholar] [CrossRef]
  48. Danielmeier, K.; Schierle, K.; Steckhan, E. A new chiral, cationic β-amino alcohol equivalent: A variable approach to enantiomerically pure building blocks for hydroxyethylene isosters. Angew. Chem. Int. Ed. Engl. 1996, 35, 2247–2248. [Google Scholar] [CrossRef]
  49. Li, W.; Hanau, C.E.; dAvignon, A.; Moeller, K.D. Conformationally restricted peptide mimetics: The incorporation of 6,5-bicyclic lactam ring skeletons into peptides. J. Org. Chem. 1995, 60, 8155–8170. [Google Scholar] [CrossRef]
  50. Moeller, K.D.; Rothfus, S.L.; Wong, P.L. Anodic amide oxidations: Fundamental studies concerning the annulation of six- and seven-membered rings onto amines. Tetrahedron 1991, 47, 583–592. [Google Scholar] [CrossRef]
  51. Papadopoulos, A.; Lewall, B.; Steckhan, E.; Ginzel, K.-D.; Knoch, F.; Nieger, M. Anodic oxidation of N-acyl and N-alkoxycarbonyl dipeptide esters as a key step for the formation of chiral heterocyclic synthetic building blocks. Tetrahedron 1991, 47, 563–572. [Google Scholar] [CrossRef]
  52. Rudd, E.J.; Finkelstein, M.; Ross, S.D. Anodic oxidations. VII. Reaction mechanism in the electrochemical oxidation of N,N-dimethylformamide in acetic acid and in methanol. J. Org. Chem. 1972, 37, 1763–1767. [Google Scholar] [CrossRef]
  53. Zhang, L.; Su, J.-H.; Wang, S.; Wan, C.; Zha, Z.; Du, J.; Wang, Z. Direct electrochemical imidation of aliphatic amines via anodic oxidation. Chem. Commun. 2011, 47, 5488–5490. [Google Scholar] [CrossRef]
  54. Gad, S.C. Ethyleneimine. In Encyclopedia of Toxicology (Third Edition); Wexler, P., Ed.; Academic Press: Oxford, UK, 2014; pp. 532–534. [Google Scholar]
  55. Gassman, P.G.; Nishiguchi, I.; Yamamoto, H. Anodic oxidation of aziridines. J. Am. Chem. Soc. 1975, 97, 1600–1602. [Google Scholar] [CrossRef]
  56. Torriero, A.A.J. Electrochemistry in Ionic Liquids. Volume 1: Fundamentals; Springer: Cham, Switzerland, 2015. [Google Scholar]
  57. Torriero, A.A.J.; Feldberg, S.; Zhang, J.; Simonov, A.; Bond, A. On choosing a reference redox system for electrochemical measurements: A cautionary tale. J. Solid State Electrochem. 2013, 17, 3021–3026. [Google Scholar] [CrossRef]
  58. Gao, Z.-N.; Zhang, J.; Liu, W.-Y. Electrocatalytic oxidation of N-acetyl-L-cysteine by acetylferrocene at glassy carbon electrode. Electroanal. Chem. Interfacial Electrochem. 2005, 580, 9–16. [Google Scholar] [CrossRef]
  59. Largeron, M.; Neudorffer, A.; Benoit, M.; Fleury, M.-B. Oxidation of primary aliphatic amines catalyzed by an electrogenerated quinonoid species: A small molecule mimic of amine oxidases. Proc. Electrochem. Soc. 2003, 12, 185–188. [Google Scholar]
  60. Shiddiky, M.J.A.; Torriero, A.A.J.; Zeng, Z.; Spiccia, L.; Bond, A.M. Highly selective and sensitive DNA assay based on electrocatalytic oxidation of ferrocene bearing zinc(ii)-cyclen complexes with diethylamine. J. Am. Chem. Soc. 2010, 132, 10053–10063. [Google Scholar] [CrossRef] [PubMed]
  61. Togni, A.; Hayashi, T. Ferrocenes: Homogeneous Catalysis, Organic Synthesis, Materials Science; Wiley: New York, NY, USA, 2008. [Google Scholar]
  62. Fleischmann, M.; Korinek, K.; Pletcher, D. The oxidation of organic compounds at a nickel anode in alkaline solution. J. Electroanal. Chem. Interfacial Electrochem. 1971, 31, 39–49. [Google Scholar] [CrossRef]
  63. Torii, S.; Inokuchi, T.; Takagishi, S.; Sato, E.; Tsujiyama, H. Conversion of cyclic aziridines and 2-amino-1-cycloalkanols to keto nitriles by using electrogenerated reactive halogen species. Chem. Lett. 1987, 16, 1469–1472. [Google Scholar] [CrossRef]
  64. Bishop, R. 6.07 Ritter-type reactions. In Comprehensive Organic Synthesis, 2nd ed.; Knochel, P., Ed.; Elsevier: Amsterdam, The Netherlands, 2014; pp. 239–295. [Google Scholar]
  65. Shono, T.; Matsumura, Y.; Hayashi, J.; Usui, M.; Yamane, S.-I.; Inoue, K. Electroorganic chemistry. 66. Electrochemical oxidation of aminals and enamines using a mediatory system. Acta Chem. Scand. 1983, B37, 491–498. [Google Scholar] [CrossRef] [Green Version]
  66. Torii, S.; Sayo, N.; Tanaka, H. Electrosynthesis of heteroatom bonds. 5. Direct cross-coupling of dialkylphosphites with amines by an iodonium ion-promoted electrolytic procedure. Tetrahedron Lett. 1979, 20, 4471–4474. [Google Scholar] [CrossRef]
  67. Okimoto, M.; Takahashi, Y.; Numata, K.; Nagata, Y.; Sasaki, G. Electrochemical oxidation of benzylic amines into the corresponding imines in the presence of catalytic amounts of KI. Synth. Commun. 2005, 35, 1989–1995. [Google Scholar] [CrossRef]
  68. Billon-Souquet, F.; Martens, T.; Royer, J. Anodic oxidation of the N-cyanomethyloxazolidine system. Tetrahedron 1996, 52, 15127–15136. [Google Scholar] [CrossRef]
  69. Torii, S.; Tanaka, H.; Hamano, S.-i.; Tada, N.; Nokami, J.; Sasaoka, M. MgBr2-promoted electrosynthesis of sulfenimines (thiooximes) from α-aminoalkanoates and diaryl/dialkyl disulfides in a CH2Cl2–H2O two-phase system. a straightforward preparation of C(6)/C(7)-sulfenimine derivatives of penicillin and cephalosporin. Chem. Lett. 1984, 13, 1823–1826. [Google Scholar] [CrossRef]
  70. Shono, T.; Matsumura, Y.; Inoue, K. Electroorganic chemistry 71. Anodic alpha-methoxylation of N-carbomethoxylated or N-acylated alpha-amino acid esters and alpha-amino-beta-lactams. J. Org. Chem. 1983, 48, 1388–1389. [Google Scholar] [CrossRef]
  71. Papadopoulos, A.; Heyer, J.; Ginzel, K.-D.; Steckhan, E. Regioselective anodic oxidation of N-acyl, N-alkoxycarbonyl, and N-(2-Nitrophenylsulfenyl) dipeptide esters. Chem. Ber. 1989, 122, 2159–2164. [Google Scholar] [CrossRef]
  72. Cãproiu, M.; Florea, C.; Galli, C.; Petride, A.; Petride, H. Oxidation of n-benzyl aziridine by molecular iodine: Competition of electron transfer and heterolytic pathways. Eur. J. Org. Chem. 2000, 2000, 1037–1043. [Google Scholar] [CrossRef]
  73. Kossaï, R.; Simonet, J.; Dauphin, G. The anodic tetramerization of the N-benzylaziridine: A chain process. Tetrahedron Lett. 1980, 21, 3575–3578. [Google Scholar] [CrossRef]
  74. Cuppoletti, A.; Dagostin, C.; Florea, C.; Galli, C.; Gentili, P.; Lanzalunga, O.; Petride, A.; Petride, H. The oxidation of n-benzylaziridine catalyzed by iron porphyrin: Radical versus electron transfer mechanism. Chem.–A Eur. J. 1999, 5, 2993–2999. [Google Scholar] [CrossRef]
  75. Largeron, M.; Chiaroni, A.; Fleury, M.-B. Environmentally friendly chemoselective oxidation of primary aliphatic amines by using a biomimetic electrocatalytic system. Chem.–A Eur. J. 2008, 14, 996–1003. [Google Scholar] [CrossRef]
  76. Dapperheld, S.; Steckhan, E.; Brinkhaus, K.-H.G.; Esch, T. Organic electron transfer systems, ii substituted triarylamine cation-radical redox systems—Synthesis, electrochemical and spectroscopic properties, Hammet behavior, and suitability as redox catalysts. Chem. Ber. 1991, 124, 2557–2567. [Google Scholar] [CrossRef]
  77. Schmidt, W.; Steckhan, E. Über organische elektronenüberträgersysteme, I. Elektrochemische und spektroskopische untersuchung bromsubstituierter triarylamin-redoxsysteme. Chem. Ber. 1980, 113, 577–585. [Google Scholar] [CrossRef]
  78. Steckhan, E. Indirect electroorganic syntheses—A modern chapter of organic electrochemistry. Angew. Chem. Int. Ed. Engl. 1986, 25, 683–701. [Google Scholar] [CrossRef]
  79. Heyer, J.; Dapperheld, S.; Steckhan, E. Convenient syntheses of 2-nitrophenylsulfen-(NPS-)imines by oxidation of NPS-protected amines and amino acid derivatives. Chem. Ber. 1988, 121, 1617–1623. [Google Scholar] [CrossRef]
  80. Pletcher, D.; Zappi, G.D. The indirect anodic oxidation of amines mediated by brominated aryl amines. J. Electroanal. Chem. Interfacial Electrochem. 1989, 265, 203–213. [Google Scholar] [CrossRef]
  81. Diamond, S.E.; Taube, H. Nucleophilic attack on cyanoformate induced by co-ordination to ruthenium ammines. J. Chem. Soc. Chem. Commun. 1974, 622–623. [Google Scholar] [CrossRef]
  82. Diamond, S.E.; Tom, G.M.; Taube, H. Ruthenium promoted oxidation of amines. J. Am. Chem. Soc. 1975, 97, 2661–2664. [Google Scholar] [CrossRef]
  83. Kwok, K.K.; Vincent, E.C.; Gibson, J.N. 36—Antineoplastic drugs. In Pharmacology and Therapeutics for Dentistry, 7th ed.; Dowd, F.J., Johnson, B.S., Mariotti, A.J., Eds.; Mosby: Maryland Heights, MI, USA, 2017; pp. 530–562. [Google Scholar]
  84. Seneca, Chapter 2-alkaloid chemistry. In Alkaloids-Secrets of Life; Aniszewski, T. (Ed.) Elsevier: Amsterdam, The Netherlands, 2007; pp. 61–139. [Google Scholar]
  85. Tabakovic, I.; Gunic, E.; Gasic, M.J. Anodic C16–C21 fragmentation of catharanthine in methanol. Synthesis of 16-methoxycleavamine. J. Chem. Soc. Perkin Trans. 1996, 2, 2741–2745. [Google Scholar] [CrossRef]
  86. Najmi, A.A.; Bhat, M.F.; Bischoff, R.; Poelarends, G.J.; Permentier, H.P. TEMPO-mediated electrochemical N-demethylation of opiate alkaloids. ChemElectroChem 2021, 8, 2590–2596. [Google Scholar] [CrossRef]
  87. Shono, T.; Toda, T.; Oshino, N. Electron transfer from nitrogen in microsomal oxidation of amine and amide. Simulation of microsomal oxidation by anodic oxidation. J. Am. Chem. Soc. 1982, 104, 2639–2641. [Google Scholar] [CrossRef]
  88. Shono, T. Synthesis of Alkaloidal Compounds Using an Electrochemical Reaction as a Key Step; Stekchan, E., Ed.; Electrochemistry III: Berlin/Heidelberg, Germany; Springer: Berlin/Heidelberg, Germany, 1988; pp. 131–151. [Google Scholar]
  89. Portis, L.C.; Klug, J.T.; Mann, C.K. Electrochemical oxidation of some phenethylamines. J. Org. Chem. 1974, 39, 3488–3494. [Google Scholar] [CrossRef]
  90. Masui, M.; Kamada, Y.; Ozaki, S. Anodic oxidation of amines. VI. Revised reaction scheme for the oxidation of ephedrine. Chem. Pharm. Bull. 1980, 28, 1619–1622. [Google Scholar] [CrossRef] [Green Version]
  91. Masui, M.; Kamada, Y.; Sasaki, E.; Ozaki, S. Anodic oxidation of amines. VII. Oxidation of beta-alkanolamines in aqueous buffer of pH 10. Chem. Pharm. Bull. 1982, 30, 1234–1243. [Google Scholar] [CrossRef] [Green Version]
  92. Shono, T.; Hamaguchi, H.; Matsumura, Y.; Yoshida, K. Syntheses of carbonyl compounds by the anodic cleavage of glycols and related compounds. Tetrahedron Lett. 1977, 18, 3625–3628. [Google Scholar] [CrossRef]
  93. Célimène, C.; Dhimane, H.; Le Bail, M.; Lhommet, G. A short and highly stereocontrolled total synthesis of (3R,5R,8aR)-3-n-butyl-5-methylindolizidine. Tetrahedron Lett. 1994, 35, 6105–6106. [Google Scholar] [CrossRef]
  94. Beal, L.M.; Moeller, K.D. A sequential electrochemical oxidation—olefin metathesis strategy for the construction of bicyclic lactam based peptidomimetics. Tetrahedron Lett. 1998, 39, 4639–4642. [Google Scholar] [CrossRef]
  95. Finkelstein, M.; Ross, S.D. Anodic oxidation of N-methylformamide and N-methylacetamide. Tetrahedron 1972, 28, 4497–4502. [Google Scholar] [CrossRef]
  96. Kardassis, G.; Brungs, P.; Nothhelfer, C.; Steckhan, E. Electrogenerated chiral cationic glycine equivalents—Part 2: Chiral 3-methoxy-2,5-morpholinediones from (S)-α-hydroxy acids and dimethyl aminomalonate. Tetrahedron 1998, 54, 3479–3488. [Google Scholar] [CrossRef]
  97. Kardassis, G.; Brungs, P.; Steckhan, E. Electrogenerated chiral cationic glycine equivalents—Part 1: The 6-methoxy derivative of cyclo(L-Pro-Gly). Tetrahedron 1998, 54, 3471–3478. [Google Scholar] [CrossRef]
  98. Moeller, K.D.; Rutledge, L.D. Anodic amide oxidations: The synthesis of two spirocyclic L-pyroglutamide building blocks. J. Org. Chem. 1992, 57, 6360–6363. [Google Scholar] [CrossRef]
  99. Moeller, K.D.; Sharif, T.; Mohammad, R.M. Electrochemical amide oxidations in the presence of monomethoxylated phenyl rings. An unexpected relationship between the chemoselectivity of the oxidation and the location of the methoxy substituent. Tetrahedron Lett. 1989, 30, 1213–1216. [Google Scholar] [CrossRef]
  100. Moeller, K.D.; Wang, P.W.; Tarazi, S.; Marzabadi, M.R.; Wong, P.L. Anodic amide oxidations in the presence of electron-rich phenyl rings: Evidence for an intramolecular electron-transfer mechanism. J. Org. Chem. 1991, 56, 1058–1067. [Google Scholar] [CrossRef]
  101. Schierle, K.; Vahle, R.; Steckhan, E. Efficient pathways to precursors of pyrrolidine azasugar structures. Eur. J. Org. Chem. 1998, 1998, 509–514. [Google Scholar] [CrossRef]
  102. Skrinjar, M.; Wistrand, L.-G. Highly stereoselective addition of alkylcopper reagents to n-acyliminium ions. Enantioselective synthesis of trans-2- butyl-5-heptylpyrrolidine. Tetrahedron Lett. 1990, 31, 1775–1778. [Google Scholar] [CrossRef]
  103. Slomczynska, U.; Chalmers, D.K.; Cornille, F.; Smythe, M.L.; Beusen, D.D.; Moeller, K.D.; Marshall, G.R. Electrochemical Cyclization of Dipeptides to Form Novel Bicyclic, Reverse-Turn Peptidomimetics. 2. Synthesis and Conformational Analysis of 6,5-Bicyclic Systems. J. Org. Chem. 1996, 61, 1198–1204. [Google Scholar] [CrossRef]
  104. Thaning, M.; Wistrand, L.-G. Synthetic applications of electrochemically produced α-methoxyamides. Part 2. Oxidation of hydroxyproline derivatives. Helv. Chim. Acta 1986, 69, 1711–1717. [Google Scholar] [CrossRef]
  105. Thaning, M.; Wistrand, L.G. Synthetic applications of electrochemically produced .alpha.-methoxy amides. 5. A short enantiodivergent synthesis of the Geissman-Waiss lactone. J. Org. Chem. 1990, 55, 1406–1408. [Google Scholar] [CrossRef]
  106. Wistrand, L.-G.; Skrinjar, M. Chirospecific synthesis of trans-2,5-disubstituted pyrrolidines via stereoselective addition of organocopper reagents to N-acyliminium ions. Tetrahedron 1991, 47, 573–582. [Google Scholar] [CrossRef]
  107. Wong, P.L.; Moeller, K.D. Anodic amide oxidations: Total syntheses of (-)-A58365A and (+-)-A58365B. J. Am. Chem. Soc. 1993, 115, 11434–11445. [Google Scholar] [CrossRef]
  108. Xu, Z.; Huang, Z.; Li, Y.; Kuniyil, R.; Zhang, C.; Ackermann, L.; Ruan, Z. Catalyst-free, direct electrochemical synthesis of annulated medium-sized lactams through C–C bond cleavage. Green Chem. 2020, 22, 1099–1104. [Google Scholar] [CrossRef]
  109. Yao, H.; Sherer, E.C.; Lu, M.; Small, J.; Martin, G.E.; Lam, Y.-h.; Chen, Q.; Helmy, R.; Liu, Y.; Chen, H. One-step regio- and stereoselective electrochemical synthesis of orexin receptor antagonist oxidative metabolites. J. Org. Chem. 2022, 87, 15011–15021. [Google Scholar] [CrossRef]
  110. Mikami, R.; Nakamura, Y.; Shida, N.; Atobe, M. Anodic substitution reaction of carbamates in a flow microreactor using a stable emulsion solution. React. Chem. Eng. 2021, 6, 2024–2028. [Google Scholar] [CrossRef]
  111. Novaes, L.F.T.; Ho, J.S.K.; Mao, K.; Liu, K.; Tanwar, M.; Neurock, M.; Villemure, E.; Terrett, J.A.; Lin, S. Exploring electrochemical C(sp3)–H oxidation for the late-stage methylation of complex molecules. J. Am. Chem. Soc. 2022, 144, 1187–1197. [Google Scholar] [CrossRef]
  112. Shono, T.; Hamaguchi, H.; Matsumura, Y. Electroorganic chemistry. XX. Anodic oxidation of carbamates. J. Am. Chem. Soc. 1975, 97, 4264–4268. [Google Scholar] [CrossRef]
  113. Shono, T.; Matsumura, Y.; Onomura, O.; Yamada, Y. A new method for regioselective synthesis of 2-substituted 1-(methoxycarbonyl)-1,2-dihydropyridines. Tetrahedron Lett. 1987, 28, 4073–4074. [Google Scholar] [CrossRef]
  114. Shono, T.; Matsumura, Y.; Tsubata, K.; Sugihara, Y. A new method of acylation at β-position of aliphatic amines. Tetrahedron Lett. 1982, 23, 1201–1204. [Google Scholar] [CrossRef]
  115. Shono, T.; Matsumura, Y.; Tsubata, K.; Sugihara, Y.; Yamane, S.; Kanazawa, T.; Aoki, T. Electroorganic chemistry. 60. Electroorganic synthesis of enamides and enecarbamates and their utilization in organic synthesis. J. Am. Chem. Soc. 1982, 104, 6697–6703. [Google Scholar] [CrossRef]
  116. Shono, T.; Terauchi, S.; Ohki, Y.; Matsumura, Y. Regioselective introduction of alkyl groups to the position α or γ to nitrogen atom of piperidine skeletons using anodic oxidation in a key step. Tetrahedron Lett. 1990, 31, 6385–6386. [Google Scholar] [CrossRef]
  117. Torii, S.; Inokuchi, T.; Kubota, M. A practical access to methyl 3,3-dimethoxypropionates and N-protected beta-aminoacrylates and beta-aminoacrylonitrile by using electrochemical procedure. J. Org. Chem. 1985, 50, 4157–4160. [Google Scholar] [CrossRef]
  118. Martínez, Á.M.; Hayrapetyan, D.; van Lingen, T.; Dyga, M.; Gooßen, L.J. Taking electrodecarboxylative etherification beyond Hofer–Moest using a radical C–O coupling strategy. Nat. Commun. 2020, 11, 4407. [Google Scholar] [CrossRef]
  119. Fuchigami, T.; Iwaoka, T.; Nonaka, T.; Sekine, T. Electrochemical oxidation of 3,3,7,7-tetraethyl-perhydro-1,5-diazocine. Chem. Lett. 1978, 7, 1437–1440. [Google Scholar] [CrossRef] [Green Version]
  120. Fuchigami, T.; Iwaoka, T.; Nonaka, T.; Sekine, T. Electrochemical generation of unstable nitrogen species. Part 4. Electrochemical and chemical oxidation of cyclic and open-chain diamines. The formation of cyclic hydrazine derivatives. Bull. Chem. Soc. Jpn. 1980, 53, 2040–2045. [Google Scholar] [CrossRef]
  121. Laube, B.L.; Asirvatham, M.R.; Mann, C.K. Electrochemical oxidation of tropanes. J. Org. Chem. 1977, 42, 670–674. [Google Scholar] [CrossRef]
  122. Fuchigami, T.; Sato, T.; Nonaka, T. Electrochemical generation of reactive nitrogen species. 10. Anodic amination of tetrahydrofuran. J. Org. Chem. 1986, 51, 366–369. [Google Scholar] [CrossRef]
  123. Frankowski, K.J.; Liu, R.; Milligan, G.L.; Moeller, K.D.; Aubé, J. Practical electrochemical anodic oxidation of polycyclic lactams for late stage functionalization. Angew. Chem. Int. Ed. 2015, 54, 10555–10558. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Masui, M.; Hara, S.; Ozaki, S. Anodic oxidation of amides and lactams using N-hydroxyphthalimide as a mediator. Chem. Pharm. Bull. 1986, 34, 975–979. [Google Scholar] [CrossRef] [Green Version]
  125. Warning, K.; Mitzlaff, M. Verfahren zur herstellung von ω,ω-dialkoxycarbonsaeureestern und -amiden. Tetrahedron Lett. 1979, 20, 1563–1564. [Google Scholar] [CrossRef]
  126. Yamamoto, K.; Kuriyama, M.; Onomura, O. Shono-type oxidation for functionalization of N-heterocycles. Chem. Rec. 2021, 21, 2239–2253. [Google Scholar] [CrossRef] [PubMed]
  127. Okita, M.; Mori, M.; Wakamatsu, T.; Ban, Y. Anodic oxidation of N-alkyl-β-lactams. Heterocycles 1985, 23, 247–250. [Google Scholar] [CrossRef]
  128. Mori, M.; Kagechika, K.; Sasai, H.; Shibasaki, M. New synthesis of 4-acetoxy-2-azetidinones by use of electrochemical oxidation. Tetrahedron 1991, 47, 531–540. [Google Scholar] [CrossRef]
  129. Weinberg, N.L.; Weinberg, H.R. Electrochemical oxidation of organic compounds. Chem. Rev. 1968, 68, 449–523. [Google Scholar] [CrossRef]
  130. Genies, E.M.; Lapkowski, M. Spectroelectrochemical evidence for an intermediate in the electropolymerization of aniline. J. Electroanal. Chem. Interfacial Electrochem. 1987, 236, 189–197. [Google Scholar] [CrossRef]
  131. Adams, R.N. Anodic oxidation pathways of aromatic hydrocarbons and amines. Acc. Chem. Res. 1969, 2, 175–180. [Google Scholar] [CrossRef]
  132. Hand, R.L.; Nelson, R.F. The anodic decomposition pathways of ortho- and meta-substituted anilines. J. Electrochem. Soc. 1978, 125, 1059–1069. [Google Scholar] [CrossRef]
  133. Hand, R.L.; Nelson, R.F. Anodic oxidation pathways of N-alkylanilines. J. Am. Chem. Soc. 1974, 96, 850–860. [Google Scholar] [CrossRef]
  134. Sharma, L.R.; Manchanda, A.K.; Singh, G.; Verma, R.S. Cyclic voltammetry of aromatic amines in aqueous and non-aqueous media. Electrochim. Acta 1982, 27, 223–233. [Google Scholar] [CrossRef]
  135. Wawzonek, S.; McIntyre, T.W. Electrolytic preparation of azobenzenes. J. Electrochem. Soc. 1972, 119, 1350. [Google Scholar] [CrossRef]
  136. Matsuda, Y.; Shono, A.; Iwakura, C.; Ohshiro, Y.; Agawa, T.; Tamura, H. Anodic oxidation of aniline in aqueous alkaline solution. Bull. Chem. Soc. Jpn. 1971, 44, 2960–2963. [Google Scholar] [CrossRef] [Green Version]
  137. Yang, H.; Bard, A.J. The application of rapid scan cyclic voltammetry and digital simulation to the study of the mechanism of diphenylamine oxidation, radical cation dimerization, and polymerization in acetonitrile. J. Electroanal. Chem. Interfacial Electrochem. 1991, 306, 87–109. [Google Scholar] [CrossRef]
  138. Andrieux, C.P.; Gallardo, I.; Junca, M. Mechanistic study of the electrochemical oxidation of some aromatic amines in the presence of bases. J. Electroanal. Chem. 1993, 354, 231–241. [Google Scholar] [CrossRef]
  139. Nishiyama, S.; Iida, Y.; Hino, S.; Yamamura, S. Anodic oxidation of N-alkylanilines. Electrochim. Acta 1997, 42, 1943–1949. [Google Scholar] [CrossRef]
  140. Yoshida, K.; Fueno, T. Anodic oxidations. VI. p-cyanation of diphenylamines. J. Org. Chem. 1972, 37, 4145–4147. [Google Scholar] [CrossRef]
  141. Mizoguchi, T.; Adams, R.N. Anodic oxidation studies of N,N-dimethylaniline. I. Voltammetric and spectroscopic investigations at platinum electrodes. J. Am. Chem. Soc. 1962, 84, 2058–2061. [Google Scholar] [CrossRef]
  142. Galus, Z.; Adams, R.N. Anodic oxidation studies of N,N-dimethylaniline. II. Stationary and rotated disk studies at inert electrodes. J. Am. Chem. Soc. 1962, 84, 2061–2065. [Google Scholar] [CrossRef]
  143. Galus, Z.; White, R.M.; Rowland, F.S.; Adams, R.N. Anodic oxidation studies of N,N-dimethylaniline. III. Tritium tracer studies of electrolysis products. J. Am. Chem. Soc. 1962, 84, 2065–2068. [Google Scholar] [CrossRef]
  144. Martins, D.; Garrido, E.M.P.J.; Borges, F.; Garrido, J.M.P.J. Voltammetric profiling of new psychoactive substances: Piperazine derivatives. J. Electroanal. Chem. 2021, 883, 115054. [Google Scholar] [CrossRef]
  145. Liu, K.; Tang, S.; Wu, T.; Wang, S.; Zou, M.; Cong, H.; Lei, A. Electrooxidative para-selective C–H/N–H cross-coupling with hydrogen evolution to synthesize triarylamine derivatives. Nat. Commun. 2019, 10, 639. [Google Scholar] [CrossRef]
  146. Nikl, J.; Ravelli, D.; Schollmeyer, D.; Waldvogel, S.R. Straightforward electrochemical sulfonylation of arenes and aniline derivatives using sodium sulfinates. ChemElectroChem 2019, 6, 4450–4455. [Google Scholar] [CrossRef] [Green Version]
  147. Yu, Y.; Fang, Y.; Tang, R.; Xu, D.; Dai, S.; Zhang, W. Electrochemical oxidative sulfonylation of N-arylamides/amine with sodium sulfinates. Asian J. Org. Chem. 2022, 11, e202100805. [Google Scholar] [CrossRef]
  148. Hand, R.; Melicharek, M.; Scoggin, D.; Stotz, R.; Carpenter, A.; Nelson, R. Electrochemical oxidation pathways of substituted dimethylanilines. Collect. Czechoslov. Chem. Commun. 1971, 36, 842–854. [Google Scholar] [CrossRef]
  149. Marcus, M.F.; Hawley, M.D. The electrochemical oxidation of p-dimethylaminophenol in aqueous solution. J. Electroanal. Chem. Interfacial Electrochem. 1968, 18, 175–183. [Google Scholar] [CrossRef]
  150. Melicharek, M.; Nelson, R.F. The electrochemical oxidation of N,N-dimethyl-p-toluidine. J. Electroanal. Chem. Interfacial Electrochem. 1970, 26, 201–209. [Google Scholar] [CrossRef]
  151. Rees, N.V.; Klymenko, O.V.; Compton, R.G.; Oyama, M. The electro-oxidation of N,N-dimethyl-p-toluidine in acetonitrile:: A microdisk voltammetry study. J. Electroanal. Chem. 2002, 531, 33–42. [Google Scholar] [CrossRef]
  152. Weinberg, N.L.; Reddy, T.B. Anodic oxidation of organic compounds. III. Effect of electrolyte on electrochemical methoxylation and dimerization of N,N-dimethylaniline. J. Am. Chem. Soc. 1968, 90, 91–94. [Google Scholar] [CrossRef]
  153. Shono, T.; Matsumura, Y.; Inoue, K.; Ohmizu, H.; Kashimura, S. Electroorganic chemistry. 62. Reaction of iminium ion with nucleophile: A versatile synthesis of tetrahydroquinolines and julolidines. J. Am. Chem. Soc. 1982, 104, 5753–5757. [Google Scholar] [CrossRef]
  154. Andreades, S.; Zahnow, E.W. Anodic cyanations of aromatic compounds. J. Am. Chem. Soc. 1969, 91, 4181–4190. [Google Scholar] [CrossRef]
  155. Michel, S.; Le Gall, E.; Hurvois, J.-P.; Moinet, C.; Tallec, A.; Uriac, P.; Toupet, L. Anodic cyanation of N-substituted 1-benzazepines: Synthesis of the corresponding α-aminonitriles. Liebigs Ann. 1997, 1997, 259–267. [Google Scholar] [CrossRef]
  156. Snead, W.K.; Remick, A.E. Studies on oxidation-reduction mechanism. II. The anodic oxidation of p-aminophenol. J. Am. Chem. Soc. 1957, 79, 6121–6127. [Google Scholar] [CrossRef]
  157. Hawley, D.; Adams, R.N. Homogeneous chemical reactions in electrode processes: Measurement of rates of follow-up chemical reactions. J. Electroanal. Chem. (1959) 1965, 10, 376–386. [Google Scholar] [CrossRef]
  158. Jackowska, K.; Bukowska, J.; Kudelski, A. Electro-oxidation of o-aminophenol studied by cyclic voltammetry and surface enhanced Raman scattering (SERS). J. Electroanal. Chem. 1993, 350, 177–187. [Google Scholar] [CrossRef]
  159. Amani, A.; Nematollahi, D. Electrochemical synthesis based on the oxidation of 1-(4-(4-hydroxyphenyl)piperazin-1-yl)ethanone in the presence of nucleophiles. J. Org. Chem. 2012, 77, 11302–11306. [Google Scholar] [CrossRef]
  160. Breising, V.M.; Kayser, J.M.; Kehl, A.; Schollmeyer, D.; Liermann, J.C.; Waldvogel, S.R. Electrochemical formation of N,N′-diarylhydrazines by dehydrogenative N–N homocoupling reaction. Chem. Commun. 2020, 56, 4348–4351. [Google Scholar] [CrossRef]
  161. Schulz, L.; Enders, M.; Elsler, B.; Schollmeyer, D.; Dyballa, K.M.; Franke, R.; Waldvogel, S.R. Reagent- and metal-free anodic C−C cross-coupling of aniline derivatives. Angew. Chem. Int. Ed. 2017, 56, 4877–4881. [Google Scholar] [CrossRef]
  162. Schulz, L.; Franke, R.; Waldvogel, S.R. Direct anodic dehydrogenative cross- and homo-coupling of formanilides. ChemElectroChem 2018, 5, 2069–2072. [Google Scholar] [CrossRef]
  163. Schulz, L.; Husmann, J.-Å.; Waldvogel, S.R. Outstandingly robust anodic dehydrogenative aniline coupling reaction. Electrochim. Acta 2020, 337, 135786. [Google Scholar] [CrossRef]
  164. Xiao, H.-L.; Yang, C.-W.; Zhang, N.-T.; Zeng, C.-C.; Hu, L.-M.; Tian, H.-Y.; Little, R.D. Electrochemical oxidation of aminophenols in the presence of benzenesulfinate. Tetrahedron 2013, 69, 658–663. [Google Scholar] [CrossRef]
  165. Kashiwagi, Y.; Anzai, J.-i. Selective electrocatalytic oxidation of N-alkyl-N-methylanilines to N-akylformanilides using nitroxyl radical. Chem. Pharm. Bull. 2001, 49, 324–326. [Google Scholar] [CrossRef] [Green Version]
  166. Hunter, D.H.; Racok, J.S.; Rey, A.W.; Zea-Ponce, Y. Oxoiminium ions for N-demethylation: 1-oxo-2,2,6,6-tetramethylpiperidinium chloride. J. Org. Chem. 1988, 53, 1278–1281. [Google Scholar] [CrossRef]
Table 1. Oxidation potential of different aliphatic and aromatic amines.
Table 1. Oxidation potential of different aliphatic and aromatic amines.
AmineSolventSupporting ElectrolyteWorking ElectrodeEox a
(V vs. SCE)
Ref.
propylamineCH3CN0.1 M Na[ClO4]Pt1.38[10]
butylamineDMF0.1 M [Bu4N][BF4]GC1.36[16]
EtOH0.1 M LiClO4GC1.22[20]
THF0.1 M LiClO4GC1.15[21]
pentylamineCH3CN0.1 M Na[ClO4]Pt1.45[10]
hexylamineDMF0.1 M [Bu4N][BF4]GC1.36[16]
nonylamineCH3CN0.1 M Na[ClO4]Pt1.48[10]
t-butylamineDMF0.1 M [Bu4N][BF4]GC1.44[16]
THF0.1 M LiClO4GC1.21[21]
CH3CN0.1 M Na[ClO4]Pt1.40[10]
butylamideTHF0.1 M LiClO4GC0.16[21]
t-butylamideTHF0.1 M LiClO4GC−0.10[21]
cyclohexylamineDMF0.1 M [Bu4N][BF4]GC1.39[16]
THF0.1 M LiClO4GC1.26[21]
cyclohexylamideTHF0.1 M LiClO4GC0.05[21]
N-methylacetamideCH3CN0.2 M NaClO4GC1.81[12]
N-acetylethylenediamineEtOH0.1 M LiClO4GC1.27[20]
dopamineEtOH0.1 M LiClO4GC1.22[20]
N-(5-aminopentyl)biotinamideEtOH0.1 M LiClO4GC1.22[20]
diethylamineCH3CN0.1 M [Bu4N][PF6]GC1.10[18]
dipropylamineCH3CN0.1 M NaClO4Pt1.00[10]
dibutylamineDMF0.1 M [Bu4N][BF4]GC1.11[16]
THF0.1 M LiClO4GC0.94[21]
CH3CN0.1 M NaClO4Pt1.07[10]
dibutylamideTHF0.1 M LiClO4GC−0.12[21]
dibenzylamineCH3CN0.1 M NaClO4Pt1.23[10]
di-isopropylamineDMF0.1 M [Bu4N][BF4]GC1.15[16]
di-isobutylamineDMF0.1 M [Bu4N][BF4]GC1.11[16]
di-sec-butylamineDMF0.1 M [Bu4N][BF4]GC1.16[16]
CH3CN0.1 M NaClO4Pt1.16[10]
dipentylamineCH3CN0.1 M NaClO4Pt1.11[10]
bis-2-ethylhexylamineDMF0.1 M [Bu4N][BF4]GC1.07[16]
N-methylbutylamineEtOH0.1 M LiClO4GC1.00[20]
N-ethylbutylamineEtOH0.1 M LiClO4GC0.99[20]
dicyclohexylamineDMF0.1 M [Bu4N][BF4]GC1.06[16]
CH3CN0.1 M [Bu4N][PF6]GC1.49[17]
N,N-dimethylacetamideCH3CN0.2 M NaClO4GC1.32[12]
trimethylamineCH3CN0.1 M NaClO4Pt1.05[10]
triethylamineEtOH0.1 M LiClO4GC0.83[20]
DMF0.1 M [Bu4N][BF4]GC0.94[16]
CH3CN0.1 M [Bu4N][PF6]GC0.88[18]
CH3CN0.1 M NaClO4Pt0.95[10]
tripropylamineDMF0.1 M [Bu4N][BF4]GC0.95[16]
CH3CN0.1 M NaClO4Pt0.93[10]
tributylamineDMF0.1 M [Bu4N][BF4]GC0.88[16]
CH3CN0.1 M NaClO4Pt0.78[10]
tripentylamineDMF0.1 M [Bu4N][BF4]GC0.91[16]
CH3CN0.1 M NaClO4Pt0.89[10]
tribenzylamineCH3CN0.1 M NaClO4Pt0.99[10]
tri-isopropylamineTHF0.1 M [Bu4N]ClO4GC0.76[19]
tri-isobutylamineDMF0.1 M [Bu4N][BF4]GC0.98[16]
N,N-dicyclohexylmethylamineCH3CN0.1 M [Bu4N][PF6]GC1.04[17]
N,N-dimethylcyclohexylamineCH3CN0.1 M [Bu4N][PF6]GC1.18[17]
N,N-dimethylbutylamineEtOH0.1 M LiClO4GC0.99[20]
4-nitrobenzylamineDMF0.1 M [Bu4N][BF4]GC1.42[16]
CH3CN0.1 M [Bu4N][BF4]GC1.58[16]
3-nitrobenzylamineDMF0.1 M [Bu4N][BF4]GC1.51[16]
CH3CN0.1 M [Bu4N][BF4]GC1.78[16]
N-methyl-3-nitrobenzylamineDMF0.1 M [Bu4N][BF4]GC1.25[16]
CH3CN0.1 M [Bu4N][BF4]GC1.33[16]
N,N-dimethyl-3-nitrobenzylamineDMF0.1 M [Bu4N][BF4]GC1.01[16]
CH3CN0.1 M [Bu4N][BF4]GC1.07[16]
pyrrolidineCH3CN0.1 M [Bu4N][PF6]GC1.16[18]
pyrroleCH3CN0.5 M NaClO4Pt1.06 *[22]
pyridineCH3CN0.5 M NaClO4Pt2.12 *[22]
N,N-dipropylpropionamideCH3CN0.2 M NaClO4GC1.26[12]
anilineCH3CN0.5 M NaClO4Pt0.90 *[23]
p-nitroanilineCH3CN0.5 M NaClO4Pt1.39 *[23]
p-bromoanilineCH3CN0.5 M NaClO4Pt0.97 *[23]
p-chloroanilineCH3CN0.5 M NaClO4Pt0.96 *[23]
p-anisidineCH3CN0.5 M NaClO4Pt0.62 *[23]
o-anisidineCH3CN0.5 M NaClO4Pt0.70 *[23]
diphenylamineCH3CN0.1 M NaClO4Pt0.83 *[24]
triphenylamineCH3CN0.1 M [Et4N]ClO4Pt0.98[25]
N,N-dimethylanilineCH3CN0.1 M [Bu4N][PF6]Pt0.76[26]
N,N-diethylanilineCH3CN0.5 M NaClO4Pt0.70 *[23]
N,N-diethyl-p-chloroanilineCH3CN0.5 M NaClO4Pt0.83 *[23]
N,N-dimethyl-p-chloroanilineCH3CN0.5 M NaClO4Pt0.85 *[23]
ethylphenylamineCH3CN0.5 M NaClO4Pt0.76 *[23]
di-4-tolylamineCH3CN0.1 M NaClO4Pt0.70 *[24]
N,N-tetramethylbenzidineCH3CN0.1 M [Pr4N]ClO4Pt0.43 *[27]
1-dimethylaminonaphthaleneCH3CN0.1 M [Pr4N]ClO4Pt0.75 *[27]
2-dimethylaminonaphthaleneCH3CN0.1 M [Pr4N]ClO4Pt0.67 *[27]
azobenzeneCH3CN0.5 M NaClO4Pt1.69 *[23]
4,4-dichloroazobenzeneCH3CN0.5 M NaClO4Pt1.80 *[23]
4,4-dimethoxyazobenzeneCH3CN0.5 M NaClO4Pt1.34 *[23]
N,N,N′,N′-tetramethyl-m-phenylenediamineCH3CN0.1 M [Pr4N]ClO4Pt0.62 *[28]
N,N,N′,N′-tetramethyl-p-phenylenediamineCH3CN0.1 M [Pr4N]ClO4Pt0.20 *[28]
N,N-dimethyl-m-anisidineCH3CN0.1 M [Pr4N]ClO4Pt0.79 *[28]
N,N-dimethyl-p-anisidineCH3CN0.1 M [Pr4N]ClO4Pt0.63 *[28]
3,4-dimethoxy-N,N-dimethylanilineCH3CN0.1 M [Pr4N]ClO4Pt0.50 *[28]
3,5-dimethoxy-N,N-dimethylanilineCH3CN0.1 M [Pr4N]ClO4Pt0.80 *[28]
N,N,N′,N′-tetramethyl-o-phenylenediamineCH3CN0.1 M [Pr4N]ClO4Pt0.58 *[28]
N,N-dimethyl-o-anisidineCH3CN0.1 M [Pr4N]ClO4Pt0.78 *[28]
2,4-dimethoxy-N,N-dimethylanilineCH3CN0.1 M [Pr4N]ClO4Pt0.57 *[28]
a oxidation peak potential is reported, except those with (*), which are the E1/2. Abbreviations: DMF = dimethylformamide; CH3CN = acetonitrile; EtOH = ethanol; THF = tetrahydrofuran; [Bu4N] = tetrabutylammonium; [Et4N] = tetraethylammonium; [Pr4N] = tetrapropylammonium; [PF6] = hexafluorophosphate; [BF4] = tetrafluoroborate; SCE = saturated calomel electrode.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mruthunjaya, A.K.V.; Torriero, A.A.J. Mechanistic Aspects of the Electrochemical Oxidation of Aliphatic Amines and Aniline Derivatives. Molecules 2023, 28, 471. https://doi.org/10.3390/molecules28020471

AMA Style

Mruthunjaya AKV, Torriero AAJ. Mechanistic Aspects of the Electrochemical Oxidation of Aliphatic Amines and Aniline Derivatives. Molecules. 2023; 28(2):471. https://doi.org/10.3390/molecules28020471

Chicago/Turabian Style

Mruthunjaya, Ashwin K. V., and Angel A. J. Torriero. 2023. "Mechanistic Aspects of the Electrochemical Oxidation of Aliphatic Amines and Aniline Derivatives" Molecules 28, no. 2: 471. https://doi.org/10.3390/molecules28020471

Article Metrics

Back to TopTop