Next Article in Journal
Hybrid Effect of Steel Bars and PAN Textile Reinforcement on Ductility of One-Way Slab Subjected to Bending
Next Article in Special Issue
Comprehensive Review on Medicinal Applications of Coumarin-Derived Imine–Metal Complexes
Previous Article in Journal
Unusual Derivatization of Methylmalonic Acid with Pentafluorobenzyl Bromide to a Tripentafluorobenzyl Derivative and Its Stable-Isotope Dilution GC-MS Measurement in Human Urine
Previous Article in Special Issue
Antimicrobial and Cytotoxic Activity of Novel Imidazolium-Based Ionic Liquids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Triflamides and Triflimides: Synthesis and Applications

by
Mikhail Y. Moskalik
* and
Vera V. Astakhova
Irkutsk Institute of Chemistry, Siberian Branch of the Russian Academy of Sciences, 664033 Irkutsk, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(16), 5201; https://doi.org/10.3390/molecules27165201
Submission received: 1 July 2022 / Revised: 10 August 2022 / Accepted: 11 August 2022 / Published: 15 August 2022
(This article belongs to the Special Issue Biologically Active Heterocyclic Compounds)

Abstract

:
Among the variety of sulfonamides, triflamides (CF3SO2NHR, TfNHR) occupy a special position in organic chemistry. Triflamides are widely used as reagents, efficient catalysts or additives in numerous reactions. The reasons for the widespread use of these compounds are their high NH-acidity, lipophilicity, catalytic activity and specific chemical properties. Their strong electron-withdrawing properties and low nucleophilicity, combined with their high NH-acidity, makes it possible to use triflamides in a vast variety of organic reactions. This review is devoted to the synthesis and use of N-trifluoromethanesulfonyl derivatives in organic chemistry, medicine, biochemistry, catalysis and agriculture. Part of the work is a review of areas and examples of the use of bis(trifluoromethanesulfonyl)imide (triflimide, (CF3SO2)2NH, Tf2NH). Being one of the strongest NH-acids, triflimide, and especially its salts, are widely used as catalysts in cycloaddition reactions, Friedel–Crafts reactions, condensation reactions, heterocyclization and many others. Triflamides act as a source of nitrogen in C-amination (sulfonamidation) reactions, the products of which are useful building blocks in organic synthesis, catalysts and ligands in metal complex catalysis, and have found applications in medicine. The addition reactions of triflamide in the presence of oxidizing agents to alkenes and dienes are considered separately.

1. Introduction

Over the past 2–3 decades, N-Trifluoromethanesulfonamides (CF3SO2NHR, TfNHR) have found wide application in organic synthesis as reagents, catalysts, additives and as substituents that transform reactivity and biological activity in a wide range of substrates. The chemistry of triflamide and triflimide derivatives was the subject of several early reviews [1,2], which confirms the high interest in such compounds. Having a strong electron-withdrawing CF3SO2-group in their structure, triflamides are among the strongest NH-acids (pKa (in H2O) for TfNH2 is 6.33, pKa (in H2O) for Tf2NH is 2.8 [1]). This property determines the use of triflamides in organic synthesis, in the production of biologically and pharmacologically active substances and in various industries. One of the most important properties of triflamide derivatives in terms of biological activity is their lipophilicity. Triflamides are widely used in the production of lithium–sulfur batteries, where N,N-dialkyl-substituted triflamide (dimethyl- or dipropyl-) is present as a solvent in the electrolyte [3]. In organic synthesis, catalysts containing a trifluoromethanesulfonamide or α-imide moiety are used (Michael [4], Friedel–Crafts [5], Diels–Alder [6], Mannich [7] and many other reactions).
This review is devoted to the latest progress in the field of applications of triflimide in organic synthesis as an active catalyst. The review also includes the reactions of triflimide and its derivatives (metal salts). As a catalyst/co-catalyst based on the Tf2N-salts of metals (Au, Ag, Fe, Li, Ca), triflimide has found application in a wide range of addition reactions, cycloaddition, intramolecular cyclization, CH-amidation, etc. Tf2NH is often used as an additive to the reaction medium, for example, in the synthesis of spiroheteropolycyclic compounds [8,9] and nitrogen-containing heterocycles [10], in catalytic (3 + 2)-annelation [11], in the oxidative synthesis of hydrodibenzofurans [12] and in condensed 2,8-O,O- or O,N-bicyclo[3.3.1]nonanes [13]. The triflimide anion is a counterion for the production of low-melting ionic liquids, used to stabilize nanoparticles, which are used in various fields, including medicine, sensors, optics and the aerospace industry [14,15]. On the basis of triflimide, various types of extractants and ionic liquids with organic cations have been obtained, which are used to isolate lanthanides and actinides from liquid waste; for example, from spent nuclear fuel [16,17].
The triflamide moiety is introduced in organic molecules in two ways. The first way is the reaction of a substrate with the activated TfNHR sulfonamide molecule. The second method is the treatment with trifluoromethanesulfonic acid anhydride or halides of the corresponding N-nucleophiles, which, as a rule, require low temperatures and the presence of additional bases.

2. Triflimide as a Catalyst in Organic Synthesis

Triflimide 1 (Tf2NH) is used as a catalytic additive in the formation of C-C and C-heteroatom bonds, due to its strong acidity, as well as its good compatibility with various organic solvents [18]. It has been widely used as a Brønsted acid for the catalysis of Friedel–Crafts reactions [19,20,21,22,23,24] and cycloaddition reactions [25,26]. In addition, triflimide is used as a Brønsted acid, for example, to obtain various bis-arylated amides from vinyl azides in moderate to quantitative yields [27] or in the hydroalkylation of arylalkenes with activated alcohols [28]. In a series of works, Ye et al. proposed different approaches to the cyclization of ynamides initiated by Tf2NH, which were applied to the synthesis of functionalized heterocyclic compounds [29,30,31,32,33].
For example, the cascade cyclization of ynamides 2 having an allyl ether moiety in the molecule made it possible to obtain various highly functionalized 3-isochromanones 3 via intramolecular alkoxylation in good to quantitative yields (68–99%) under mild conditions (Scheme 1) [30]:
Based on these works, it was proposed to activate simple alkynes and to synthesize a number of highly functionalized isoindolinones [34]. The intramolecular Tf2NH-initiated cascade cyclization of N-methoxybenzamides 4 gives rise to the formation of fused chromanes or isoindolinones 5 in good yields and with a high regioselectivity (Scheme 2). The reaction meets the requirements of green chemistry by providing an atom-economical alternative to transition metal catalysis.
Frontier et al. [35] described a simple intermolecular reaction of the carboamination of alkynols 6 with cyclic aminals 7 in the absence of metal-based catalysts. Acidic conditions (Tf2NH) allowed it to stereoselectively synthesize ether-condensed cyclic enamines 8.
The proposed reaction mechanism involves the formation of aminals by combining alkynols with aminals via the capture of iminium ions. Aminals give oxonium ions and vinyl cations, respectively, through the ring opening and subsequent reaction with alkyne. Finally, the intramolecular amination leads to the formation of cyclic enamines. Various bicyclic heterocycles were obtained in a good yield and with good diastereoselectivity (Scheme 3):
The one-pot assembling of the 9,10-dihydroacridine scaffolds 11 used in OLED devices was achieved by the selective ortho-C-alkenylation of diarylamines 10 with arylalkynes 9 followed by the intramolecular hydroarylation of the olefin formed as an intermediate. The reaction was carried out in hexafluoroisopropanol with triflimide as a catalyst which launched the reaction (Scheme 4) [36]:
Arylalkynes with electron-donating groups in the para-position give 9,10-substituted dihydroacridines in a 73 to 98% yield. Noteworthy, in the case of the nitrile and nitro groups, a complete conversion of the starting alkyne is achieved only at high temperatures of the reaction mixture. The effect of the diarylamine structure on the course of the reaction was also investigated, and it was shown that the presence of electron-donating or electron-withdrawing groups only slightly affected the reaction, as the target products were obtained in 57–95% yields.
An efficient method was developed for the nucleophilic introduction of a difluorinated carbanion (from 13) into α,β-enones 12 with the formation of 2,2-difluoro-1,5-diketones 14, the regioselectivity of the reaction being determined by the TMSNTf2 (silylbistriflimide) or Tf2NH 1 catalyst. It is the strong electron-withdrawing properties and bulky substituents in the TMSNTf2 generated in situ that were responsible for the 1,4-addition route. 2,4,6-Triaryl-substituted 3-fluoropyridines 15 can be obtained by the one-pot method (Scheme 5) [37]:
However, with other bulky catalysts, such as trimethylsilyl trifluoromethylsulfonate (TMSOTf) or tert-butyldimethylsilyl trifluoromethanesulfonate (TBSOTf), which are also effective in this reaction, 1,2-addition by-products are also formed as minor admixtures. The electronic effects of the substituents in α,β-enone do not affect the yield of the products. The yield is slightly reduced in the presence of bulky substituents.
Triflimide is used in some cyclization in Diels–Alder reactions of 4-oxopent-2-enoates [38], in Michael addition reactions to substitute 3-vinyl-1,2,4-triazines and the subsequent cyclization to tetrahydro-[1,6]-naphthyridines [39] and in the synthesis of 2,3-dihydro-1H-benzo[e]indoles and 2,3-dihydrobenzofurans using Tf2NH. The yield of the latter was 64% (with Tf2NH) but while using In(OTf)3, the yield increased to 80% [40]. The synthesis of mono- and bis-γ-lactams by the Mannich method (the addition of 2,5-bis(trimethylsilyloxy)furan to imines) occurs in good yields in the presence triflimide [41]. The intramolecular hydroacyloxylation of non-activated alkenes proceeds well in hexafluoroisopropanol not only with triflimide, but also in the presence of Ca(NTf2)2+nBu4NPF6 [42]. The aromatization of dibenzonorcaradienes to dibenzo[f,h]isocoumarins proceeds with TfOH in yields up to 99%. Replacing the triflic acid with triflimide also showed excellent results, affording the products in a 95–98% yield [43]. Enantioselective (4 + 2)-cycloaddition in the presence of triflimide gives chiral 1,2-amino alcohols, 1,2-diamines and β-amino acids in yields of 92–98% [44].
The Tf2NH-catalyzed (3,3)-sulfonium rearrangement of vinyl sulfoxides 17 or 18 in the presence of ynamides 16 is an enantioselective and stereodivergent strategy for the synthesis of acyclic polysubstituted 1,4-dicarbonyls 19 or 20 (Scheme 6) [45]:
The catalytic aldol reaction of silyl enol ethers is a universal method for C-C bond formation. Gati et al. [46] proposed the triflimide catalyzed syn-stereoselective aldol reaction for the synthesis of α,β-dioxyaldehydes 23 and 1,2,3-triols 24 from (Z)-tris(trimethylsilyl)silyl enol ethers 21 and aldehydes 22 (Scheme 7):
Iodobenzene acts as a co-catalyst that stabilizes the silylenium cation formed in situ, because the additive seemed to be playing a critical role in affecting the rate of the reaction.
The Mukaiyama aldol reaction in the presence of “supersilyl” tris(trimethylsilyl)silyl enol ethers 26 with ketone 25 leads diastereoselectively to α,β-dioxyaldehydes, whereas the same reaction in the presence of Tf2NH 1 (the reaction of silylketene acetals with ketones) leads to siloxycarbonyl compound 28 [47] (Scheme 8):
This efficient methodology allows for a rapid and stereoselective construction of mono-, bis- and tris-hydroxyaldehydes by mono-, double- and triple-cross-aldol processes, respectively, to produce polyketide-like scaffolds that are particularly useful for the construction of complex natural polyketides [47].
The Peterson olefination in the presence of triflimide 1 [48], as well as the synthesis of α-CF3 and α-CF2H amines 30 by the aminofluorination of gem-difluoroalkenes and monofluoroalkenes 29, respectively, was described (Scheme 9) [49]. Selectfluor was used as a source of electrophilic fluorine and acetonitrile as a source of nitrogen.
Triflimide 1 is used as a catalyst in the isoprenylation of aliphatic aldehydes via the (3,3)-sigmatropic rearrangement of N-Boc-N-(1,1-dimethylallyl)hydrazones [50], in the synthesis of amides from vinyl azides and alcohols [51] and in the rearrangement of N-(1-trimethylsilyl)allylhydrazones with the formation of the corresponding vinylsilanes and cyclopropanes [52]. In addition, triflimide activates the silylium catalyst in the reactions of the selective functionalization of azines with the formation of N-silylated dihydropyridines [53]. A number of works can be mentioned in which Tf2NH was successfully used as a catalyst. These include the synthesis of poly-l-lactide in CO2 under plasticization conditions [54], the nucleophilic C-arylation of halopurines leading to N7-substituted purine biaryls [55], reactions of the diastereoselective intramolecular hydride shift in the presence of alkenes [56], the preparation of amides from vinyl azides and alcohols [50], the synthesis of polysubstituted naphthalenes by the reaction of arylacetaldehydes with alkynes (benzenylation reaction) [57,58], glycosylation reactions [59], three-component regioselective synthesis of tetrahydrofuro[2,3-d]oxazoles [60], etc.

3. Triflimide Derivatives in Organic Synthesis

3.1. AgNTf2

The direct C–H amidation of substituted benzamides 31 with trichloroethoxycarbonyl azide (TrocN3) 32 was reported in the presence of AgNTf2 33 [61]. When cesium acetate is added, the reactions proceed efficiently and with high regioselectivity, affording various functionalized quinazoline-2,4(1H,3H)-diones 34, which are important building blocks and key synthetic intermediates from a biological and medical point of view. During the reactions, two new C-N bonds are formed by the successive rupture of the C-H and N-H bonds (Scheme 10)
N-Alkylated benzamides 31 with electron-withdrawing or -donating groups in the para-position of the phenyl ring cyclize to the corresponding products in good yields. In the same way proceeds the regioselective C-H amidation of meta-substituted benzamides 31. However, substituents in the meta position affect both the steric hindrances and electron density of the reaction center of the substrate [61]. The method can be successfully applied to a wide range of oximes, aryl- and alkylsulfonamides as a source of nitrogen [62]. Catalysis by chiral cyclopentadienyl complexes of iridium in combination with AgNTf2 33 is used for CH-arylation of tetralone derivatives with arylboronic esters [63].
The diastereoselective synthesis of spirocyclic pyrrole-2-one-dienone systems was conducted in the presence of AgNTf2 33 in combination with a gold-based complex [64]. A similar procedure was used for the synthesis of diarylmethine-substituted enones [65]. The combination of AgNTf2 33 and the AuCl-based ligand was also successfully used in the cycloisomerization of 1,6-enyne to 3-azabicyclo[4.1.0]heptene [66], as well as in the asymmetric synthesis of fused bicyclic N,O- and O,O-acetals [67]. The (C6F5)3PAuCl/AgNTf2 catalyzed cyclization of N-tosyl-protected 5-benzyl-6-((trimethylsilyl)ethynyl)-1,2,3,4-tetrahydropyridines obtained by the Sonogashira reaction in the presence of enol triflate lactam derivatives gives tetrahydrobenzo[g]quinolines, the skeletal framework of which is a recurring motif in natural products [68].
In 2019, an efficient and stereoselective method was developed for the synthesis of (Z)-β-halogenated enamides by the Ritter type Au-catalyzed reaction in the presence of BrettPhosAuCl and AgNTf2 33 using haloalkynes as substrates [69]. The regioselectivity of the reaction was controlled by bulky substituents in the substrates. Other combinations of AgNTf2 with Au+-based complexes also showed the efficiency in various processes, for example, in the [4 + 3]-annelation of anthranils with 1,5-enynes to form tetrahydrobenzoazepine derivatives [70,71]. The chemoselectivity of the reaction depended on the type of the used alkynes. Silver triflimide AgNTf2 33 was used in one of the stages of the synthesis of new fluorinated symmetric and asymmetric imidazolium salts, as well as of their complexes with various metals [72] and in the diastereoselective synthesis of benzo[5,6]oxepino[2,3-c]pyrroles via the [5 + 2]-annulation of the donor–acceptor type of arylvinyldiazosuccinimide with ketones [73]. The iodination reaction of a wide range of arenes (derivatives of anisole, aniline, acetanilide and phenol) in the presence of AgNTf2 33 and N-iodosuccinimide is known [74].

3.2. Gold Triflimide

Gold triflimides 35 are widely used in organic synthesis. Gold α-iminocarbene complexes have demonstrated good catalytic efficiency in the synthesis of N-heterocycles in the last few years [75]. [1,3]Dioxino[5,4-b]indoles 38 have been synthesized by the [4 + 2]-cycloaddition of the 3-indolylidene-Au-carbenium intermediate 39 to aldehydes 37 (Scheme 11). The reaction presumably occurs due to the presence of a hydroxyl group in the starting 3-(2-azidophenyl)prop-2-yn-1-ols 36. On the other hand, the presence of electron-withdrawing groups such as 3,5-dichloro or 3-cyano in the phenyl ring facilitates the reaction by excluding the competitive attack by the gold-activated alkyne:
A number of syntheses of the oxazino- [76] and pyrazinoindoles [77], pyrroloindoles [78], carbazoles condensed with benzofuran, 1,3H-dibenzo[a,h]carbazoles [79], pyridine derivatives [80], pyrroles [81,82,83,84,85], azepines [86] and other polycyclic compounds [87,88,89,90,91,92,93] are known in the literature. In contrast to intramolecular reactions, less is known about the intermolecular formation of gold carbene α-imino intermediates. Nevertheless, these reactions were applied to the synthesis of α,β-unsaturated amidines [94] or 2-aminopyrroles [95] using highly polarized alkynes, for example, ynamides.
In the presence of Ph3PAuNTf2 35, the regioselective functionalization of arenes and heteroarenes (derivatives of benzenes, phenols, ethers, indoles, pyrroles, furans and thiophenes) 43 with acetylenes 41 was performed using N-alkenoxypyridinium salt 42 as electrophilic alkylating agents for the synthesis of α-aryl- and heteroaryl ketones 44 (Scheme 12) [96]:
N−O bond cleavage might occur prior to the attack by the arene. After N−O bond cleavage, an electrophilic carbocation species is likely produced [96].
Using Ph3PAuNTf2 35 makes it possible to synthesize the derivatives of 2-substituted 4-oxo-4-arylbutanal from alkynes and vinyl sulfoxide; five-, six- and seven-membered cycloalkyl-1-ones, for example, the derivatives of tetrahydrocycloalkyl[b]pyrrole, which are pharmaceutical structural units, were easily obtained from 1-cycloalkenyl sulfoxides [97]. The use of Ph3PAuNTf2 35 allowed it to synthesize cyclohept[b]pyrroles from diynes and pyrroles [98] and to perform glycosylation reactions [99,100]. The combination of Ph3PAuNTf2 35 with N-iodosuccinimide made it possible to develop a highly efficient synthesis of arenes labeled with radioactive iodine. The reaction represents the first example of the use of homogeneous catalysis in selective synthesis using radioactive materials. The method was used to obtain meta-[125I]iodobenzylguanidine, a radiopharmaceutical used for imaging and tumor therapy [101].
The synthesis of unsymmetrical esters using benzyl, t-butyl alcohols as alkylating reagents, also catalyzed by Ph3PAuNTf2 35, was described [102]. An interesting type of reactions proceeding in the presence of gold triflimides complexes are the so-called NT-reactions (Nitrene Transfer Reaction). Among various NTRs, in a vast majority of studies, the azides, azirines, isoxazoles, anthranils, pyridinium azaylides and sulfilimines were used as the reagents. These reactions exhibit different activity and selectivity, depending on the chemical nature of the reactants. However, some of the NTRs have significant disadvantages. For example, azides are potentially explosive, and the ylides are poorly reactive in the gold-catalyzed reactions and are not easily available. Shcherbakov et al. [103] have recently demonstrated the gold catalyzed nitrene transfer reaction from benzofuroxans 45 to N-allylynamides 46, resulting in the formation of various 3-azabicyclo[3.1.0]hexanes 47 (Scheme 13):
This highly selective annulation occurs under mild conditions (5 mol % Ph3PAuNTf2, PhCl, 60 °C) and is applicable to substrates with various functional groups (21 examples, yield ≤ 96%) [103].

3.3. LiNTf2, Ca(NTf2)2 and Fe(NTf2)3

In the presence of LiNTf2 48, various oxetanes are efficiently opened by C-nucleophiles; for example, silylketene acetals under these conditions give a spectrum of saturated 1,5-oxygen-containing molecules that are part of many natural compounds (for example, polyketides) [104]. It should be noted that when LiNTf2 is replaced by TBSNTf2 (TBS = = tert-butyldimethylsilyl) as a catalyst, the reaction leads to the formation of 1,3,7-oxygen-containing products [104].
The formation of oxetane ethers 5152 under the conditions of the Friedel–Crafts reaction between oxetanols 49 and phenols 50 in the presence of LiNTf2 was studied in detail by Bull et al. [105]. The introduction of para-substituents in the phenyl ring shows a direct dependence of the reaction course and the structure of the products on the nature of the nucleophile, namely, the formation of the kinetic products of O-alkylation versus the thermodynamic products of C-alkylation 53 (Scheme 14) [105]:
For electron-deficient 4-cyanophenol, the only observed product was the product of O-alkylation, while for X = halogen and electron-releasing substituents, after 1 h, a mixture of C- and O-alkylated products was formed, and after 20 h, only the C-alkylation products were observed in the mixture. In addition, a number of oxetane ester derivatives were isolated, which are new potential bioisosteres for esters of carboxylic acids [105]. When the reaction was monitored using 1H NMR spectroscopy without a nucleophile present, a rapid degradation of the starting material occurred. A small amount of an aldehyde side product was identified; however, the majority of the material was unaccounted for. An insoluble precipitate was observed, indicating the possible formation of a polymeric species under the reaction conditions resulting from ring opening of the oxetane, promoted by the formation of the oxetane carbocation in the absence of a better nucleophile than the substrate itself [105].
The chemo-, regio- and stereoselective addition of triflimide (from LiNTf2 48) to alkynes 54 is known. The reaction represents the second sample of the rare class of vinyl triflimides 55 in good yields (Scheme 15) [106]:
N-vinyl triflimides were first described in the work [107].
The reaction was performed by adding phenylacetylene to the dichloromethane solution of 1.5 eq. LiNTf2 48 and Bu4NNTf2 to increase the solubility of LiNTf2 48. The regioselectivity of the addition is determined by the cation-stabilizing effect of the α-aryl substituent [108]: (Scheme 16).
A similar reaction of Si-substituted alkynes 59 and Tf2NH 1 with the formation of vinyl triflimides 60 is known. The process proceeds in dichloromethane with slight heating and does not require the presence of additional catalysts (Scheme 17) [109]:
In the presence of calcium triflimide 66, the annulation reaction of an aldehyde 61, amine 62, alkene 63 and alkyne 64 takes place. The process proceeds with the formation of fused pyrrolo[1,2a]quinolones 65 with exclusive syn-diastereoselectivity and good yields (Scheme 18):
This tandem annulation process delivers products that generally comprise four or five rings fused in an angular fashion [110].
For example, in the reactions of furylcarbinol with amines under the action of Ca(NTf2)2, the Aza-Piancatelli rearrangement occurs, and after the addition of Et3N to the reaction mixture, an intramolecular Michael addition proceeds with the formation of the corresponding tetrahydrobenzo[b]azepines in 33–93% yields [111]. The reactions of N,O-acetals with vinylboronic acids in the presence of Ca(NTf2)2 lead to the formation of isoindolinones in good to quantitative yields [112].
Iron(III) triflimide salt 67 (formed in situ from FeCl3 and AgNTf2) is an efficient catalyst for addition reactions [113]. An alternative method for preparing Fe(NTf2)3 that avoids the use of additional metal salts consists of dissolving FeCl3 in a readily available and inexpensive ionic liquid [BMIM]-NTf2 (1-butyl-3-methylimidazolium bis(trifluoromethanesulfonyl)imide). The combination of FeCl3 with the ionic liquid resistant to air oxygen accelerates the reaction, and the FeCl3+[BMIM]NTf2 system has proven to be ideal for the direct iodination of aromatic compounds 67 in the presence of N-iodosuccinimide (NIS) with good yields (58–91%) [114] (Scheme 19):
The reaction with N-chlorosuccinimide (NCS) proceeds similarly [100,115]. The yields of the corresponding chlorinated arenes (26 examples) ranged from 53 to 97%. The reaction was used for the mono- and di-chlorination of a number of target products, such as nitrofungin, the anti-bacterial agent chloroxylenol and the herbicide chloroxynil [115].
In the presence of Fe(III) and Cu(I), a regioselective reaction of the para-amination of the activated arenes 69 occurs via bromination with N-bromosuccinimide (NBS) (yields 51–95%) [116] (Scheme 20):
The reaction involves the bromination of an aryl substrate in the presence of Fe(NTf2)3 followed by an N-arylation reaction catalyzed by Cu(I). A similar system was used for the synthesis of 2-arylbenzoxazoles and 2-arylbenzothiazoles from N-arylbenzamides. [117].

4. Synthesis of Biologically Active Triflamide Derivatives

Thrombosis is the main pathogenesis that causes the low curability of ischemic stroke and is often the cause of death and disability worldwide. Metformin 72, a biguanidine derivative, is a drug for the treatment of type 2 diabetes mellitus, which alleviates the course of ischemic stroke in patients with diabetes mellitus. Based on triflamide and metformin, a promising drug for the treatment of stroke was obtained [118] (Scheme 21):
The compound is prepared by the simple treatment of metformin 72 with trifluoromethanesulfonyl chloride 71 in anhydrous acetone in the presence of catalytic amounts of KOH. The target product was obtained in an 85% yield. The compound inhibits the formation of human platelets, including the reduction of platelet aggregation, adhesion and clot retraction, strongly inhibits the formation of blood clots in arteries, reduces the size and compactness of blood clots in stroke, reduces damage to nerve function and mortality and does not cause severe toxicity and tissue damage [118]. As shown later [119], this modified metformin exhibits selective biological activity against breast cancer cells (MFC-7). Together with its anti-thrombotic properties, which is very important in the treatment of cancer, the compound is a promising drug for the therapy of cancer [119,120].
N-Trifluoromethanesulfonyl-substituted anilines 75 proved to be effective substrates for biocatalytic hydroxylation with the formation of 4-aminophenols in the presence of a number of cytochromes of the P450BM3 family. The reactions proceeded with 100% conversion. Similar results were shown by N-trifluoroacetyl protection at the nitrogen atom of aniline. Ac- or Boc-derivatives of anilines showed only a 17 and 66% conversion, respectively [121] (Scheme 22):
P450-catalysed arene hydroxylation is accepted to occur via an NIH-shift within an iminium intermediate to give a dienone that re-aromatizes to phenol [121]. Biocatalytic hydroxylation reactions are used for the synthesis of drugs, agrochemicals and their metabolites. In the case of N-trifluoromethanesulfonyl derivatives, the reaction proved to be effective both in preparative and screening variants [121].
The reaction of 5-methyl-2-phenyl-4,5-dihydrooxazole 76 with trifluoromethanesulfonic acid anhydride 77 leads to the formation of trifluoromethanesulfonamido)prop-2-yl esters of phenyl-substituted derivatives of benzoic acid 78 [122] (Scheme 23):
The product yield varies from 56 to 88%. The reaction does not require the presence of bases. Derivatives with R = 4-Cl, 4-NO2 and 4-CF3 show high cytotoxicity against six human cancer cell lines (U251 (glioblastoma), PC-3 (prostate adenocarcinoma), K-562 (chronic myelogenous leukemia), HCT-15 (colorectal adenocarcinoma), MCF-7 (breast adenocarcinoma), SKLU-1 (lung adenocarcinoma) and the 4-CF3 derivative was active against human gum cancer cells (FGH  (gingival fibroblastoma) [122].
A triflamide derivative of diphenylpyrimidine 82 is known, which exhibits excellent activity against the proliferation of pancreatic carcinoma cells (AsPC-1, Panc-1, BxPC-3), lymphoblastic leukemia cells (Ramos) and some lung cancer cells (Scheme 24):
4-Aminephenol was converted to N-phenyl triflamide 80 intermediates under the action of trifluoromethane sulfonic anhydride. Then, 80 were reacted with 4-fluoronitrobenzene to form the intermediate, which were conveniently converted to the amine derivative 81 by using the Fe-NH4Cl reduction condition. Additionally, under the action of the p-toluenesulfonic acid reagent, 81 was reacted with the 2-chlorine pyrimidine derivative to generate the title molecule 82 [123].
On the basis of triflamide 83, a derivative of oseltamivir 84, an anti-viral drug used to treat various types of influenza, was obtained (Scheme 25):
Coupling reactions of acid 84 with triflamide 83 were carried out, followed by the removal of the Boc protecting group with TFA, to afford acyl triflamide 85. The amino group in 85 was further elaborated to the guanidino group by treatment with 1,3-di-Boc-2-(trifluoromethylsulfonyl)guanidine to afford GOC-sulfonamides 86 after the removal of the Boc groups. The compound 86 exhibited high inhibitory activity against the H1N1 influenza virus, with the final yield of the product being 39% [123,124].
Type 2 diabetes mellitus is a progressive metabolic disorder characterized by high blood glucose and high endogenous insulin levels. Diabetes mellitus causes serious vascular complications, heart disease, kidney failure and blindness [125]. Glucagon-like peptide-1 (GLP-1) is a potent anti-hyperglycemic hormone that induces the glucose-dependent stimulation of insulin secretion while simultaneously inhibiting glucagon secretion. However, active GLP-1 is rapidly degraded by the dipeptidyl peptidase-4 (DPP-4) enzyme. Therefore, the inhibition of DPP-4 is a new approach to the treatment of type 2 diabetes. Based on triflamide, a DPP-4 inhibitor 90 in the low micromolar range was obtained, the pharmacokinetic profile of which was suitable for clinical use (Scheme 26):
Compound 90 was synthesized according to Scheme 26. The condensation of 87 with 88 provided compound 89, which was then reduced by H2 and condensed with sulfonyl chloride and was followed by the deprotection of Boc by treatment with TFA in CH2Cl2, which provided target compound 90 [125].
Triflamide derivatives have been obtained that exhibit anti-diabetic activity as inhibitors of aldose reductase (ALR2), aldehyde reductase (ALR1) and antioxidant activity. These N-acyl triflamide derivatives 94 are obtained on the basis of the quinoxalinone 91 framework [126] (Scheme 27):
3-Chloro-quinoxalin-2(1H)-one 91 was alkylated with methyl bromoacetate to form methyl ester 92 as a key intermediate. Then, 92 was subjected to the Heck coupling reaction with the corresponding styrenes and then to hydrolysis with lithium hydroxide, and the carboxylic acid 93 was afforded. Finally, 93 was treated with triflamide 83 in the presence of EDC·HCl and DMAP to obtain the desired compound 94.
The yield of the final reaction products was 57–70%. The product 94 with R = 3,4-(OH)2 exhibited the highest inhibitory activity. It also exhibited extraordinary antioxidant activity, which was even higher than that of the commercial drug Trolox (6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid), a water-soluble analogue of vitamin E [126].
The triflamide derivative of indomethacin 97 is known, which is promising in the treatment of prostate cancer, showing high selectivity and good inhibitory properties in the treatment of this disease. The yield of the product was 70% [127] (Scheme 28):
Compound 96 was obtained from 4-chloro-N-(4-methoxyphenyl)- benzohydrazide hydrochloride (obtained from 95) by the refluxing with slight excess of 4-oxobutanoic acid in AcOH, respectively. Then, 96 and triflamide 83 were dissolved in 2 mL of 1,2-dichloro ethane (DCE) under stirring to obtain compound 97.
Triflamide has been used to synthesize the selective inhibitor of matrix metalloproteinase-12 (MMP-12) 100. The inhibition of the reactions of this enzyme often plays an important role in the treatment of lung, inflammatory and cardiovascular diseases [128] (Scheme 29):
Diarylsulfide derivative 100 was prepared as reported in Scheme 29. Compound 98 was protected as an ethyl ester by treatment with thionyl chloride (SOCl2) in ethanol. This step was necessary to improve the subsequent cross-coupling reaction, which exhibited a poor yield if conducted on carboxylic acid. A palladium-catalyzed cross-coupling reaction (under Suzuki conditions) of protected aryl bromide 98 with 4-methoxyphenylboronic acid afforded biphenyl derivative 99. Compound 99 was converted into sulfonamide 100 by condensation with triflamide 83, respectively, in the presence of N-(3-(dimethylamino)propyl)-N′-ethylcarbodiimide (EDC) and 4- dimethylaminopyridine (DMAP) using dichloromethane as a solvent.
The yield of the target product 100 was 80%. A similar sulfone was also obtained, but its selectivity and inhibitory activity were much lower [128].
Chiral complex 103 was obtained on the basis of N-trifluoromethanesulfonyl substituted amino acids 102 and Au(I). The complex 103 exhibited in vitro cytotoxicity against breast cancer cells with limited toxicity to healthy epithelial cells [129] (Scheme 30):
Ligand 102, derived from the polar amino acid proline’ Secondary amine 101, was alkylated with methyl 3-bromopropanoate16 and the ester was subsequently saponified to the corresponding carboxylic acid. An EDCmediated amidification reaction afforded the corresponding triflic propionamide 102. Complex 103 was derived from the dimethylsulfide gold chloride complex; the two different structures depend on the initial presence of a silver(I) salt. The unusual dimeric complex 6 was obtained in one synthetic operation by ligand metathesis between a silver amide salt generated in situ from ligand 102. The complex 103 was obtained in a quantitative yield. These complexes of triflamide and gold(I) derivatives have the ability to selectively accumulate in adenocarcinoma cells, which makes it possible to effectively treat cancer in elderly patients and for slow-growing tumors, that is, just in those cases when traditional cancer therapy aimed at combating rapidly dividing cells cannot be used [129].
Triflamide derivatives are known, on the basis of which non-steroidal anti-inflammatory drugs were obtained, similar in action to nimesulide or celecoxib, which are safe for the gastrointestinal tract and cardiovascular system. The resulting compounds 109 are characterized by a high anti-inflammatory activity, similar to or higher than that of nimesulide or celecoxib, as well as a high COX-2/COX-1 selectivity [130] (Scheme 31):
The synthetic pathway for compound 109 can be summarized as follows: 3-bromopyridine 104 is first oxidized by a mixture of acetic acid and hydrogen peroxide to afford 3-bromopyridine N-oxide 105, which is then nitrated at the 4- position by a nitric and sulfuric acid medium to provide the key intermediate 3-bromo-4-nitropyridine N-oxide 106. The NH-bridge is achieved by the reaction of intermediate 106 with properly substituted anilines 107. The 4-aminopyridine intermediates 108 are obtained after the simultaneous reduction of the nitro and the pyridine N-oxide moieties using iron in an acetic acid and water medium. The pyridine analogs of nimesulide 109 are obtained by the reaction of the aminopyridine intermediates 108 with trifluoromethylsulfonyl chloride in acetonitrile in the presence of potassium carbonate. The products 109 were obtained in a 35–77% yield [130].
Triflamide derivatives exhibit biological activity as progesterone receptor antagonists. The product 113 presented in Scheme 32 is a potentially effective drug in the treatment of diseases of the female reproductive system, including endometriosis [131] (Scheme 32):
The synthetic route towards compound 113 is described in Scheme 32. The deprotonation of cyclohexanol 111 with NaH in THF followed by the addition of fluorobenzonitrile 110 furnished cyclohexylamine 112 in a good yield. Compound 113 was synthesized by the triflation of cyclohexylamine 112 using triflic anhydride and triethylamine in DCM at −60 °C. The yield of the final product 113 was 92%. The product 113 also inhibited drug metabolism by cytochromes P450, CYP 2C [131].
N-Trifluoromethanesulfonyl-substituted derivatives 117 were obtained, exhibiting anti-mycobacterial activity, and could potentially become new anti-microbial drugs, including for the treatment of tuberculosis [132] (Scheme 33):
The synthetic route towards compound 117 is presented in Scheme 33. Acetyl chloride was added to a solution of 114 in methanol under nitrogen. The reaction was stirred for 3 h at room temperature. Sodium hydride and benzyl bromide were added. The reaction was cooled in an ice bath, quenched by the addition of methanol and then concentrated in vacuo. Then, the residue was dissolved in a mixture of water and acetic acid to afford hemiacetal 115. Compound 115 and triflamide 83 were stirred at room temperature in dry diethyl ether in the presence TMSOTf to give compound 116. Compound 117 was obtained by reduction with 10% activated Pd/C. The N-furanosyl triflamide 117 was obtained in a 44% yield [132]. Compound 117 was configurationally stable in aqueous solutions, in contrast to its other analogues obtained from alkylsulfonamides. N-Acyltriflamides also exhibited anti-mycobacterial activity [133].
Triflamide derivatives exhibit biological activity against the causative agent of sleeping sickness (African trypanosomiasis) [134]. Currently, there are five drugs for the treatment of this disease, including suramin, pentamidine, melarsoprol, eflornithine and nifurtimox [135]. However, their use is accompanied by a number of serious side effects and difficulties: (1) high toxicity; (2) the necessity to be injected intramuscularly or intravenously, which creates difficulties in an epidemic area with limited medical resources; (3) a narrow antitrypanosomous spectrum of action; and (4) high cost. In general, these drugs are not effective in treating the disease, and there is an urgent need to develop more effective and inexpensive chemotherapeutic agents for the treatment of trypanosomiasis. The treatment of this disease involves blocking the processes of polymerization/depolymerization of the tubulin protein, which is necessary for the division of pathogen cells and their movement. On the basis of the triflamide derivative, a selective tubulin inhibitor 121 was obtained, which can be used to treat sleeping sickness [134] (Scheme 34):
Triflamide derivative 119 was prepared from aryl-substituted 2-amino-5-nitrophenol 118 by adding it with trifluoromethanesulfonic chloride in anhydrous DCM and K2CO3. Compound 119 was dissolved in acetone; then, Zn and FeCl3 were added into the solution. When the reaction completed, the corresponding benzoyl chloride was added and product 121 was collected by filtration and purified by recrystallization in ethanol/water. The yield of the final product 121 of the reaction was 47% [134].
The derivative of triflamide 126 is known to exhibit biological activity and act as an inhibitor of the transporter derivatives of uric acid (hURAT1). In medicine, there are only three options for drugs of this kind, although they are extremely necessary in the treatment of hyperuricemia, which subsequently causes many diseases, such as gout, arterial hypertension, chronic kidney disease and some cardiovascular diseases [136] (Scheme 35):
Triflamide compound 126 was synthesized by following Scheme 35. Bis(pinacolato)diboron reacted with 122 under the catalyst of Pd(dppf)Cl2, which gave compound 123. Further Suzuki coupling of 123 with 4-amino-3-bromopyridine 124 provided 125, which subsequently reacted with CF3SO2Cl under the conditions of Et3N in DCM to afford compound 126 [136].
Based on triflamide, a promising inhibitor of the α-amylase enzyme 130 was obtained [137] (Scheme 36):
Sulfonohydrazide-substituted indazole 130 was synthesized by a multi-step reaction. In the first step, 4-oxyindazole 129 was formed by reacting dimedone 127, dimethylformamide dimethylacetal (DMF-DMA) and phenylhydrazine 128 in the presence of a catalytic amount of CuCl2 in ethanol for 3 h. In the next step, 4-oxyindazole 129 was treated with a sulfonylhydrazide derivative in ethanol in the presence of pyridine as a catalyst and was refluxed for 2 h to obtain the desired sulfonohydrazide-substituted indazole 130 (Scheme 36). Reducing the activity of the α-amylase enzyme is one of the best treatments for type 2 diabetes. For example, acarbose, voglibose and miglitol are commercially available α-amylase enzyme inhibitors used to treat type II diabetes mellitus. However, these agents have some side effects such as flatulence, diarrhea and abdominal discomfort, so other anti-diabetic agents are always recommended for greater effectiveness [137].
Triflamide derivatives are used for the synthesis of inhibitors of biochemical processes by introducing peptidomimetics into biochemical reactions. A triflamide-containing component of the proteasome complex 133 was obtained, exhibiting activity similar to chymotrypsin. Similar compounds are used for cancer therapy. The product 133 yield was 80% [138] (Scheme 37):
The synthesis of the target compound is presented in Scheme 37. The treatment of compound 131 with hydrazine generated compound 132, which was treated with trifluoromethylsulfonyl chloride in a basic media to generate sulfonamide 133 over two steps [138].
Scheme 38 shows a simple synthesis of a compound 137 exhibiting high anti-retroviral activity. The compound showed good activity against HIV at nanomolar concentrations, being an effective inhibitor of HIV-1 replication [139] (Scheme 38):
The synthesis of triflamide derivative 137 was achieved from the key intermediate, N-(4-amino-2-methylphenyl)-4-chloro-phthalimine 136, which was synthesized from 4-chlorophthalimine 135 and 4-chloro-3- methylaniline 134. Further, 136, when treated with appropriate CF3SO2Cl in the presence of TEA and DCM at room temperature, gave the respective sulfonamide 137.
The resulting isoindolindione 137 gave a good performance in overcoming the hematoencephalic barrier. Obtaining such drugs is important because there is a demand in medicine for drugs for highly active anti-retroviral therapy, which is necessary to curb the progression of HIV disease and increase the survival of HIV-infected patients, as well as due to the high resistance of retroviruses [139].

5. Triflamide Derivatives in Organic Synthesis

Arylbenzyltrifluoromethanesulfonamides 139 are starting compounds for the synthesis of phenanthridine derivatives 140, which are among the most important basic nitrogen-containing heteroaromatic structures. Phenanthridine fragments are part of natural compounds and have powerful anti-bacterial and anti-tumor activity [140]. O-Arylbenzyltrifluoromethanesulfonamides 139 are obtained by the reduction of the corresponding CN-diaryls 138 followed by treatment with triflic anhydride [140] (Scheme 39):
Then, the product undergoes intramolecular oxidative cyclization at the aromatic ring in the presence of DIH in 1,2-dichloroethane under irradiation with a tungsten lamp with the formation of 5-trifluoromethanesulfonyl-5,6-dihydrophenanthridine, which, by the action of t-BuOK in THF, gives the target phenanthridine 140 (79%) [140].
An olefination reaction of 2-methylquinoline in the presence of triflamide was shown to proceed in the presence of aldehydes under the action of microwave radiation with the formation of 2-vinylquinolines 144 [141] (Scheme 40):
Not only derivatives of aliphatic aldehydes or benzaldehydes 142, but also heteroaromatic aldehydes (pyridinecarboxaldehyde and thiophenecarboxaldehyde) can be involved in the reaction. Triflamide 83 is added in the excess of 20% with respect to the aldehyde. As seen in Scheme 40, the microwave irradiation of TfNH2 83 gives the corresponding aldimine in situ. The corresponding alkene 144 is formed by the elimination of the triflamide molecule from 143. Some of the obtained alkenes have high anti-malarial biological activity [141].
Various triflamide compounds are widely used as organyl catalysts that show high stereoselectivity and asymmetric transformations [142,143,144,145,146,147,148,149].
Triflamide derivatives 147 are used in the enantioselective cross-coupling of glycine 146 derivatives with ketones and aldehydes 145 [143] (Scheme 41):
The photoinduced process includes the oxidation of glycine derivatives to an imine intermediate, which enters the asymmetric Mannich reaction with an enamine 149 intermediate adduct formed in situ from a ketone or aldehyde 145 and a chiral TfNH-containing organic catalyst 147. The method allows one to create a new C–C bond with the formation of new stereocenters without the additional functionalization of the substrates [143].
Similar to organic catalysts, 2-(trifluoromethanesulfonamidoalkyl)pyrrolidines and their D-prolinamides were used in the addition of aldehydes to β-nitroalkenes at room temperature. The reaction in this case proceeded without additional reagents and catalysts and led to the formation of γ-nitroaldehydes in a quantitative yield and high enantio- and diastereoselectivity [144].
N-Propyltriflamide 151 reacts with activated alkenes 152 in the presence of an Ir(III) complex as a photocatalyst and a base (quinuclidine) [150] (Scheme 42):
The more stable nitrogen-centered radical present on triflamide acts as an intermolecular hydrogen-atom abstractor to the deprotonated triflamides in the solution, which possess highly activated α-CH bonds due to their anionic character. This pivot in reactivity allows for the selective functionalization of both α- and δ-CH bonds depending on the installed nitrogen protecting group. The presence of a trifluoromethanesulfonyl group at the nitrogen atom ensures the complete deprotonation of the NH bond and makes the α-C-H bond more “hydride” and susceptible to hydrogen-atom transfer (HAT) under the action of quinuclidine and a photocatalyst. The reaction allows for the formation of a new C-C bond due to the combination of α-amino radicals and electron-deficient alkenes. The yields of 153 vary from 32 to 63% [150].
The reaction of 2H-azirine 154 with a slight excess of Tf2O in chloroform leads to the formation of the N-acetophenone derivative of triflamide 157 [151] (Scheme 43):
2H-Azirine 154 is activated by a nucleophilic attack by the lone nitrogen pair on the highly electrophilic Tf2O, resulting in the formation of the triflate of the N-triflylazirinium cation 155. In the presence of water in the reaction medium, the iminium ion is hydrolyzed to N-triflyl α-aminoacetophenone 157 (71%). In anhydrous medium, when 2-chloropyridines are added, this 2H-azirine reacts in the presence of Tf2O to give imidazo[1,2-a]pyridines [151].
A unique one-pot reaction of the radical trifluoromethanesulfonation and trifluoromethylation of imines 158 is known, which proceeds with the formation of CF3-substituted N-vinyl triflamides 159 [152] (Scheme 44):
The reaction starts with the formation of N-vinyltriflamides, which are the sources of the CF3 group. In this case, the N-vinylatriflamides obtained at the first stage of the reaction further react as bifunctional reagents, acting as both sources of trifluoromethyl radicals and acceptors of these radicals. The yields vary from 34 to 61% [152].
Based on the triflamide derivatives, a new class of block of polystyrene-block-poly(ethylene oxide-co-hydroxyethyl glycidyl ether)diblock copolymers 163 is developed [153] (Scheme 45):
The resulting materials are currently being used to produce the next generation of single-ion-conducting polymer electrolytes (SICPEs) that are capable of competing with known Tf2NLi-based materials. In solid single-ion-conducting polymer electrolytes, the anion is covalently attached to the polymer and, ideally, only the cation contributes to the conductivity. The use of these electrolytes effectively eliminates many of the current safety and performance issues of the current liquid and salt-in-polymer electrolytes.
N-Phenethylenetriflamides 164 react with 1,3-dienes 166 in the presence of catalytic amounts of Pd(OAc)2 and Cu(OAc)2/O2 as an oxidizing agent [154] (Scheme 46):
The reaction proceeds chemo-, regio- and diastereoselectively with the formation of 2,3,4,5-tetrahydro-1H-benzo[d]azepines 166. The use of O2 as a co-oxidant makes it possible to reduce the amount of the oxidizing agent (Cu(OAc)2). 3-Benzazepines are present in a wide variety of natural products and important pharmaceuticals. These compounds are among the most convenient in terms of structural proximity and selectivity for dopamine D1 receptors, which regulate cell growth and development. As drugs for the treatment of CNS diseases, dopaminergic 3-benzazepines have the selective agonist or antagonist properties of dopamine D1 receptors, which have led to the development of pharmaceutical drugs against Parkinson’s disease, leukemia, cocaine addiction and obesity [154]. Triflamide derivatives 166 containing a 3-benzazepine moiety can be used to treat Alzheimer’s disease [155]. It should be noted that the synthesis of the N-alkyl-substituted triflamides presented in the scheme above is carried out by the reaction of TfNH2 with the corresponding alcohol in the presence of Cs2CO3 and an Ir-centered catalyst [156]. Such substituted sulfonamides exhibit a number of different types of biological activity (e.g., anti-inflammatory activity) [156].
With the participation of triflamide 83, the rhodium(III)-catalyzed intermolecular reaction of the C–H amination of ketoxime 167 in the presence of iodobenzene diacetate takes place [157] (Scheme 47):
In this case, under the conditions of an oxidative reaction, triflamide acts as a source of nitrogen during amination. The product 168 is formed in a 51% yield [157].
In the presence of NaOCl, the C–H amination reaction of 3-substituted indoles 169 with chiral triflamide-containing amino acids 170 proceeds. A feature of the reaction is the formation of a new class of atropisomers 171 [158] (Scheme 48):
The yield of the product 171 is 50–92%. Indoles are among the most important heteroarenes that are present in natural products and pharmaceuticals. The compounds are of great interest in the field of synthetic and medicinal chemistry [158].
A unique property of triflamide derivatives is the ability to aminate non-activated C-H bonds, which is a great advantage in the synthesis of N-containing molecules due to the high efficiency and atom economy of such methods. A very unusual reaction of selective intramolecular amination at the C–H bond in the presence of silver salts resulting in functionalized heterocyclic products was published [159] (Scheme 49):
The method is effective for the transformation of non-functionalized groups in organic molecules and for the creation of complex structural units in natural and bioactive molecules; the reaction is most effective in the amination of primary sp3 carbon atoms as the reactions proceed chemo- and regioselectively in comparison with the existing methods of direct amination (the Hoffmann–Löffler–Freitag reaction and the nitrene insertion reaction) [159].
Over the past few years, the reactions of triflamide with alkenes and dienes in the presence of oxidizing agents have been thoroughly studied in our group. Triflamide 83 reacts with styrene derivatives 174 and NBS in acetonitrile to form a product 175 with two different amine functional groups, N-trifluoromethanesulfonyl and acetamide. Obviously, solvent molecules (CH3CN) are involved in the reaction in this case [160] (Scheme 50):
Product yields are 76–94%.
With alkenes 176178 that do not have aryl groups, the reaction proceeds with the formation of N-sulfonylamidines 179181 in quantitative yields [160] (Scheme 51):
Amidines 179181 and acetamides 175 can be formed independently from the common intermediate bromonium ion 182, which could be opened via either the CH2–Br or CH–Br bond splitting, leading, respectively, to amidines 179181 or the products of hydrolysis (acetamides) 175 [160] (Scheme 52):
Unexpectedly, triflamide 83 reacts with norbornene 185 in acetonitrile with the insertion of the solvent and skeletal rearrangement, leading to the formation of iodine-containing acetamidine 186 and the tricyclic product 187 [161] (Scheme 53):
If NBS or NIS are used as the oxidizing agent, only halogen-amidines are formed, similar to those presented in Scheme 51 [161].
The reaction of triflamide 83 with 1,5-hexadiene 188 was also studied. Two products were obtained in the total isolated yield of 91%; the corresponding pyrrolidine 189 and 3,8-bis(trifluoromethanesulfonyl)-3,8-diazabicyclo[3.2.1]octane 190 formed in the ratio of 3:2 [162] (Scheme 54):
We have also proposed a simple one-pot synthesis of the 3,6-diazabicyclo[3.1.0]hexane framework based on the reaction of triflamide 83 with 2,3-dimethylbuta-1,3-diene 191 and 2,5-dimethylhexa-2,4-diene 193 in the presence of the t-BuOCl + NaI system [163]. The reaction with 2,3-dimethylbuta-1,3-diene 191 leads to a single product 192 in an 80% yield (Scheme 55):
A similar reaction of triflamide with 2,5-dimethylhexa-2,4-diene 193 carried out at –30 °C afforded the corresponding bicyclic product 194 in a 37% yield and a 1:1 mixture of pyrrolidine diastereomers 195 in a 54% yield. The mechanism of the reaction of triflamide 83 with dienes 191 or 193 is supposedly heterocyclization to 3-pyrrolines and the subsequent aziridination of 3-pyrrolines and results in 3,6-diazabicyclo[3.1.0]hexanes 192 or 194. Pyrrolidines 195 in this case are formed by the nucleophilic attack of the aziridine ring by Cl¯ or I¯ anions, presented in the reaction mixture, which is accompanied by the opening of the aziridine ring in the bicyclic product [163].

6. Conclusions

To summarize, triflimide is a unique reagent in organic synthesis. Possessing strong acidity and low nucleophilicity, it is successfully used for the design of new organic molecules, making it possible to form new C–C and C–heteroatom bonds. Given the low electron density of the nitrogen atom in Tf2NH and the steric hindrances around the nitrogen atom, it is a weak nucleophile for addition reactions. Due to these properties, the Tf2N¯ ion is actively used as a harmless counterion for cationic catalysts based on various metals to create substances that are important from a biological, pharmaceutical and industrial point of view. It is the ease of use, small loads and mild reaction conditions that make it possible to use triflimide both as a direct reagent in various reactions and as a catalyst/co-catalyst. Triflimide itself, as well as its derivatives, are effective in cascade cyclizations, cycloaddition, olefination, iodination, amination, etc. All these reactions lead to a huge number of new synthetically and biologically important objects, which undoubtedly make a significant contribution not only to fundamental organic chemistry but also in many additional applications.
Some triflamides demonstrate high anti-diabetic activity, high cytotoxicity for human cancer cell lines, anti-mycobacterial activity and anti-HIV activity. Triflamide derivatives are used as non-steroidal anti-inflammatory and anti-viral drugs, drugs for the treatment of hyperuricemia and effective drugs in the treatment of the female reproductive system diseases.
The introduction of a triflamide in aryl-containing molecules makes the modification of such substrates for biocatalytic hydroxylation possible; some triflamide compounds possess extraordinary antioxidant activity. Various triflamide compounds are widely used as peptidomimetics, organic catalysts showing high stereoselectivity and asymmetric transformations. The oxidative addition of triflamide to unsaturated substrates is a convenient method for amination and further heterocyclization under mild conditions.
Triflamides in all their various forms have certainly influenced most areas of synthetic organic chemistry in the last 10 years. Special physical and chemical properties of triflamides are used very widely, especially in medicine and pharmaceuticals for the development and synthesis of drugs and prodrugs containing the triflamide group as a key structural unit that determines biological activity, which will certainly be widely used in the future. In addition, in recent years there has been a growing interest in the use of triflamides as powerful electrophiles in a wide range of reactions. Triflamides are used as a hydrophobic solvent in lithium–sulfur batteries. It can be easily argued that the most promising field of application of triflamides are electrochemical energy storage systems. Triflamides can greatly influence the reactivity of chemical reactions compared to alkyl- and arylsulfonamides. Thus, this modifying effect will be widely used in synthetic organic chemistry and catalysis.

Author Contributions

Conceptualization, writing, original draft preparation, M.Y.M. and V.V.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (RSF) (Project 22-13-00036).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

Abbreviations

DMEDA1,2-Dimethylethylenediamine
DMAP4-Dimethylaminopyridine
HBTUHexafluorophosphate Benzotriazole Tetramethyl Uronium
EDClEDC-1-Ethyl-3-(3-dimethylaminopropyl)carbodiimide
HOBtHydroxybenzotriazole
DIH1,3-Diiodo-5,5-Dimethylhydantoin
dF-CH3-ppy2-(2,4-difluorophenyl)-5-methylpyridine
dtbbpy4,4′-Di-tert-butyl-2,2′-dipyridyl
TTBPTri-Tert-Butylphenol
TFETrifluoroethanol
Phenphenethylamine
DCE1,2-Dichloroethane
HFIPHexafluoroisopropanol
XPhosDicyclohexyl[2′,4′,6′-tris(propan-2-yl)[1,1′-biphenyl]-2-yl]phosphane
DCMDichloromethane
TEAtriethanolamine
Dppf1,1′-Bis(diphenylphosphino)ferrocene
Bpy2,2′-Bipyridine

References

  1. Shainyan, B.A.; Tolstikova, L.L. Trifluoromethanesulfonamides and Related Compounds. Chem. Rev. 2013, 113, 699–733. [Google Scholar] [CrossRef] [PubMed]
  2. Antoniotti, S.; Dalla, V.; Duñach, E. Metal Triflimidates: Better than Metal Triflates as Catalysts in Organic Synthesis—The Effect of a Highly Delocalized Counteranion. Angew. Chem. Int. Ed. 2010, 49, 7860–7888. [Google Scholar] [CrossRef]
  3. Shyamsunder, A.; Beichel, W.; Klose, P.; Pang, Q.; Scherer, H.; Hoffmann, A.; Murphy, G.K.; Krossing, I.; Nazar, L.F. Inhibiting Polysulfide Shuttle in Lithium–Sulfur Batteries through Low-Ion-Pairing Salts and a Triflamide Solvent. Angew. Chem. Int. Ed. 2017, 56, 6192–6197. [Google Scholar] [CrossRef] [PubMed]
  4. Wabnitz, T.C.; Spencer, J.B. A General, Brønsted Acid-Catalyzed Hetero-Michael Addition of Nitrogen, Oxygen, and Sulfur Nucleophiles. Org. Lett. 2003, 5, 2141–2144. [Google Scholar] [CrossRef]
  5. Croft, R.A.; Mousseau, J.J.; Choi, C.; Bull, J.A. Structurally Divergent Lithium Catalyzed Friedel–Crafts Reactions on Oxetan-3-ols: Synthesis of 3,3-Diaryloxetanes and 2,3-Dihydrobenzofurans. Chem. Eur. J. 2016, 22, 16271–16276. [Google Scholar] [CrossRef] [PubMed]
  6. Jung, M.E.; Guzaev, M. Trimethylaluminum–Triflimide Complexes for the Catalysis of Highly Hindered Diels–Alder Reactions. Org. Lett. 2012, 14, 5169–5171. [Google Scholar] [CrossRef]
  7. Kano, T.; Sakamoto, R.; Yamaguchi, Y.; Itoh, K.; Maruoka, K. One-pot cross double-Mannich reaction of acetaldehyde catalyzed by a binaphthyl-based amino sulfonamide. Chem. Commun. 2013, 49, 1118–1120. [Google Scholar] [CrossRef] [PubMed]
  8. Fan, W.-T.; Li, N.-K.; Xu, L.; Qiao, C.; Wang, X.-W. Organo-Catalyzed Asymmetric Michael–Hemiketalization–Oxa-Pictet–Spengler Cyclization for Bridged and Spiro Heterocyclic Skeletons: Oxocarbenium Ion as a Key Intermediate. Org. Lett. 2017, 19, 6626–6629. [Google Scholar] [CrossRef]
  9. Parthasarathy, K.; Praveen, C.; Jeyaveeran, J.C.; Prince, A.A.M. Gold catalyzed double condensation reaction: Synthesis, antimicrobial and cytotoxicity of spirooxindole derivatives. Bioorg. Med. Chem. Lett. 2016, 26, 4310–4317. [Google Scholar] [CrossRef]
  10. Feng, Q.; Wu, A.; Zhang, X.; Song, L.; Sun, J. An unusual formal migrative cycloaddition of aurone-derived azadienes: Synthesis of benzofuran-fused nitrogen heterocycles. Chem. Sci. 2021, 12, 7953–7957. [Google Scholar] [CrossRef]
  11. Pirenne, V.; Robert, E.G.L.; Waser, J. Catalytic (3 + 2) annulation of donor–acceptor aminocyclopropane monoesters and indoles. Chem. Sci. 2021, 12, 8706–8712. [Google Scholar] [CrossRef] [PubMed]
  12. Wong, T.H.M.; Li, X.; Ma, D.; Sun, J. HNTf2-Catalyzed Synthesis of Hydrodibenzofurans by an Epoxidation/Semipinacol Rearrangement Cascade. Org. Lett. 2020, 22, 1951–1954. [Google Scholar] [CrossRef]
  13. Fan, W.-T.; Yang, X.-P.; Lv, H.-P.; Wang, X.-W.; Wang, Z. Chiral Binaphthyl Box-Copper-Catalyzed Enantioselective Tandem Michael–Ketalization Annulations for Optically Active Aryl and Heteroaryl Fused Bicyclicnonanes. Org. Lett. 2020, 22, 3936–3941. [Google Scholar] [CrossRef] [PubMed]
  14. Avirdi, E.; Hooshmand, S.E.; Sepahvand, H.; Vishwanathan, V.; Bahadur, I.; Katata-Seru, L.M.; Varma, R.S. Ionic liquids-assisted greener preparation of silver nanoparticles. Curr. Opin. Green Sustain. Chem. 2022, 33, 100581. [Google Scholar] [CrossRef]
  15. Tohidi, M.; Ghanbari, A.; Honarasa, F. Synthesis of Copper and Silver Nanoparticles by Using Microwave-Assisted Ionic Liquid Crystal Method and Their Application for Nonenzymatic Hydrogen Peroxide Determination. Electrocatalysis 2021, 12, 350–361. [Google Scholar] [CrossRef]
  16. De Jesus, K.; Rodriguez, R.; Baek, D.L.; Fox, R.V.; Pashikanti, S.; Sharma, K. Extraction of lanthanides and actinides present in spent nuclear fuel and in electronic waste. J. Mol. Liq. 2021, 336, 116006. [Google Scholar] [CrossRef]
  17. Gujar, R.B.; Yadav, A.G.; Mohapatra, P.K.; Valsala, T.P.; Sathe, D.B.; Bhatt, R.B.; Verboom, W. Comparative uptake studies on trivalent f-cations from acidic feeds using two extraction chromatography resins containing a diglycolamide in molecular diluent and ionic liquid. J. Chromatogr. A 2021, 1641, 461999. [Google Scholar] [CrossRef]
  18. Zhao, W.; Sun, J. Triflimide (HNTf2) in Organic Synthesis. Chem. Rev. 2018, 118, 10349–10392. [Google Scholar] [CrossRef]
  19. Zheng, H.; Dong, K.; Wherritt, D.; Arman, H.; Doyle, M.P. Brønsted Acid Catalyzed Friedel–Crafts-Type Coupling and Dedinitrogenation Reactions of Vinyldiazo Compounds. Angew. Chem. Int. Ed. 2020, 59, 13613–13617. [Google Scholar] [CrossRef]
  20. Liu, M.; Zhang, J.; Zhou, H.; Yang, H.; Xia, C.; Jiang, G. Efficient hydroarylation and hydroalkenylation of vinylarenes by Brønsted acid catalysis. RSC Adv. 2016, 6, 76780–76784. [Google Scholar] [CrossRef]
  21. Reddel, J.C.T.; Wang, W.; Koukounas, K.; Thomson, R.J. Triflimide-catalyzed allylsilane annulations of benzylic alcohols for the divergent synthesis of indanes and tetralins. Chem. Sci. 2017, 8, 2156–2160. [Google Scholar] [CrossRef] [PubMed]
  22. Zheng, H.; Wang, K.; De Angelis, L.; Arman, H.D.; Doyle, M.P. Brønsted Acid Catalyzed Oxocarbenium-Olefin Metathesis/Rearrangements of 1H-Isochromene Acetals with Vinyl Diazo Compounds. J. Am. Chem. Soc. 2021, 143, 15391–15399. [Google Scholar] [CrossRef] [PubMed]
  23. Zheng, H.; Wang, K.; Faghihi, I.; Griffith, W.P.; Arman, H.; Doyle, M.P. Diverse Reactions of Vinyl Diazo Compounds with Quinone Oxonium Ions, Quinone Imine Ketals, and Eschenmoser’s Salt. ACS Catal. 2021, 11, 9869–9874. [Google Scholar] [CrossRef]
  24. Hu, W.; Zhang, F.; Chen, C.; Qi, T.; Shen, Y.; Qian, G.; Rong, Z. Metal-free C8–H functionalization of quinoline N-oxides with ynamides. Chem. Commun. 2021, 57, 6995–6998. [Google Scholar] [CrossRef]
  25. Wang, Y.; Song, L.-J.; Zhang, X.; Sun, J. Metal-Free [2 + 2 + 2] Cycloaddition of Ynamides and Nitriles: Mild and Regioselective Synthesis of Fully Substituted Pyridines. Angew. Chem. Int. Ed. 2016, 55, 9704–9708. [Google Scholar] [CrossRef]
  26. Zhao, W.; Qian, H.; Li, Z.; Sun, J. Catalytic Ring Expansion of Cyclic Hemiaminals for the Synthesis of Medium-Ring Lactams. Angew. Chem. Int. Ed. 2015, 54, 10005–10008. [Google Scholar] [CrossRef]
  27. Rathod, J.; Sharma, B.M.; Mali, P.S.; Kumar, P. Tf2NH-Catalyzed 1,6-Conjugate Addition of Vinyl Azides with p-Quinone Methides: A Mild and Efficient Method for the Synthesis of β-Bis-Arylamides. Synthesis 2017, 49, 5224–5230. [Google Scholar] [CrossRef]
  28. Gandamana, D.A.; Gagosz, F.; Chiba, S. Diastereoselective hydroalkylation of aryl alkenes enabled by Remote hydride transfer. Tetrahedron 2020, 76, 131272. [Google Scholar] [CrossRef]
  29. Zhang, Y.-Q.; Zhu, X.-Q.; Chen, Y.-B.; Tan, T.-D.; Yang, M.-Y.; Ye, L.-W. Synthesis of Isothiochroman-3-ones via Metal-Free Oxidative Cyclization of Alkynyl Thioethers. Org. Lett. 2018, 20, 7721–7725. [Google Scholar] [CrossRef]
  30. Zhang, Y.-Q.; Zhu, X.-Q.; Xu, Y.; Bu, H.-Z.; Wang, J.-L.; Zhai, T.-Y.; Zhou, J.-M.; Ye, L.-W. Synthesis of functionalized 3-isochromanones via metal-free intramolecular alkoxylation-initiated cascade cyclization. Green Chem. 2019, 21, 3023–3028. [Google Scholar] [CrossRef]
  31. Li, L.; Zhu, X.-Q.; Zhang, Y.-Q.; Bu, H.-Z.; Yuan, P.; Chen, J.; Su, J.; Deng, X.; Ye, L.-W. Metal-free alkene carbooxygenation following tandem intramolecular alkoxylation/Claisen rearrangement: Stereocontrolled access to bridged[4.2.1]lactones. Chem. Sci. 2019, 10, 3123–3129. [Google Scholar] [CrossRef] [PubMed]
  32. Zhou, B.; Zhang, Y.-Q.; Zhang, K.; Yang, M.-Y.; Chen, Y.-B.; Li, Y.; Peng, Q.; Zhu, S.-F.; Zhou, Q.-L.; Ye, L.-W. Stereoselective synthesis of medium lactams enabled by metal-free hydroalkoxylation/stereospecific[1,3]-rearrangement. Nat. Commun. 2019, 10, 3234. [Google Scholar] [CrossRef] [PubMed]
  33. Bu, H.-Z.; Li, H.-H.; Luo, W.-F.; Luo, C.; Qian, P.-C.; Ye, L.-W. Synthesis of 2H-Chromenes via Unexpected [4 + 2] Annulation of Alkynyl Thioethers with o-Hydroxybenzyl Alcohols. Org. Lett. 2020, 22, 648–652. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, L.-B.; Wang, Z.-C.; Sun, S.-Z.; Ni, S.-F.; Wen, L.-R.; Li, M. Metal-Free Catalyzed Cyclization of N-Methoxybenzamides to Construct Quaternary Carbon-Containing Isoindolinones. Chin. J. Chem. 2021, 39, 903–908. [Google Scholar] [CrossRef]
  35. Abdul-Rashed, S.; Alachouzos, G.; Brennessel, W.W.; Frontier, A.J. One-Pot Double-Annulation Strategy for the Synthesis of Unusual Fused Bis-Heterocycles. Org. Lett. 2020, 22, 4350–4354. [Google Scholar] [CrossRef]
  36. Wang, S.; Force, G.; Carpentier, J.-F.; Sarazin, Y.; Bour, C.; Gandon, V.; Lebœuf, D. Modular Synthesis of 9,10-Dihydroacridines through an ortho-C Alkenylation/Hydroarylation Sequence between Anilines and Aryl Alkynes in Hexafluoroisopropanol. Org. Lett. 2021, 23, 2565–2570. [Google Scholar] [CrossRef]
  37. Li, J.; Liu, S.; Zhong, R.; Yang, Y.; He, Y.; Yang, J.; Ma, Y.; Wang, Z. Reversal of Regioselectivity in Nucleophilic Difluoroalkylation of α,β-Enones Employing In Situ-Formed Sterically Encumbered Silylium Catalyst. Org. Lett. 2021, 23, 5859–5864. [Google Scholar] [CrossRef]
  38. Zhang, S.-L.; Lu, Y.; Li, Y.-H.; Wang, K.-Y.; Chen, J.-H.; Yang, Z. Catalytic and Enantioselective Diels–Alder Reactions of (E)-4-Oxopent-2-enoates. Org. Lett. 2017, 19, 3986–3989. [Google Scholar] [CrossRef]
  39. Buttard, F.; Berthonneau, C.; Hiebel, M.-A.; Brière, J.-F.; Suzenet, F. Organocatalytic aza-Michael Reaction to 3-Vinyl-1,2,4-triazines as a Valuable Bifunctional Platform. J. Org. Chem. 2019, 84, 3702–3714. [Google Scholar] [CrossRef]
  40. Liu, S.; Zang, Y.; Huang, H.; Sun, J. In(OTf)3-Catalyzed Synthesis of 2,3-Dihydro-1H-benzo[e]indoles and 2,3-Dihydrobenzofurans via [3 + 2] Annulation. Org. Lett. 2020, 22, 8219–8223. [Google Scholar] [CrossRef]
  41. Laws, S.W.; Howard, S.Y.; Mato, R.; Meng, S.; Fettinger, J.C.; Shaw, J.T. Organocatalytic Mukaiyama Mannich Reactions of 2,5-Bis(trimethylsilyloxy)furan. Org. Lett. 2019, 21, 5073–5077. [Google Scholar] [CrossRef] [PubMed]
  42. Qi, C.; Yang, S.; Gandon, V.; Lebœuf, D. Calcium(II)- and Triflimide-Catalyzed Intramolecular Hydroacyloxylation of Unactivated Alkenes in Hexafluoroisopropanol. Org. Lett. 2019, 21, 7405–7409. [Google Scholar] [CrossRef]
  43. Xu, X.; Shao, J.; Wang, S.R. Tuning the Cyclopropane Ring-Opening Reaction over Electronic Bias by Fusion to a Pre-Aromatic Ring: TfOH-Promoted Aromatization of Dibenzonorcaradienes to Dibenzo[f,h]isocoumarins. J. Org. Chem. 2022, 87, 3794–3798. [Google Scholar] [CrossRef] [PubMed]
  44. Mahender Reddy, K.; Thirupathi, B.; Corey, E.J. Useful Applications of Enantioselective (4 + 2)-Cycloaddition Reactions to the Synthesis of Chiral 1,2-Amino Alcohols, 1,2-Diamines, and β-Amino Acids. Org. Lett. 2017, 19, 4956–4959. [Google Scholar] [CrossRef] [PubMed]
  45. Kaldre, D.; Klose, I.; Maulide, N. Stereodivergent synthesis of 1,4-dicarbonyls by traceless charge–accelerated sulfonium rearrangement. Science 2018, 361, 664–667. [Google Scholar] [CrossRef] [PubMed]
  46. Gati, W.; Yamamoto, H. A highly diastereoselective “super silyl” governed aldol reaction: Synthesis of α,β-dioxyaldehydes and 1,2,3-triols. Chem. Sci. 2016, 7, 394–399. [Google Scholar] [CrossRef]
  47. Bae, H.Y.; List, B. Triflimide: An Overlooked High-Performance Catalyst of the Mukaiyama Aldol Reaction of Silyl Ketene Acetals with Ketones. Chem. Eur. J. 2018, 24, 13767–13772. [Google Scholar] [CrossRef]
  48. Britten, T.K.; McLaughlin, M.G. Brønsted Acid Catalyzed Peterson Olefinations. J. Org. Chem. 2020, 85, 301–305. [Google Scholar] [CrossRef]
  49. Yang, L.; Fan, W.-X.; Lin, E.; Tan, D.-H.; Li, Q.; Wang, H. Synthesis of α-CF3 and α-CF2H amines via the aminofluorination of fluorinated alkenes. Chem. Commun. 2018, 54, 5907–5910. [Google Scholar] [CrossRef]
  50. Heerdegen, D.; Junker, J.; Dittrich, S.; Mayer, P.; Bracher, F. Traceless Isoprenylation of Aldehydes via N-Boc-N-(1,1-dimethylallyl)hydrazones. Eur. J. Org. Chem. 2020, 2020, 3680–3687. [Google Scholar] [CrossRef]
  51. Zhang, F.-L.; Zhu, X.; Chiba, S. Tf2NH-Catalyzed Amide Synthesis from Vinyl Azides and Alcohols. Org. Lett. 2015, 17, 3138–3141. [Google Scholar] [CrossRef] [PubMed]
  52. Dittrich, S.; Bracher, F. Triflimide-Catalysed Rearrangement of N-(1-Trimethylsilyl)allylhydrazones Results in the Formation of Vinylsilanes and Cyclopropanes. Eur. J. Org. Chem. 2015, 2015, 8024–8033. [Google Scholar] [CrossRef]
  53. Obradors, C.; List, B. Azine Activation via Silylium Catalysis. J. Am. Chem. Soc. 2021, 143, 6817–6822. [Google Scholar] [CrossRef] [PubMed]
  54. Mase, N.; Moniruzzaman; Yamamoto, S.; Sato, K.; Narumi, T.; Yanai, H. Epimerization-suppressed organocatalytic synthesis of poly-l-lactide in supercritical carbon dioxide under plasticizing conditions. Tetrahedron Lett. 2019, 60, 150987. [Google Scholar] [CrossRef]
  55. Shoji, T.; Fukushima, K.; Menjo, T.; Yamada, Y.; Hanasaki, T.; Kikushima, K.; Takenaga, N.; Dohi, T. Triflimide-Promoted Nucleophilic C-Arylation of Halopurines to Access N7-Substituted Purine Biaryls. Chem. Pharm. Bull. 2021, 69, 886–891. [Google Scholar] [CrossRef]
  56. Wang, B.; Gandamana, D.A.; Gagosz, F.; Chiba, S. Diastereoselective Intramolecular Hydride Transfer under Brønsted Acid Catalysis. Org. Lett. 2019, 21, 2298–2301. [Google Scholar] [CrossRef]
  57. Ponra, S.; Vitale, M.R.; Michelet, V.; Ratovelomanana-Vidal, V. Practical synthesis of polysubstituted naphthalene derivatives via HNTf2-catalyzed benzannulation reaction. Arkivoc 2015, 2016, 62–81. [Google Scholar] [CrossRef]
  58. Ponra, S.; Vitale, M.R.; Michelet, V.; Ratovelomanana-Vidal, V. HNTf2-Catalyzed Regioselective Preparation of Polysubstituted Naphthalene Derivatives Through Alkyne–Aldehyde Coupling. J. Org. Chem. 2015, 80, 3250–3257. [Google Scholar] [CrossRef]
  59. Arihara, R.; Kakita, K.; Suzuki, N.; Nakamura, S.; Hashimoto, S. Glycosylation with 2-Acetamido-2-deoxyglycosyl Donors at a Low Temperature: Scope of the Non-Oxazoline Method. J. Org. Chem. 2015, 80, 4259–4277. [Google Scholar] [CrossRef]
  60. Ohura, A.; Itoh, T.; Ishida, H.; Saito, A.; Yamamoto, K. Three-Component Regioselective Synthesis of Tetrahydrofuro[2,3-d]oxazoles and Their Efficient Conversion to Oxazoles. Asian J. Org. Chem. 2017, 6, 673–676. [Google Scholar] [CrossRef]
  61. Zhang, T.; Wang, Z.; Hu, X.; Yu, M.; Deng, T.; Li, G.; Lu, H. Cesium Carboxylate-Promoted Iridium Catalyzed C–H Amidation/Cyclization with 2,2,2-Trichloroethoxycarbonyl Azide. J. Org. Chem. 2016, 81, 4898–4905. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, W.; Yang, D.; Wang, W.; Wang, S.; Zhao, H. Iridium(III)-Catalyzed Directed ortho-C(sp2)–H Amidation of Arenes with Sulfonamides. Eur. J. Org. Chem. 2018, 2018, 2071–2077. [Google Scholar] [CrossRef]
  63. Woźniak, Ł.; Cramer, N. Atropo-Enantioselective Oxidation-Enabled Iridium(III)-Catalyzed C−H Arylations with Aryl Boronic Esters. Angew. Chem. Int. Ed. 2021, 60, 18532–18536. [Google Scholar] [CrossRef] [PubMed]
  64. Nechaev, A.A.; Van Hecke, K.; Zaman, M.; Kashtanov, S.; Ungur, L.; Pereshivko, O.P.; Peshkov, V.A.; Van der Eycken, E.V. Gold-Catalyzed Post-Ugi Ipso-Cyclization with Switchable Diastereoselectivity. J. Org. Chem. 2018, 83, 8170–8182. [Google Scholar] [CrossRef]
  65. Sharma, B.M.; Rathod, J.; Gonnade, R.G.; Kumar, P. Harnessing Nucleophilicity of Allenol Ester with p-Quinone Methides via Gold Catalysis: Application to the Synthesis of Diarylmethine-Substituted Enones. J. Org. Chem. 2018, 83, 9353–9363. [Google Scholar] [CrossRef]
  66. Demmer, C.S.; Voituriez, A.; Marinetti, A. Catalytic uses of helicenes displaying phosphorus functions. C. R. Chim. 2017, 20, 860–879. [Google Scholar] [CrossRef]
  67. Hu, B.; Li, J.; Cao, W.; Lin, Q.; Yang, J.; Lin, L.; Liu, X.; Feng, X. Asymmetric Synthesis of Fused Bicyclic N,O- and O,O-Acetals via Cascade Reaction by Gold(I)/N,N′-Dioxide-Nickel(II) Bimetallic Relay Catalysis. Adv. Synth. Catal. 2018, 360, 2831–2835. [Google Scholar] [CrossRef]
  68. Nejrotti, S.; Ghinato, S.; Gini, E.C.; Scarpi, D.; Occhiato, E.G.; Maranzana, A.; Prandi, C. Gold(I)-Catalysed Hydroarylation of Lactam-Derived Enynes as an Entry to Tetrahydrobenzo[g]quinolines. Eur. J. Org. Chem. 2020, 2020, 646–653. [Google Scholar] [CrossRef]
  69. Liu, C.; Yang, F. Au-Catalyzed Stereoselective Ritter Reaction of Haloalkynes with Nitriles for (Z)-β-Halogenated Enamides. Eur. J. Org. Chem. 2019, 2019, 6867–6870. [Google Scholar] [CrossRef]
  70. Singh, R.R.; Skaria, M.; Chen, L.-Y.; Cheng, M.-J.; Liu, R.-S. Gold-catalyzed (4 + 3)-annulations of 2-alkenyl-1-alkynylbenzenes with anthranils with alkyne-dependent chemoselectivity: Skeletal rearrangement versus non-rearrangement. Chem. Sci. 2019, 10, 1201–1206. [Google Scholar] [CrossRef]
  71. Kardile, R.D.; Chao, T.-H.; Cheng, M.-J.; Liu, R.-S. Gold(I)-Catalyzed Highly Diastereo- and Enantioselective Cyclization–[4+3] Annulation Cascades between 2-(1-Alkynyl)-2-alken-1-ones and Anthranils. Angew. Chem. Int. Ed. 2020, 59, 10396–10400. [Google Scholar] [CrossRef] [PubMed]
  72. Delgado-Rebollo, M.; García-Morales, C.; Maya, C.; Prieto, A.; Echavarren, A.M.; Pérez, P.J. Coinage metal complexes bearing fluorinated N-Heterocyclic carbene ligands. J. Organomet. Chem. 2019, 898, 120856. [Google Scholar] [CrossRef]
  73. Laha, D.; Meher, K.B.; Bankar, O.S.; Bhat, R.G. Silver-Catalyzed One-Pot Access to Diastereoselective Benzo[5,6]oxepino[2,3-c]pyrroles via Formal (5 + 2)-Annulation of Donor-/Acceptor-Type Aryl Vinyl diazosuccinimide with Ketones. Asian J. Org. Chem. 2022, 11, e202200062. [Google Scholar] [CrossRef]
  74. Racys, D.T.; Sharif, S.A.I.; Pimlott, S.L.; Sutherland, A. Silver(I)-Catalyzed Iodination of Arenes: Tuning the Lewis Acidity of N-Iodosuccinimide Activation. J. Org. Chem. 2016, 81, 772–780. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, X.; Sun, X.; Fan, H.; Li, P.; Lyu, C.; Rao, W. Gold-Catalyzed Chemoselective Synthesis of Heterocycles from 3-(2-Azidophenyl)prop-2-yn-1-ols and Aldehydes. Eur. J. Org. Chem. 2016, 2016, 4265–4268. [Google Scholar] [CrossRef]
  76. Lonca, G.H.; Tejo, C.; Chan, H.L.; Chiba, S.; Gagosz, F. Gold(i)-catalyzed 6-endo-dig azide–yne cyclization: Efficient access to 2H-1,3-oxazines. Chem. Commun. 2017, 53, 736–739. [Google Scholar] [CrossRef]
  77. Shen, C.-H.; Pan, Y.; Yu, Y.-F.; Wang, Z.-S.; He, W.; Li, T.; Ye, L.-W. Facile and efficient synthesis of [1,4]oxazino[3,2-b]indoles and 1H-pyrazino[2,3-b]indoles through gold-catalyzed cascade cyclization of (azido)ynamides. J. Organomet. Chem. 2015, 795, 63–67. [Google Scholar] [CrossRef]
  78. Shen, W.-B.; Sun, Q.; Li, L.; Liu, X.; Zhou, B.; Yan, J.-Z.; Lu, X.; Ye, L.-W. Divergent synthesis of N-heterocycles via controllable cyclization of azido-diynes catalyzed by copper and gold. Nat. Commun. 2017, 8, 1748. [Google Scholar] [CrossRef]
  79. Li, N.; Lian, X.-L.; Li, Y.-H.; Wang, T.-Y.; Han, Z.-Y.; Zhang, L.; Gong, L.-Z. Gold-Catalyzed Direct Assembly of Aryl-Annulated Carbazoles from 2-Alkynyl Arylazides and Alkynes. Org. Lett. 2016, 18, 4178–4181. [Google Scholar] [CrossRef]
  80. Galenko, A.V.; Shakirova, F.M.; Galenko, E.E.; Novikov, M.S.; Khlebnikov, A.F. Fe(II)/Au(I) Relay Catalyzed Propargylisoxazole to Pyridine Isomerization: Access to 6-Halonicotinates. J. Org. Chem. 2017, 82, 5367–5379. [Google Scholar] [CrossRef]
  81. Wu, Y.; Zhu, L.; Yu, Y.; Luo, X.; Huang, X. Polysubstituted 2-Aminopyrrole Synthesis via Gold-Catalyzed Intermolecular Nitrene Transfer from Vinyl Azide to Ynamide: Reaction Scope and Mechanistic Insights. J. Org. Chem. 2015, 80, 11407–11416. [Google Scholar] [CrossRef] [PubMed]
  82. Pawar, S.K.; Sahani, R.L.; Liu, R.-S. Diversity in Gold-Catalyzed Formal Cycloadditions of Ynamides with Azidoalkenes or 2H-Azirines: [3+2] versus [4 + 3] Cycloadditions. Chem. Eur. J. 2015, 21, 10843–10850. [Google Scholar] [CrossRef] [PubMed]
  83. Jin, H.; Huang, L.; Xie, J.; Rudolph, M.; Rominger, F.; Hashmi, A.S.K. Gold-Catalyzed C—H Annulation of Anthranils with Alkynes: A Facile, Flexible, and Atom-Economical Synthesis of Unprotected 7-Acylindoles. Angew. Chem. Int. Ed. 2016, 55, 794–797. [Google Scholar] [CrossRef] [PubMed]
  84. Zhou, A.-H.; He, Q.; Shu, C.; Yu, Y.-F.; Liu, S.; Zhao, T.; Zhang, W.; Lu, X.; Ye, L.-W. Atom-economic generation of gold carbenes: Gold-catalyzed formal [3 + 2] cycloaddition between ynamides and isoxazoles. Chem. Sci. 2015, 6, 1265–1271. [Google Scholar] [CrossRef]
  85. Xiao, X.-Y.; Zhou, A.-H.; Shu, C.; Pan, F.; Li, T.; Ye, L.-W. Atom-Economic Synthesis of Fully Substituted 2-Aminopyrroles via Gold-Catalyzed Formal [3 + 2] Cycloaddition between Ynamides and Isoxazoles. Chem. Asian J. 2015, 10, 1854–1858. [Google Scholar] [CrossRef]
  86. Giri, S.S.; Liu, R.-S. Gold-catalyzed [4 + 3]- and [4 + 2]-annulations of 3-en-1-ynamides with isoxazoles via novel 6π-electrocyclizations of 3-azahepta trienyl cations. Chem. Sci. 2018, 9, 2991–2995. [Google Scholar] [CrossRef]
  87. Pan, Y.; Chen, G.-W.; Shen, C.-H.; He, W.; Ye, L.-W. Synthesis of fused isoquinolines via gold-catalyzed tandem alkyne amination/intramolecular O–H insertion. Org. Chem. Front. 2016, 3, 491–495. [Google Scholar] [CrossRef]
  88. Shu, C.; Wang, Y.-H.; Zhou, B.; Li, X.-L.; Ping, Y.-F.; Lu, X.; Ye, L.-W. Generation of α-Imino Gold Carbenes through Gold-Catalyzed Intermolecular Reaction of Azides with Ynamides. J. Am. Chem. Soc. 2015, 137, 9567–9570. [Google Scholar] [CrossRef]
  89. Zhou, B.; Zhang, Y.-Q.; Liu, X.; Ye, L.-W. Gold-catalyzed intermolecular reaction of ynamides with 3-indolyl azides via an unexpected 1,2-alkyl migration. Sci. Bull. 2017, 62, 1201–1206. [Google Scholar] [CrossRef]
  90. Jin, H.; Tian, B.; Song, X.; Xie, J.; Rudolph, M.; Rominger, F.; Hashmi, A.S.K. Gold-Catalyzed Synthesis of Quinolines from Propargyl Silyl Ethers and Anthranils through the Umpolung of a Gold Carbene Carbon. Angew. Chem. Int. Ed. 2016, 55, 12688–12692. [Google Scholar] [CrossRef]
  91. González, J.; Santamaría, J.; Suárez-Sobrino, Á.L.; Ballesteros, A. One-Pot and Regioselective Gold-Catalyzed Synthesis of 2-Imidazolyl-1-pyrazolylbenzenes from 1-Propargyl-1H-benzotriazoles, Alkynes and Nitriles through α-Imino Gold(I) Carbene Complexes. Adv. Synth. Catal. 2016, 358, 1398–1403. [Google Scholar] [CrossRef]
  92. Allegue, D.; González, J.; Fernández, S.; Santamaría, J.; Ballesteros, A. Regiodivergent Control in the Gold(I) Catalyzed Synthesis of 7-Pyrazolylindoles from 1-Propargyl-1H-benzotriazoles and Ynamides through α-Imino Gold(I) Carbene Complexes. Adv. Synth. Catal. 2019, 361, 758–768. [Google Scholar] [CrossRef]
  93. Mei, L.-Y.; Wei, Y.; Tang, X.-Y.; Shi, M. Catalyst-Dependent Stereodivergent and Regioselective Synthesis of Indole-Fused Heterocycles through Formal Cycloadditions of Indolyl-Allenes. J. Am. Chem. Soc. 2015, 137, 8131–8137. [Google Scholar] [CrossRef] [PubMed]
  94. Ruan, P.-P.; Li, H.-H.; Liu, X.; Zhang, T.; Zuo, S.-X.; Zhu, C.; Ye, L.-W. Synthesis of α,β-Unsaturated Amidines through Gold-Catalyzed Intermolecular Reaction of Azides with Ynamides. J. Org. Chem. 2017, 82, 9119–9125. [Google Scholar] [CrossRef] [PubMed]
  95. Shu, C.; Wang, Y.-H.; Shen, C.-H.; Ruan, P.-P.; Lu, X.; Ye, L.-W. Gold-Catalyzed Intermolecular Ynamide Amination-Initiated Aza-Nazarov Cyclization: Access to Functionalized 2-Aminopyrroles. Org. Lett. 2016, 18, 3254–3257. [Google Scholar] [CrossRef]
  96. Zhai, R.L.; Xue, Y.S.; Liang, T.; Mi, J.J.; Xu, Z. Regioselective Arene and Heteroarene Functionalization: N-Alkenoxypyridinium Salts as Electrophilic Alkylating Agents for the Synthesis of α-Aryl/α-Heteroaryl Ketones. J. Org. Chem. 2018, 83, 10051–10059. [Google Scholar] [CrossRef]
  97. Zhou, W.; Voituriez, A. Gold(I)-Catalyzed Synthesis of Highly Substituted 1,4-Dicarbonyl Derivatives via Sulfonium (3,3)-Sigmatropic Rearrangement. Org. Lett. 2020, 23, 247–252. [Google Scholar] [CrossRef]
  98. Hamada, N.; Yoshida, Y.; Oishi, S.; Ohno, H. Gold-Catalyzed Cascade Reaction of Skipped Diynes for the Construction of a Cyclohepta[b]pyrrole Scaffold. Org. Lett. 2017, 19, 3875–3878. [Google Scholar] [CrossRef]
  99. Li, W.; Yu, B. Gold-catalyzed glycosylation in the synthesis of complex carbohydrate-containing natural products. Chem. Soc. Rev. 2018, 47, 7954–7984. [Google Scholar] [CrossRef]
  100. Yu, B. Gold(I)-Catalyzed Glycosylation with Glycosyl o-Alkynylbenzoates as Donors. Acc. Chem. Res. 2018, 51, 507–516. [Google Scholar] [CrossRef]
  101. Webster, S.; O’Rourke, K.M.; Fletcher, C.; Pimlott, S.L.; Sutherland, A.; Lee, A.-L. Rapid Iododeboronation with and without Gold Catalysis: Application to Radiolabelling of Arenes. Chem. Eur. J. 2018, 24, 937–943. [Google Scholar] [CrossRef] [PubMed]
  102. Liu, Y.; Wang, X.; Wang, Y.; Du, C.; Shi, H.; Jin, S.; Jiang, C.; Xiao, J.; Cheng, M. Gold(I)-Catalyzed Synthesis of Unsymmetrical Ethers Using Alcohols as Alkylating Reagents. Adv. Synth. Catal. 2015, 357, 1029–1036. [Google Scholar] [CrossRef]
  103. Shcherbakov, N.V.; Dar’in, D.V.; Kukushkin, V.Y.; Dubovtsev, A.Y. Gold-Catalyzed Nitrene Transfer from Benzofuroxans to N-Allylynamides: Synthesis of 3-Azabicyclo[3.1.0]hexanes. J. Org. Chem. 2021, 86, 12964–12972. [Google Scholar] [CrossRef] [PubMed]
  104. Huang, H.; Zhang, T.; Sun, J. Mild C–C Bond Formation via Lewis Acid Catalyzed Oxetane Ring Opening with Soft Carbon Nucleophiles. Angew. Chem. Int. Ed. 2021, 60, 2668–2673. [Google Scholar] [CrossRef] [PubMed]
  105. Croft, R.A.; Mousseau, J.J.; Choi, C.; Bull, J.A. Oxetane ethers are formed reversibly in the lithium-catalyzed Friedel–Crafts alkylation of phenols with oxetanols: Synthesis of dihydrobenzofurans, diaryloxetanes, and oxetane ethers. Tetrahedron 2018, 74, 5427–5435. [Google Scholar] [CrossRef]
  106. Schroeder, S.; Strauch, C.; Gaelings, N.; Niggemann, M. Vinyl Triflimides—A Case of Assisted Vinyl Cation Formation. Angew. Chem. Int. Ed. 2019, 58, 5119–5123. [Google Scholar] [CrossRef]
  107. Ochiai, M.; Okubo, T.; Miyamoto, K. Weakly Nucleophilic Conjugate Bases of Superacids as Powerful Nucleophiles in Vinylic Bimolecular Nucleophilic Substitutions of Simple β-Alkylvinyl(aryl)-λ3-bromanes. J. Am. Chem. Soc. 2011, 133, 3342–3344. [Google Scholar] [CrossRef]
  108. Chuchmareva, M.; Strauch, C.; Schröder, S.; Collong, A.; Niggemann, M. Regio- and Stereoselective N-addition to an Open Bromo Vinyl Cation. Tetrahedron Lett. 2021, 74, 153173. [Google Scholar] [CrossRef]
  109. Hao, L.; Pan, Q.; Zhang, C.; Wang, S.; Wang, W.; Zhang, J.; Bai, L.; Wang, Y. General, Straightforward, and Atom-Economical Synthesis of Vinyl Triflimides. Chem. Eur. J. 2021, 27, 12272–12275. [Google Scholar] [CrossRef]
  110. Dada, R.; Sulthan, M.; Yaragorla, S. Calcium-Catalyzed Stereoselective Tandem [4 + 2] and [3 + 2] Annulation Reaction for the Synthesis of Dihydropyrrolo[1,2-a]quinolines. Org. Lett. 2019, 22, 279–283. [Google Scholar] [CrossRef]
  111. Wang, S.; Guillot, R.; Carpentier, J.-F.; Sarazin, Y.; Bour, C.; Gandon, V.; Lebœuf, D. Synthesis of Bridged Tetrahydrobenzo[b]azepines and Derivatives through an Aza-Piancatelli Cyclization/Michael Addition Sequence. Angew. Chem. Int. Ed. 2020, 59, 1134–1138. [Google Scholar] [CrossRef] [PubMed]
  112. Qi, C.; Gandon, V.; Lebœuf, D. Calcium(II)-Catalyzed Alkenylation of N-Acyliminiums and Related Ions with Vinylboronic Acids. Adv. Synth. Catal. 2017, 359, 2671–2675. [Google Scholar] [CrossRef]
  113. Cabrero-Antonino, J.R.; Leyva-Pérez, A.; Corma, A. Iron(III) Triflimide as a Catalytic Substitute for Gold(I) in Hydroaddition Reactions to Unsaturated Carbon–Carbon Bonds. Chem. Eur. J. 2013, 19, 8627–8633. [Google Scholar] [CrossRef] [PubMed]
  114. Racys, D.T.; Warrilow, C.E.; Pimlott, S.L.; Sutherland, A. Highly Regioselective Iodination of Arenes via Iron(III)-Catalyzed Activation of N-Iodosuccinimide. Org. Lett. 2015, 17, 4782–4785. [Google Scholar] [CrossRef]
  115. Mostafa, M.A.B.; Bowley, R.M.; Racys, D.T.; Henry, M.C.; Sutherland, A. Iron(III)-Catalyzed Chlorination of Activated Arenes. J. Org. Chem. 2017, 82, 7529–7537. [Google Scholar] [CrossRef] [PubMed]
  116. Mostafa, M.A.B.; Calder, E.D.D.; Racys, D.T.; Sutherland, A. Intermolecular Aryl C−H Amination through Sequential Iron and Copper Catalysis. Chem. Eur. J. 2017, 23, 1044–1047. [Google Scholar] [CrossRef] [PubMed]
  117. Henry, M.C.; Abbinante, V.M.; Sutherland, A. Iron-Catalyzed Regioselective Synthesis of 2-Arylbenzoxazoles and 2-Arylbenzothiazoles via Alternative Reaction Pathways. Eur. J. Org. Chem. 2020, 2020, 2819–2826. [Google Scholar] [CrossRef]
  118. Xin, G.; Ming, Y.; Ji, C.; Wei, Z.; Li, S.; Morris-Natschke, S.L.; Zhang, X.; Yu, K.; Li, Y.; Zhang, B.; et al. Novel potent antiplatelet thrombotic agent derived from biguanide for ischemic stroke. Eur. J. Med. Chem. 2020, 200, 112462. [Google Scholar] [CrossRef]
  119. Markowicz-Piasecka, M.; Huttunen, J.; Zajda, A.; Sikora, J.; Huttunen, K.M. Sulfonamide metformin derivatives induce mitochondrial-associated apoptosis and cell cycle arrest in breast cancer cells. Chem.-Biol. Interact. 2022, 352, 109795. [Google Scholar] [CrossRef]
  120. Zajda, A.; Sikora, J.; Huttunen, K.M.; Markowicz-Piasecka, M. Structural Comparison of Sulfonamide-Based Derivatives That Can Improve Anti-Coagulation Properties of Metformin. Int. J. Mol. Sci. 2022, 23, 4132. [Google Scholar] [CrossRef]
  121. O’Hanlon, J.A.; Ren, X.; Morris, M.; Wong, L.L.; Robertson, J. Hydroxylation of anilides by engineered cytochrome P450BM3. Org. Biomol. Chem. 2017, 15, 8780–8787. [Google Scholar] [CrossRef] [PubMed]
  122. Gómez-García, O.; Gómez, E.; Monzón-González, C.; Ramírez-Apan, T.; Álvarez-Toledano, C. An Efficient Strategy for the Synthesis of 1-(Trifluoromethylsulfonamido)propan-2-yl Esters and the Evaluation of Their Cytotoxic Activity. Chem. Pharm. Bull. 2017, 65, 248–252. [Google Scholar] [CrossRef] [PubMed]
  123. Hong, B.-T.; Chen, C.-L.; Fang, J.-M.; Tsai, K.-C.; Wang, S.-Y.; Huang, W.-I.; Cheng, Y.-S.E.; Wong, C.-H. Oseltamivir hydroxamate and acyl sulfonamide derivatives as influenza neuraminidase inhibitors. Bioorg. Med. Chem. 2014, 22, 6647–6654. [Google Scholar] [CrossRef] [PubMed]
  124. Hu, Y.; Chen, B.; Lei, Z.; Zhao, H.; Zhu, H.; Quan, P.; Tian, Y. Synthesis and Biological Evaluation of NH2-Sulfonyl Oseltamivir Analogues as Influenza Neuraminidase Inhibitors. Molecules 2019, 24, 2176. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, S.; Su, M.; Wang, J.; Li, Z.; Zhang, L.; Ji, X.; Li, J.; Li, J.; Liu, H. (R)-3-Amino-1-((3aS,7aS)-octahydro-1H-indol-1-yl)-4-(2,4,5-trifluorophenyl)butan-1-one derivatives as potent inhibitors of dipeptidyl peptidase-4: Design, synthesis, biological evaluation, and molecular modeling. Bioorg. Med. Chem. 2014, 22, 6684–6693. [Google Scholar] [CrossRef]
  126. Ji, Y.; Chen, X.; Chen, H.; Zhang, X.; Fan, Z.; Xie, L.; Ma, B.; Zhu, C. Designing of acyl sulphonamide based quinoxalinones as multifunctional aldose reductase inhibitors. Bioorg. Med. Chem. 2019, 27, 1658–1669. [Google Scholar] [CrossRef]
  127. Liedtke, A.J.; Adeniji, A.O.; Chen, M.; Byrns, M.C.; Jin, Y.; Christianson, D.W.; Marnett, L.J.; Penning, T.M. Development of Potent and Selective Indomethacin Analogues for the Inhibition of AKR1C3 (Type 5 17β-Hydroxysteroid Dehydrogenase/Prostaglandin F Synthase) in Castrate-Resistant Prostate Cancer. J. Med. Chem. 2013, 56, 2429–2446. [Google Scholar] [CrossRef]
  128. Nuti, E.; Cuffaro, D.; Bernardini, E.; Camodeca, C.; Panelli, L.; Chaves, S.; Ciccone, L.; Tepshi, L.; Vera, L.; Orlandini, E.; et al. Development of Thioaryl-Based Matrix Metalloproteinase-12 Inhibitors with Alternative Zinc-Binding Groups: Synthesis, Potentiometric, NMR, and Crystallographic Studies. J. Med. Chem. 2018, 61, 4421–4435. [Google Scholar] [CrossRef]
  129. Newcombe, S.; Bobin, M.; Shrikhande, A.; Gallop, C.; Pace, Y.; Yong, H.; Gates, R.; Chaudhuri, S.; Roe, M.; Hoffmann, E.; et al. Gold amides as anticancer drugs: Synthesis and activity studies. Org. Biomol. Chem. 2013, 11, 3255–3260. [Google Scholar] [CrossRef]
  130. Renard, J.-F.; Lecomte, F.; Hubert, P.; de Leval, X.; Pirotte, B. N-(3-Arylaminopyridin-4-yl)alkanesulfonamides as pyridine analogs of nimesulide: Cyclooxygenases inhibition, anti-inflammatory studies and insight on metabolism. Eur. J. Med. Chem. 2014, 74, 12–22. [Google Scholar] [CrossRef]
  131. Skerratt, S.E.; de Groot, M.J.; Phillips, C. Discovery of a novel binding pocket for CYP 2C9 inhibitors: Crystallography, pharmacophore modelling and inhibitor SAR. Med. Chem. Commun. 2016, 7, 813–819. [Google Scholar] [CrossRef]
  132. Suthagar, K.; Polson, M.I.J.; Fairbanks, A.J. Unexpected furanose/pyranose equilibration of N-glycosyl sulfonamides, sulfamides and sulfamates. Org. Biomol. Chem. 2015, 13, 6573–6579. [Google Scholar] [CrossRef] [PubMed]
  133. Winton, V.J.; Aldrich, C.; Kiessling, L.L. Carboxylate Surrogates Enhance the Antimycobacterial Activity of UDP-Galactopyranose Mutase Probes. ACS Infect. Dis. 2016, 2, 538–543. [Google Scholar] [CrossRef]
  134. Zhao, A.; Li, Y.; Orahoske, C.M.; Schnur, B.; Sabbagh, A.; Zhang, W.; Li, B.; Su, B. Lead optimization of selective tubulin inhibitors as anti-trypanosomal agents. Bioorg. Med. Chem. 2019, 27, 1517–1528. [Google Scholar] [CrossRef] [PubMed]
  135. Steverding, D.; Tyler, K.M. Novel antitrypanosomal agents. Expert Opin. Investig. Drugs 2005, 14, 939–955. [Google Scholar] [CrossRef] [PubMed]
  136. Yang, X.; Pang, X.; Fan, L.; Li, X.; Chen, Y. Synthesis and evaluation of sulfonamide derivatives as potent Human Uric Acid Transporter 1 (hURAT1) inhibitors. Bioorg. Med. Chem. Lett. 2017, 27, 1919–1922. [Google Scholar] [CrossRef] [PubMed]
  137. Rafique, R.; Khan, K.M.; Chigurupati, S.; Wadood, A.; Rehman, A.U.; Salar, U.; Venugopal, V.; Shamim, S.; Taha, M.; Perveen, S. Synthesis, in vitro α-amylase inhibitory, and radicals (DPPH & ABTS) scavenging potentials of new N-sulfonohydrazide substituted indazoles. Bioorg. Chem. 2020, 94, 103410. [Google Scholar] [CrossRef]
  138. Dunn, D.; Iqbal, M.; Chatterjee, S. Synthesis and activity of proteasome inhibitors. J. Enzyme Inhib. Med. Chem. 2014, 29, 55–58. [Google Scholar] [CrossRef]
  139. Kumari, G.; Singh, R.K. Molecular Modeling, Synthesis, and Anti-HIV Activity of Novel Isoindolinedione Analogues as Potent Non-nucleoside Reverse Transcriptase Inhibitors. Chem. Biol. Drug Des. 2016, 87, 200–212. [Google Scholar] [CrossRef]
  140. Yanai, K.; Togo, H. Preparation of phenanthridines from N-(o-arylbenzyl)trifluoromethanesulfonamides with 1,3-diiodo-5,5-dimethylhydantoin. Tetrahedron 2020, 76, 131503. [Google Scholar] [CrossRef]
  141. Huang, G.; Solano, C.M.; Su, Y.; Ezzat, N.; Matsui, S.; Huang, L.; Chakrabarti, D.; Yuan, Y. Microwave-assisted, rapid synthesis of 2-vinylquinolines and evaluation of their antimalarial activity. Tetrahedron Lett. 2019, 60, 1736–1740. [Google Scholar] [CrossRef] [PubMed]
  142. Takeshima, A.; Kano, T.; Maruoka, K. Synthesis of Phenylcyclopropane-Based Secondary Amine Catalysts and Their Applications in Enamine Catalysis. Org. Lett. 2019, 21, 8071–8074. [Google Scholar] [CrossRef] [PubMed]
  143. Yang, X.; Xie, Z.; Li, Y.; Zhang, Y.-Q. Enantioselective aerobic oxidative cross-dehydrogenative coupling of glycine derivatives with ketones and aldehydes via cooperative photoredox catalysis and organocatalysis. Chem. Sci. 2020, 11, 4741–4746. [Google Scholar] [CrossRef] [PubMed]
  144. Gorde, A.B.; Ramapanicker, R. d-Prolyl-2-(trifluoromethylsulfonamidopropyl)pyrrolidine: An Organocatalyst for Asymmetric Michael Addition of Aldehydes to β-Nitroalkenes at Ambient Conditions. J. Org. Chem. 2019, 84, 1523–1533. [Google Scholar] [CrossRef]
  145. Skrzyńska, A.; Frankowski, S.; Topolska, A.; Dyguda, M.; Gao, X.-Y.; Xu, C.-J.; Chen, Y.-C.; Albrecht, Ł. Enantioselective H-bond-directed vinylogous iminium ion strategy for the functionalization of vinyl-substituted heteroaryl aldehydes. Chem. Commun. 2021, 57, 1667–1670. [Google Scholar] [CrossRef]
  146. Ramchandra, G.T. Synthesis of Axially Chiral Amines and its Application in Asymmetric Catalysis. IJRAR Int. J. Res. Anal. Rev. 2019, 6, 392y–408y. [Google Scholar]
  147. Schwengers, S.A.; De, C.K.; Grossmann, O.; Grimm, J.A.A.; Sadlowski, N.R.; Gerosa, G.G.; List, B. Unified Approach to Imidodiphosphate-Type Brønsted Acids with Tunable Confinement and Acidity. J. Am. Chem. Soc. 2021, 143, 14835–14844. [Google Scholar] [CrossRef]
  148. Yuan, Y.-H.; Han, X.; Zhu, F.-P.; Tian, J.-M.; Zhang, F.-M.; Zhang, X.-M.; Tu, Y.-Q.; Wang, S.-H.; Guo, X. Development of bifunctional organocatalysts and application to asymmetric total synthesis of naucleofficine I and II. Nat. Commun. 2019, 10, 3394. [Google Scholar] [CrossRef]
  149. Suto, T.; Yanagita, Y.; Nagashima, Y.; Takikawa, S.; Kurosu, Y.; Matsuo, N.; Miura, K.; Simizu, S.; Sato, T.; Chida, N. Unified Total Synthesis of Madangamine Alkaloids. Bull. Chem. Soc. Jpn. 2019, 92, 545–571. [Google Scholar] [CrossRef]
  150. Ashley, M.A.; Yamauchi, C.; Chu, J.C.K.; Otsuka, S.; Yorimitsu, H.; Rovis, T. Photoredox-Catalyzed Site-Selective α-C(sp3)−H Alkylation of Primary Amine Derivatives. Angew. Chem. Int. Ed. 2019, 58, 4002–4006. [Google Scholar] [CrossRef]
  151. Vuillermet, F.; Bourret, J.; Pelletier, G. Synthesis of Imidazo[1,2-a]pyridines: Triflic Anhydride-Mediated Annulation of 2H-Azirines with 2-Chloropyridines. J. Org. Chem. 2021, 86, 388–402. [Google Scholar] [CrossRef] [PubMed]
  152. Kawamoto, T.; Ikawa, K.; Kamimura, A. One-Pot Synthesis of CF3-Substituted Vinyl Trifluoromethanesulfonamides from Imines and Trifluoromethanesulfonic Anhydride. J. Org. Chem. 2021, 86, 15818–15824. [Google Scholar] [CrossRef] [PubMed]
  153. Dreier, P.; Pipertzis, A.; Spyridakou, M.; Mathes, R.; Floudas, G.; Frey, H. Introduction of Trifluoromethanesulfonamide Groups in Poly(ethylene oxide): Ionic Conductivity of Single-Ion-Conducting Block Copolymer Electrolytes. Macromolecules 2022, 55, 1342–1353. [Google Scholar] [CrossRef]
  154. Velasco-Rubio, Á.; Varela, J.A.; Saá, C. Palladium-Catalyzed [5 + 2] Heteroannulation of Phenethylamides with 1,3-Dienes to Dopaminergic 3-Benzazepines. Org. Lett. 2020, 22, 3591–3595. [Google Scholar] [CrossRef] [PubMed]
  155. Markus, A.; Schepmann, D.; Wünsch, B. Synthesis of oxazolo-annulated 3-benzazepines designed by merging two negative allosteric NMDA receptor modulators. Arch. Pharm. 2022, 355, 2200020. [Google Scholar] [CrossRef]
  156. Li, X.; Peters, B.B.C.; Tan, M.; He, L.; Yang, J.; Andersson, P.G.; Zhou, T. Mono-N-Alkylation of Sulfonamides with Alcohols Catalyzed by Iridium N-Heterocyclic Carbene-Phosphine Complexes. Asian J. Org. Chem. 2022, 11, e202200178. [Google Scholar] [CrossRef]
  157. Wang, N.; Liu, L.; Xu, W.; Zhang, M.; Huang, Z.; Shi, D.; Zhao, Y. Rhodium(III)-Catalyzed Oxidative Annulation of Ketoximes with Sulfonamide: A Direct Approach to Indazoles. Org. Lett. 2019, 21, 365–368. [Google Scholar] [CrossRef]
  158. Li, Z.; Zhang, H.; Yu, S. NaClO-Promoted Atroposelective Couplings of 3-Substituted Indoles with Amino Acid Derivatives. Org. Lett. 2019, 21, 4754–4758. [Google Scholar] [CrossRef]
  159. Yang, M.; Su, B.; Wang, Y.; Chen, K.; Jiang, X.; Zhang, Y.-F.; Zhang, X.-S.; Chen, G.; Cheng, Y.; Cao, Z.; et al. Silver-catalysed direct amination of unactivated C–H bonds of functionalized molecules. Nat. Commun. 2014, 5, 4707. [Google Scholar] [CrossRef]
  160. Moskalik, M.Y.; Garagan, I.A.; Astakhova, V.V.; Sterkhova, I.V.; Shainyan, B.A. Solvent-dependent oxidative triflamidation of alkenes and N(O)-Heterocyclization of the products. Tetrahedron 2021, 88, 132145. [Google Scholar] [CrossRef]
  161. Moskalik, M.Y.; Shainyan, B.A.; Ushakov, I.A.; Sterkhova, I.V.; Astakhova, V.V. Oxidant effect, skeletal rearrangements and solvent interception in oxidative triflamidation of norbornene and 2,5-norbornadiene. Tetrahedron 2020, 76, 131018. [Google Scholar] [CrossRef]
  162. Shainyan, B.A.; Moskalik, M.Y.; Astakhova, V.V.; Schilde, U. Novel design of 3,8-diazabicyclo[3.2.1]octane framework in oxidative sulfonamidation of 1,5-hexadiene. Tetrahedron 2014, 70, 4547–4551. [Google Scholar] [CrossRef]
  163. Moskalik, M.Y.; Astakhova, V.V.; Schilde, U.; Sterkhova, I.V.; Shainyan, B.A. Assembling of 3,6-diazabicyclo[3.1.0]hexane framework in oxidative triflamidation of substituted buta-1,3-dienes. Tetrahedron 2014, 70, 8636–8641. [Google Scholar] [CrossRef]
Scheme 1. Metal-free intramolecular alkoxylation-initiated cascade cyclization in the presence of Tf2NH.
Scheme 1. Metal-free intramolecular alkoxylation-initiated cascade cyclization in the presence of Tf2NH.
Molecules 27 05201 sch001
Scheme 2. Intramolecular cyclization of N-methoxybenzamides in the presence of Tf2NH.
Scheme 2. Intramolecular cyclization of N-methoxybenzamides in the presence of Tf2NH.
Molecules 27 05201 sch002
Scheme 3. One-pot synthesis of bicyclic heterocycles by the reaction of carboamination of alkynols with cyclic aminals using metal-free approach.
Scheme 3. One-pot synthesis of bicyclic heterocycles by the reaction of carboamination of alkynols with cyclic aminals using metal-free approach.
Molecules 27 05201 sch003
Scheme 4. One-pot assembly of 9,10-dihydroacridine scaffolds.
Scheme 4. One-pot assembly of 9,10-dihydroacridine scaffolds.
Molecules 27 05201 sch004
Scheme 5. Difluoroalkylation of α,β-enones in the presence of TMSNTf2 or Tf2NH.
Scheme 5. Difluoroalkylation of α,β-enones in the presence of TMSNTf2 or Tf2NH.
Molecules 27 05201 sch005
Scheme 6. Reaction of vinyl sulfoxides with ynamides in the presence of Tf2NH in water.
Scheme 6. Reaction of vinyl sulfoxides with ynamides in the presence of Tf2NH in water.
Molecules 27 05201 sch006
Scheme 7. Aldol reaction of (Z)-tris(trimethylsilyl)silyl enol ethers in the presence of triflimide.
Scheme 7. Aldol reaction of (Z)-tris(trimethylsilyl)silyl enol ethers in the presence of triflimide.
Molecules 27 05201 sch007
Scheme 8. Mukaiyama aldol reaction in the presence of Tf2NH.
Scheme 8. Mukaiyama aldol reaction in the presence of Tf2NH.
Molecules 27 05201 sch008
Scheme 9. Peterson olefination and aminofluorination of gem-difluoroalkenes and monofluoroalkenes in the presence of Tf2NH.
Scheme 9. Peterson olefination and aminofluorination of gem-difluoroalkenes and monofluoroalkenes in the presence of Tf2NH.
Molecules 27 05201 sch009
Scheme 10. Cyclization of ortho- and para-substituted benzamides 31 with trichloroethoxycarbonyl azide 32 in the presence of a complex based on silver triflimide 33.
Scheme 10. Cyclization of ortho- and para-substituted benzamides 31 with trichloroethoxycarbonyl azide 32 in the presence of a complex based on silver triflimide 33.
Molecules 27 05201 sch010
Scheme 11. Preparation of [1,3]dioxino[5,4-b]indole derivatives.
Scheme 11. Preparation of [1,3]dioxino[5,4-b]indole derivatives.
Molecules 27 05201 sch011
Scheme 12. Regioselective functionalization of arenes and heteroarenes to α-aryl- and α-heteroaryl ketones.
Scheme 12. Regioselective functionalization of arenes and heteroarenes to α-aryl- and α-heteroaryl ketones.
Molecules 27 05201 sch012
Scheme 13. Synthesis of 3 azabicyclo[3.1.0]hexanes from benzofuroxans and N-allylynamides catalyzed by Au+.
Scheme 13. Synthesis of 3 azabicyclo[3.1.0]hexanes from benzofuroxans and N-allylynamides catalyzed by Au+.
Molecules 27 05201 sch013
Scheme 14. Friedel–Crafts reaction of oxetanols and phenols in the presence of LiNTf2.
Scheme 14. Friedel–Crafts reaction of oxetanols and phenols in the presence of LiNTf2.
Molecules 27 05201 sch014
Scheme 15. Hydroaminosulfonation of phenylacetylene.
Scheme 15. Hydroaminosulfonation of phenylacetylene.
Molecules 27 05201 sch015
Scheme 16. Synthesis of bromvinyl triflimide.
Scheme 16. Synthesis of bromvinyl triflimide.
Molecules 27 05201 sch016
Scheme 17. Reactions of aryl-, vinyl-, alkyl- and silyl-substituted terminal alkynes with triflimide.
Scheme 17. Reactions of aryl-, vinyl-, alkyl- and silyl-substituted terminal alkynes with triflimide.
Molecules 27 05201 sch017
Scheme 18. Calcium triflimide-catalyzed stereoselective tandem of [4 + 2]- and [3 + 2]-annulation for the synthesis of ropyrrolo[1,2a]quinolines.
Scheme 18. Calcium triflimide-catalyzed stereoselective tandem of [4 + 2]- and [3 + 2]-annulation for the synthesis of ropyrrolo[1,2a]quinolines.
Molecules 27 05201 sch018
Scheme 19. Fe(NTf2)3-catalyzed iodination of arenes.
Scheme 19. Fe(NTf2)3-catalyzed iodination of arenes.
Molecules 27 05201 sch019
Scheme 20. Strategy for regioselective CH-amination of arenes in the presence of Fe(NTf2)3/CuI.
Scheme 20. Strategy for regioselective CH-amination of arenes in the presence of Fe(NTf2)3/CuI.
Molecules 27 05201 sch020
Scheme 21. Synthesis of N-trifluoromethanesulfonyl derivative of metformin.
Scheme 21. Synthesis of N-trifluoromethanesulfonyl derivative of metformin.
Molecules 27 05201 sch021
Scheme 22. Biocatalytic hydroxylation of N-trifluoromethanesulfonyl substituted aniline.
Scheme 22. Biocatalytic hydroxylation of N-trifluoromethanesulfonyl substituted aniline.
Molecules 27 05201 sch022
Scheme 23. Ring-opening of the 5-methyl-2-phenyl-4,5-dihydrooxazole system by triflic anhydride.
Scheme 23. Ring-opening of the 5-methyl-2-phenyl-4,5-dihydrooxazole system by triflic anhydride.
Molecules 27 05201 sch023
Scheme 24. Synthesis of triflamide-containing anti-pancreatic carcinoma derivative.
Scheme 24. Synthesis of triflamide-containing anti-pancreatic carcinoma derivative.
Molecules 27 05201 sch024
Scheme 25. Synthesis of oseltamivir acyl triflamide derivative.
Scheme 25. Synthesis of oseltamivir acyl triflamide derivative.
Molecules 27 05201 sch025
Scheme 26. Synthesis of DPP-4 inhibitor.
Scheme 26. Synthesis of DPP-4 inhibitor.
Molecules 27 05201 sch026
Scheme 27. Synthesis of quinoxalinone-based triflamide derivatives.
Scheme 27. Synthesis of quinoxalinone-based triflamide derivatives.
Molecules 27 05201 sch027
Scheme 28. Synthesis of indomethacin triflamide analogues.
Scheme 28. Synthesis of indomethacin triflamide analogues.
Molecules 27 05201 sch028
Scheme 29. Synthesis of triflamide-containing selective matrix metalloproteinase-12 (MMP-12) inhibitor.
Scheme 29. Synthesis of triflamide-containing selective matrix metalloproteinase-12 (MMP-12) inhibitor.
Molecules 27 05201 sch029
Scheme 30. Synthesis of chiral complexes based on N-trifluoromethanesulfonyl-substituted amino acids and Au(I).
Scheme 30. Synthesis of chiral complexes based on N-trifluoromethanesulfonyl-substituted amino acids and Au(I).
Molecules 27 05201 sch030
Scheme 31. Synthesis of non-steroidal anti-inflammatory drugs-derivatives of triflamide.
Scheme 31. Synthesis of non-steroidal anti-inflammatory drugs-derivatives of triflamide.
Molecules 27 05201 sch031
Scheme 32. Synthesis of trifluoromethanesulfonyl-based antagonists of the progesterone receptor.
Scheme 32. Synthesis of trifluoromethanesulfonyl-based antagonists of the progesterone receptor.
Molecules 27 05201 sch032
Scheme 33. Synthesis of N-glycosyl triflamide.
Scheme 33. Synthesis of N-glycosyl triflamide.
Molecules 27 05201 sch033
Scheme 34. Synthesis of trifluoromethanesulfonyl-based tubulin inhibitor.
Scheme 34. Synthesis of trifluoromethanesulfonyl-based tubulin inhibitor.
Molecules 27 05201 sch034
Scheme 35. Synthesis of triflamide-containing hURAT1 inhibitor.
Scheme 35. Synthesis of triflamide-containing hURAT1 inhibitor.
Molecules 27 05201 sch035
Scheme 36. Synthesis of triflamide-containing α-amylase enzyme inhibitor.
Scheme 36. Synthesis of triflamide-containing α-amylase enzyme inhibitor.
Molecules 27 05201 sch036
Scheme 37. Synthesis of triflamide-containing peptidomimetic.
Scheme 37. Synthesis of triflamide-containing peptidomimetic.
Molecules 27 05201 sch037
Scheme 38. Synthesis of triflamide-containing anti-retroviral agent.
Scheme 38. Synthesis of triflamide-containing anti-retroviral agent.
Molecules 27 05201 sch038
Scheme 39. Preparation of phenanthridine from N-(o-arylbenzyl) trifluoromethanesulfonamides with 1,3-diiodo-5,5-dimethylhydantoin.
Scheme 39. Preparation of phenanthridine from N-(o-arylbenzyl) trifluoromethanesulfonamides with 1,3-diiodo-5,5-dimethylhydantoin.
Molecules 27 05201 sch039
Scheme 40. Microwave-assisted synthesis of 2-vinylquinolines in the presence of TfNH2.
Scheme 40. Microwave-assisted synthesis of 2-vinylquinolines in the presence of TfNH2.
Molecules 27 05201 sch040
Scheme 41. Asymmetric Mannich-type reaction for the C−H Alkylation in the presence of TfNH-containing organyl catalyst.
Scheme 41. Asymmetric Mannich-type reaction for the C−H Alkylation in the presence of TfNH-containing organyl catalyst.
Molecules 27 05201 sch041
Scheme 42. C-H Alkylation of N-propyl substituted Triflamide.
Scheme 42. C-H Alkylation of N-propyl substituted Triflamide.
Molecules 27 05201 sch042
Scheme 43. Formation of N-Triflyl acetophenone derivative.
Scheme 43. Formation of N-Triflyl acetophenone derivative.
Molecules 27 05201 sch043
Scheme 44. One-pot synthesis of CF3-substituted vinyl trifluoromethanesulfonamides from imines and triflic anhydride.
Scheme 44. One-pot synthesis of CF3-substituted vinyl trifluoromethanesulfonamides from imines and triflic anhydride.
Molecules 27 05201 sch044
Scheme 45. Synthesis of tert-butyl((trifluoromethyl)sulfonyl)carbamate (BTFSA) and synthesis route for the incorporation of TFSA groups at the polyether block.
Scheme 45. Synthesis of tert-butyl((trifluoromethyl)sulfonyl)carbamate (BTFSA) and synthesis route for the incorporation of TFSA groups at the polyether block.
Molecules 27 05201 sch045
Scheme 46. Reaction of Phenethyltriflamides with 1,3-Dienes to dopaminergic 3-Benzazepines.
Scheme 46. Reaction of Phenethyltriflamides with 1,3-Dienes to dopaminergic 3-Benzazepines.
Molecules 27 05201 sch046
Scheme 47. Catalyzed oxidative annulation of ketoximes with sulfonamide.
Scheme 47. Catalyzed oxidative annulation of ketoximes with sulfonamide.
Molecules 27 05201 sch047
Scheme 48. NaClO-promoted amination of 3-substituted indoles with TfNH-containg amino acid derivatives.
Scheme 48. NaClO-promoted amination of 3-substituted indoles with TfNH-containg amino acid derivatives.
Molecules 27 05201 sch048
Scheme 49. Silver-catalyzed direct amination of unactivated C–H bonds.
Scheme 49. Silver-catalyzed direct amination of unactivated C–H bonds.
Molecules 27 05201 sch049
Scheme 50. Amination of styrenes with TfNH2 in the presence of NBS.
Scheme 50. Amination of styrenes with TfNH2 in the presence of NBS.
Molecules 27 05201 sch050
Scheme 51. Amination of vinylcyclohexane, hexa-1,5-diene and cyclohexa-1,4-diene with TfNH2 in the presence of NBS.
Scheme 51. Amination of vinylcyclohexane, hexa-1,5-diene and cyclohexa-1,4-diene with TfNH2 in the presence of NBS.
Molecules 27 05201 sch051
Scheme 52. Formation of amidines or acetamides from alkenes and TfNH2 in the presence of NBS.
Scheme 52. Formation of amidines or acetamides from alkenes and TfNH2 in the presence of NBS.
Molecules 27 05201 sch052
Scheme 53. Reaction of norbornene with TfNH2 in the presence of t-BuOCl + NaI in CH3CN.
Scheme 53. Reaction of norbornene with TfNH2 in the presence of t-BuOCl + NaI in CH3CN.
Molecules 27 05201 sch053
Scheme 54. Reaction of 1,5-hexadiene 188 with TfNH2 83 in the presence of t-BuOCl + NaI in CH3CN.
Scheme 54. Reaction of 1,5-hexadiene 188 with TfNH2 83 in the presence of t-BuOCl + NaI in CH3CN.
Molecules 27 05201 sch054
Scheme 55. Heterocyclization of 2,3-dimethylbuta-1,3-diene and 2,5-dimethylhexa-2,4-diene. with TfNH2 in the presence of t-BuOCl + NaI in CH3CN.
Scheme 55. Heterocyclization of 2,3-dimethylbuta-1,3-diene and 2,5-dimethylhexa-2,4-diene. with TfNH2 in the presence of t-BuOCl + NaI in CH3CN.
Molecules 27 05201 sch055
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Moskalik, M.Y.; Astakhova, V.V. Triflamides and Triflimides: Synthesis and Applications. Molecules 2022, 27, 5201. https://doi.org/10.3390/molecules27165201

AMA Style

Moskalik MY, Astakhova VV. Triflamides and Triflimides: Synthesis and Applications. Molecules. 2022; 27(16):5201. https://doi.org/10.3390/molecules27165201

Chicago/Turabian Style

Moskalik, Mikhail Y., and Vera V. Astakhova. 2022. "Triflamides and Triflimides: Synthesis and Applications" Molecules 27, no. 16: 5201. https://doi.org/10.3390/molecules27165201

Article Metrics

Back to TopTop