Next Article in Journal
Kinetic Characterization of Novel HIV-1 Entry Inhibitors: Discovery of a Relationship between Off-Rate and Potency
Next Article in Special Issue
Transition-Metal-Free Activation of Amide Bond by Arynes
Previous Article in Journal
Probing Protein-Protein Interactions Using Asymmetric Labeling and Carbonyl-Carbon Selective Heteronuclear NMR Spectroscopy
Previous Article in Special Issue
Asymmetric Primaquine and Halogenaniline Fumardiamides as Novel Biologically Active Michael Acceptors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Uses of N,N-Dimethylformamide and N,N-Dimethylacetamide as Reagents

Institut de Chimie Moléculaire de Reims, CNRS—Université de Reims Champagne-Ardenne, B.P. 1039, 51687 Reims CEDEX 2, France
*
Author to whom correspondence should be addressed.
Molecules 2018, 23(8), 1939; https://doi.org/10.3390/molecules23081939
Submission received: 13 July 2018 / Revised: 30 July 2018 / Accepted: 31 July 2018 / Published: 3 August 2018
(This article belongs to the Special Issue Amide Bond Activation)

Abstract

:
N,N-Dimethylformamide and N,N-dimethylacetamide are multipurpose reagents which deliver their own H, C, N and O atoms for the synthesis of a variety of compounds under a number of different experimental conditions. The review mainly highlights the corresponding literature published over the last years.

Graphical Abstract

1. Introduction

The organic, organometallic and bioorganic transformations are extensively carried out in N,N-dimethylformamide (DMF) or N,N-dimethylacetamide (DMAc). These two polar solvents are not only use for their dissolution properties, but also as multipurpose reagents. They participate in a number of processes and serve as a source of various building blocks giving one or more of their own atoms (Scheme 1).
In 2009, one of us reviewed the different roles of DMF, highlighting that DMF is much more than a solvent [1]. Subsequently, this topic has been documented by the teams of Jiao [2] and Sing [3]. For of a book devoted to solvents as reagents in organic synthesis, we wrote a chapter summarizing the reactions consuming DMF and DMAc as carbon, hydrogen, nitrogen and/or oxygen sources [4]. This book chapter tentatively covered the literature up to middle 2015. The present mini-review focuses on recent reactions which involve DM (DM = DMF or DMAc) as a reagent although some key older papers are also included for context. Processes which necessitate the prerequisite synthesis of DM derivatives such as the Vilsmeier-Haack reagents [5] and DMF dimethyl acetal [6] are not surveyed, but a few reactions of the present review involve the in-situ formation of a Vilsmeier-type intermediate (Vilsmeier-type reagents have been extensively used. Search on 26 June 2018 for “Vilsmeier” with SciFinder led to 4379 entries). Color equations, based on literature proposals, are used to easily visualize the DM atom origin. When uncertainty is expressed by the authors or suspected by us, the atom is typed in italic. Mechanistic schemes are not reported, but Scheme 2 [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35] summarizes different proposed reactions of DM with the corresponding literature references, where DM acts as either a nucleophilic or electrophilic reagent, or leads to neutral, ionic or radical species. The review is divided in Sections depending on the DM fragment(s) which is (are) incorporated into the reaction product.

2. C Fragment

Aerobic carbonylation under nickel/copper or palladium/silver synergistic catalysis occurred efficiently using the Me group of DMF as the C source, affording cyclic carbonylated compounds, via the directing group-assisted activation of a C(sp2)–H or C(sp3)-H bond (Equations (1) and (2) [36], Equations (3) and (4) [37]). Shifting from DMF to DMAc greatly decreased the yields (Equations (1) and (3)).
Molecules 23 01939 i001
Molecules 23 01939 i002
The Me group of DMF was also involved in the cyanation of the C(sp2)-H bond of arenes catalyzed with an heterogeneous copper catalyst (Equation (5)) [38].
Molecules 23 01939 i003

3. CH Fragment

Treatment of indole at 130 °C with suprastoichiometric amounts of CuI, t-BuOOH and AcOH in DMAc under air afforded the corresponding C3-formylation product (Equation (6)) [39]. Such a reaction also occurred with N-methylindole using CuI and CF3CO2H in DMAc under oxygen [40]. The CH fragment came from the NMe2 moiety [39,40]. In DMF, both procedures led to C3-cyanation (see below, Equation (25)).
Molecules 23 01939 i004
Cycloadditions leading to symmetrical tetrasubstituted pyridines using the Me group of DMF as the CH source have been carried out using either ketoxime carboxylates with Ru catalysis (Equation (7)) [41], or arones with both iodine and ammonium persulfate mediation (Equation (8)) [42].
Molecules 23 01939 i005
2,4-Diarylpyridines were also synthetized from Ru-catalyzed reaction of acetophenones with ammonium acetate and DMF as source of the N and CH atoms, respectively (Equation (9)) [43].
Molecules 23 01939 i006
With sodium azide as the nitrogen source, DMAc was superior to DMF to deliver the CH fragment of the copper-catalyzed domino reactions of aryl halides which led to imidazo [1,2-c]quinazolines, quinazolinones or imidazo[4,5-c]quinolones (Equations (10)–(12)) [44].
Molecules 23 01939 i007
In contrast to the above examples, the Cu-catalyzed cyclization leading to 6H-chromeno[4,3-b]quinolin-6-ones (Equation (13)) with incorporation of a CH belonging to DMF occurred with low yield when t-BuOOH was the oxidant. Shifting to t-butyl perbenzoate allowed an effective reaction [45]. For the synthesis of 4-acyl-1,2,3-triazoles from Cu-catalyzed cycloaddition to acetophenones (Equation (14)), K2S2O8 was superior to t-BuOOH, (t-BuO)2 and (PhCO2)2 [46]. Yields decreased with DMAc instead of DMF (Equations (13) and (14)).
Molecules 23 01939 i008
The insertion of a CH from the NMe2 of DM under metal-free conditions has been reported for the synthesis of cyclic compounds such as:
-
pyrimidines from t-BuOOH-mediated reaction between acetophenones, amidines and DMF (Equation (15)) [47],
Molecules 23 01939 i009
-
substituted phenols also from three components cycloadditions (Equation (16)) [48],
Molecules 23 01939 i010
-
3-acylindoles from 2-alkenylanilines (Equation (17)) [49],
Molecules 23 01939 i011
-
benzimidazoles and benzothiazole from o-phenylenediamine or 2-aminobenzenethiol through carbon dioxide-mediated cyclization (Equation (18)) [50].
Molecules 23 01939 i012

4. CH2 Fragment

The coupling of indoles or imidazo[1,2-a]pyridines to afford heterodiarylmethanes with DMF as the methylenating reagent occurred in fair to high yields with a CuI catalyst associated to t-BuOOH [51] or K2S2O8 [52] (Equations (19)–(21)). Use of DMAc was less efficient.
Molecules 23 01939 i013
In DMAc, the I2/t-BuOOH association catalyzed the formation of methylene-bridged bis-1,3-dicarbonyl compounds from aryl β-ketoesters or β-ketoamides (Equation (22)). Lower yields were obtained with I2/K2S2O8 in DMAc or DMF [9]. Subjection 1,3-diphenylpropane-1,3-dione or ethyl 3-oxobutanoate to the I2/t-BuOOH/DMAc did not afford the bridged compounds.
Molecules 23 01939 i014
A Mannich reaction leading to β-amino ketones with DMF as the formaldehyde source has been reported in the presence of t-BuOOH and catalytic amounts of an N-heterocyclic carbene, SnCl2 and NEt3 (Equation (23)) [11].
Molecules 23 01939 i015
The study of an unexpected reaction due to the oxidation of DMAc with aqueous t-BuOOH (Equation (24)) showed the formation of MeCONMe(CH2OH) and MeCON(CH2OH)2, these unusual species delivering the methylene group [12].
Molecules 23 01939 i016

5. NC Fragment

While CuI under oxidative and acidic conditions led, in DMAc, to the C3-formylation of indole and N-methylindole (Equation (6)), reactions in DMF led to C3-cyanations (Equation (25)) [39,40]. Cyanation of electron-rich arenes and benzaldehydes was also carried out (Equation (26)) [39]. Monitoring the course of the reaction indicated a cyanation arising via the formyl compounds [39,40]. Moreover, 3-iodo indole could also be involved in the formation of the cyano product [40].
Molecules 23 01939 i017
Laser ablation of silver nitrate in DMF led to silver cyanide (Equation (27)) [53].
Molecules 23 01939 i018

6. NMe2 Fragment

This chapter is divided in sections corresponding to the type of function reacting with DM.

6.1. Aryl Halides

Refluxing chloropyridines in DM or DMAc afforded the corresponding aminopyridines (Equation (28)) [25]. The amination of aryl chlorides and 3-pyridinyl chloride with DMF occurred at room temperature in the presence of potassium t-butoxide and a carbenic palladium catalyst (Equation (29)) [27].
Molecules 23 01939 i019

6.2. Alkylarenes

Oxidation of methylarenes and ethylarenes at 80 °C in DMF using catalytic amounts of both I2 and NaOH [31] or n-Bu4NI [32] associated to aqueous t-BuOOH under air led to benzylic oxidation and incorporation of the NMe2 fragment to afford benzamides (Equation (30)) or α-ketoamides (Equation (31)).
Molecules 23 01939 i020

6.3. Alkenes

Hydrocarbonylation of terminal alkenes and norbornene followed by acyl metathesis with DM occurred under Pd catalysis, CO pressure and in the presence of ammonium chloride or N-methyl-2-pyrrolidone hydrochloride (NMP·HCl) (Equations (32) and (33)) [54]. From alkenes, the selectivity towards linear and branched products depended on the catalytic system (Equation (32)). DMF and DMAc afforded similar results.
Molecules 23 01939 i021

6.4. Acids

Copper, palladium and ruthenium catalysts associated to oxidants and DMF were used for the amidation of cinnamic acids [29] and carboxylic acids [55] (Equations (34) and (35)). N,N-Dimethylbenzamide was one of the products obtained from the CuII-catalyzed oxidation of flavonol [56].
Molecules 23 01939 i022
Metal-free conditions and DMF were used for:
-
the amidation of acids promoted with propylphosphonic anhydride associated to HCl at 130 °C (Equation (36)) [15],
-
the amination of acids employing a hypervalent iodine reagent at room temperature (Equation (37)) [35]. Mesityliodine diacetate was superior to the other hypervalent iodine reagents, while oxidants such as I2, t-BuOOH, NaIO4 or K2S2O8 did not mediate the amidation reaction [35].
Molecules 23 01939 i023
Treatment of arylacetic and cinnamic acids with base, sulfur and DMF at 100–120 °C led to decarboxylative thioamidation (Equations (38) and (39)) [7]. Inhibition of the process in the presence of TEMPO or BHT indicated a radical involvement in the transformations.
Molecules 23 01939 i024

6.5. Carbonylated Compounds

Reaction of 2-arylquinazolin-4(3H)-ones with TsCl and t-BuOK in DM provided the corresponding 4-(dimethylamino)quinazolines in good yields, especially in DMF (Equation (40)). These reactions occurred via the formation of the 2-aryl-4-(tosyloxy)quinazolines [57].
Molecules 23 01939 i025
Various amides have been synthetized from aldehydes and DMF using t-BuOOH and a recyclable heterogeneous catalyst—a carbon–nitrogen embedded cobalt nanoparticle denoted as Co@C-N600 (Equation (41)) [33]. The same transformation of benzaldehydes was subsequently reported using Co/Al hydrotalcite-derived catalysts [58].
Molecules 23 01939 i026
Copper oxide and iodine mediated the reaction of acetophenones with sulfur and DMF to afford α-arylketothioamides (Equation (42)) via the formation of α-iodoacetophenones [59].
Molecules 23 01939 i027
Elemental sulfur and the NMe2 moiety of DMF or DMAc was also used for the DBU-promoted synthesis of thioamides from aldehydes (Equation (43)) or arones (Equation (44)) [60], the latter undergoing an efficient Willgerodt-Kindler reaction [61,62].
Molecules 23 01939 i028

6.6. Benzyl Amines

The recyclable Co/Al catalysts used above in DMF for the amidation of benzaldehydes also led to benzamides from benzylamines and t-BuOOH (Equation (45)). These transformations would involve benzaldehydes as intermediates [58].
Molecules 23 01939 i029

6.7. Nitriles

NaOH mediated, at room temperature, the efficient reaction of the CN group of 4-oxo-2,4-diphenylbutanenitrile with DMF to afford the corresponding γ-ketoamide (Equation (46)) [63]. Such compounds were also obtained from the domino reaction of chalcones with malononitrile and NaOH in DMF [63].
Molecules 23 01939 i030

6.8. Sulfur Compounds

Sulfonamides were synthetized:
-
from thiophenols, DMF and air via an oxygen-activated radical process mediated by copper salts and cinnamic acid (Equation (47)) [64],
Molecules 23 01939 i031
or reaction of sodium sulfonates with N-iodosuccinimide and DMF pretreated with t-BuOK (Equation (48)) via, probably, sulfonyl iodides (Equation (49)) [65].
Molecules 23 01939 i032

7. O Fragment

DMF delivered its oxygen atom to 1,2-cyclic sulfamidates via nucleophilic displacement at the quaternary center to afford, after hydrolysis, an aminoalcohol (Equation (50)) [17].
Molecules 23 01939 i033
DMF was also the oxygen source leading to an imidazolinone from the reaction with the Cu-carbene complex and the borate salt depicted in Equation (51) [66].
Molecules 23 01939 i034
The I2/CuO association allowed the α-hydroxylation of arones in abstracting, via the α-iodoarone, the oxygen atom of DMF (Equation (52)) [18].
Molecules 23 01939 i035

8. C=O Fragment

With DMF as the CO surrogate, quinazolinones have been prepared at 140–150 °C
-
via C(sp2)-H bond activation and annulation using Pd/C [67] or Pd(OAc)2 [8], in the presence of K2S2O8, CF3CO2H and O2 (Equation (53)),
Molecules 23 01939 i036
or carbon dioxide-mediated cyclization of 2-aminobenzonitrile (Equation (54)) [50]. This latter reaction would involve a Vilsmeier-Haack type intermediate and did not occur with DMAc.
Molecules 23 01939 i037
A carbonylative Suzuki-type reaction leading to diarylketones arose in DMF under Ni catalysis at 100 °C (Equation (55)), IPr = bis(2,6-diisopropylphenyl)imidazol-2-ylidene) [13], or Pd catalysis and UV light assistance at room temperature (Equation (56)) [68].
Molecules 23 01939 i038
Catalytic amounts of a ruthenium pincer complex and t-BuOK led, at 165 °C, to symmetric and unsymmetric N,N’-disubstituted ureas from primary amines and DMF (Equation (57)) [69].
Molecules 23 01939 i039
At 120 °C under CuBr2 catalysis, o-iodoanilines reacted with potassium sulfide and DMF, leading to benzothiazolones (Equation (58)) [70].
Molecules 23 01939 i040

9. C=ONMe2 Fragment

Potassium persulfate-promoted the reaction of pyridines with DMF to provide N,N-dimethylpicolinamides (Equation (59)) [71], while the oxidative carbamoylation of isoquinoline N-oxides, with also DMF, was catalyzed by PdII in the presence of ytterbium oxide as base and tetrabutylammonium acetate, the latter mediating the N-O reduction (Equation (60)) [72].
Molecules 23 01939 i041
Alkynylation of DMF leading to N,N-dimethylamides was produced with peroxides and either an hypervalent alkynyl iodide under Ag catalysis (Equation (61)) [73], or a terminal alkyne under Cu catalysis (Equation (62)) [34].
Molecules 23 01939 i042
A peroxide was also used for the carbamoylation of 4-arylcoumarin with DMF (Equation (63)) [74].
Molecules 23 01939 i043
α-Ketoamides were obtained from the domino reaction of toluenes with DMF using (t-BuO)2, Cs2CO3 and catalytic amounts of n-Bu4NI (Equation (64)) [75].
Molecules 23 01939 i044
At 100 °C in DMF, Cu catalyst associated to t-BuOOH led to unsymmetrical ureas from 2-oxindoles (Equation (65)) [76]. The peroxide would mediate the cleavage reaction, and was the oxygen source of the benzylic carbonyl. That resulted in a ketoamine which undergone the Cu-catalyzed reaction with DMF/t-BuOOH, leading to the urea.
Molecules 23 01939 i045

10. H Fragment

Semihydrogenation of diaryl alkynes occurred under Ru (Equation (66)) [77] and Pd [78] catalysis with DMF and water as hydrogen source.
Molecules 23 01939 i046
Cobalt porphyrins catalyzed the hydrogenation transfer from DMF to the C(sp3)-C(sp3) bond of [2.2]paracyclophane (Equation (67)) [79]. DMF was also involved in the Ni-catalyzed intramolecular hydroarylations depicted in Equation (68) [14].
Molecules 23 01939 i047

11. RC Fragment

New metal-catalyzed and metal-free conditions involving the CH of the formyl group of DMF have been reported for cyclizations leading to heterocycles (Equations (69) [80,81,82] and (70) [83]).
Molecules 23 01939 i048
2-Methylbenzimidazoles were obtained from PhSiH3-assisted delivery of the CMe of DMAc to benzene-1,2-diamines (Equation (71)) [82].
Molecules 23 01939 i049
Addition of p-tolyllithium to DMF followed by reaction with hydroxylamine hydrochloride afforded 4-methylbenzaldehyde oxime (Equation (72)). The latter underwent cycloaddition with diphenylphosphoryl azide or, in the presence of Oxone®, with diethylacetylene dicarboxylate to provide the corresponding 5-aryltetrazole or 3-arylisoxazole, respectively [84].
Molecules 23 01939 i050

12. RCNMe2 Fragment

Dihydropyrrolizino[3,2-b]indol-10-ones were isolated in fair to high yields from a Cs2CO3-promoted domino reaction leading to the formation of three bonds with incorporation of the HCNMe of DMF. Such a reaction-type with incorporation of the MeCNMe also occurred in DMAc but with a low yield (Equation (73)) [26].
Molecules 23 01939 i051

13. RC-O Fragment

A 1:1 mixture of CuO and I2 led the α-formyloxylation or α-acetoxylation of methylketones by DMF or DMAc, respectively (Equation (74)). α-Iodoketones would be the intermediates as indicated by the reaction of 2-iodo-1-(4-methoxyphenyl)ethanone (Equation (75)). Traces of water delivered the carbonyl oxygen [16].
Molecules 23 01939 i052
Stereoinversion of the secondary alcohols of a number of carbocyclic substrates was carried out via their triflylation followed by treatment with aqueous DMF (Equation (76)) and subsequent methanolysis [85]. A one pot stereoinversion process was reported.
Molecules 23 01939 i053
The formyl group of DMF was involved in the triflic anhydride-mediated domino reaction depicted in Equation (77) [86].
Molecules 23 01939 i054
Chloroformyloxylation and chloroacetoxylation of olefinic substrates were performed with PhICl2 and either wet DMF or DMAc (Equation (78)) [87]. Styrenes suffered also difunctionalization using aryl diazonium salts and Ru photocatalysis in wet DM (Equation (79)) [28].
Molecules 23 01939 i055
Molecules 23 01939 i056

14. RC=O Fragment

Metal-catalyzed or CO2-mediated C–N and N–H bond metathesis reactions between primary or secondary amines and DM provided the transamidation products, that is formamides or acetamides (Equation (80)) [88,89,90,91].
Molecules 23 01939 i057
Formylation of aromatic substrates resulted from their treatment with LDA [92] or n-BuLi [93] and, subsequently, DMF (Equation (81)).
Molecules 23 01939 i058
Graphene oxide reacted with DBU and DM to afford, in presence of trace of water, N-(3-(2-oxoazepan-1-yl)propyl)formamide or the corresponding acetamide (Equation (82)) [94].
Molecules 23 01939 i059

15. RC=ON(CH2)Me Fragment

N-Amidoalkylation of imidazoles and 1,2,3-triazoles with DM effectively arose under various experimental conditions (Equations (83) [95] and (84) [10]).
Molecules 23 01939 i060

16. HC-ONMe2 Fragment

Catalysis with Mo(CO)6 of the reduction of DMF with triethylsilane afforded a siloxymethylamine (Equation (85)), which was used as a Mannich reagent [96].
Molecules 23 01939 i061

17. RC and O Fragment

Arynes, which are easily obtained from, for example 2-(trimethylsilyl)phenyl trifluoromethane-sulfonate, undergone a [2 + 2] cyclization with DM giving a benzoxetene and its isomer, the ortho-quinone methide (Scheme 3). Trapping of these intermediates provides various products, which contain the formyl or acetyl CH part and the O atom of DM (Equations (86) [19], (87) [20], (88) [21] and (89) [97]), or the HCNMe2 and O fragments of DMF (see Section 18).
Molecules 23 01939 i062
Molecules 23 01939 i063
At 115 °C in wet toluene, the hexadehydro-Diels–Alder of tetraynes depicted in Equation (89) was in-situ followed by [2 + 2] cycloaddition reaction with DM leading to multifunctionalized salicylaldehydes and salicylketones [97].
Molecules 23 01939 i064

18. RCNMe2 and O Fragment

All atoms of DMF were inserted in the polyfunctionalized compounds produced from domino reactions involving formation of arynes, cycloaddition with DMF and subsequent trapping with α-chloro β-diesters (Equation (90)) [22], aroyl cyanides (Equation (91)) [23] or diesters of acetylenedicarboxylic acid (Equation (92)) [24].
Molecules 23 01939 i065

19. HC=O and HC Fragment

Lithium thioanisole biscarbanion reacted with two molecules of DMF to afford benzo[b]thiophene-2-carbaldehyde (Equation (93)) [98].
Molecules 23 01939 i066

20. H and NMe2 Fragment

The reaction of DMF with sodium and subsequent addition of terminal activated alkynes afforded the corresponding hydroamination compounds (Equation (94)) [99].
Molecules 23 01939 i067

21. H and C=ONMe2 Fragment

Semicarbazides have been synthetized from additions, mediated with (t-BuO)2 and catalytic amounts of both NaI and PhCOCl, of the H and CONMe2 moieties of DMF to the extremities of the N=N bond of azoarenes (Equation (95)) [100]. The role of NaI and PhCOCl is not clear and, furthermore, exchange of NaI for imidazole led to formylhydrazines (Equation (96)) [100]. The corresponding acetylhydrazine was not formed in DMAc (Equation (96)). The (t-BuO)2/NaI/PhCOCl/DMF system led to the addition of H and CONMe2 to the N=C bond of N-benzylideneaniline but with low yield (Equation (97)) [100].
Molecules 23 01939 i068
The Ru-catalyzed hydrocarbamoylative cyclization of 1,6-diynes proceeded in DMF to afford cyclic α,β,δ,γ-unsaturated amides (Equation (98)) [101,102,103].
Molecules 23 01939 i069

22. H, C=ONMe2 and NMe2 Fragment

Re2(CO)8[μ-η2-C(H)=C(H)Bu](μ-H) undergone reaction with DMF leading to hexenyl/CONMe2 and CO/HNMe2 exchange of ligands (Equation (99)) [104].
Molecules 23 01939 i070

23. C=ONMe2 and CH Fragment

Couplings between amidines, styrenes and fragments of two molecules of DMF in the presence of t-BuOOH and a PdII catalyst provided pyrimidine carboxamides (Equation (100)) [105]. DMAc may also be the CH source as exemplified with the formation of the N,N-diethyl-2,4-diphenylpyrimidine-5-carboxamide when N,N-diethylformamide was the source of the amide moiety (Equation (101)).
Molecules 23 01939 i071

24. Reducing or Stabilizing Agent

DMF is a powerful reducing agent of metal salts, hence its use for the preparation of metal colloids [106]. In wet DMF, PdCl2 led to carbamic acid and Pd(0) nanoparticles (Equation (102)) [107]. The latter have been associated with the metal-organic framework Cu2(BDC)2(DABCO) (BDC = 1,4-benzenedicarboxylate), leading to a catalytic system with high activity and recyclability for the aerobic oxidation of benzyl alcohols to aldehydes [108] and Suzuki-Miyaura cross-coupling reactions [107].
Molecules 23 01939 i072
In addition, DMF can act as stabilizing agent of metal colloids to afford effective and recyclable catalysts, based for examples:
-
on iron for the hydrosilylation of alkenes (Equation (103)) [109],
Molecules 23 01939 i073
-
on palladium for the synthesis of 2,3-disubstituted indoles from 2-halooanilines and alkynes (Equation (104)) [110],
Molecules 23 01939 i074
-
on copper for Sonogashira–Hagihara cross-coupling reactions (Equation (105)) [111],
Molecules 23 01939 i075
-
on iridium for methylation of alcohols (Equation (106)) and amines (Equation (107)), using methanol as the C1 source [112].
Molecules 23 01939 i076
Thermal decomposition of DMF leads to CO, which reacts with water under CuFe2O4 catalysis to produce hydrogen [113]. In the presence of 2-nitroanilines, this water gas shift reaction was part of a domino reaction involving the reduction of the nitro group followed by cyclisation into benzimidazoles using a CH from the NMe2 of DMF (Equation (108)) [113]. Such cyclisation is above documented under different experimental conditions (Equation (18)) [50].
Molecules 23 01939 i077

25. Conclusions

This minireview highlights recent uses of DMF and DMAc as sources of building blocks in various reactions of the organic synthesis. We assume that new uses of these multipurpose reagents will be reported.

Author Contributions

J. Muzart collected the literature references. J. Muzart and J. Le Bras contributed to the manuscript writing.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Muzart, J. N,N-Dimethylformamide: Much more than a solvent. Tetrahedron 2009, 65, 8313–8323. [Google Scholar] [CrossRef]
  2. Ding, S.; Jiao, N. N,N-Dimethylformamide: A multipurpose building block. Angew. Chem. Int. Ed. 2012, 51, 9226–9231. [Google Scholar] [CrossRef] [PubMed]
  3. Batra, A.; Singh, P.; Singh, K.N. Cross dehydrogenative coupling (CDC) reactions of N,N-disubstituted formamides, benzaldehydes and cycloalkanes. Eur. J. Org. Chem. 2016, 2016, 4927–4947. [Google Scholar] [CrossRef]
  4. Le Bras, J.; Muzart, J. N,N-Dimethylformamide and N,N-dimethylacetamide as carbon, hydrogen, nitrogen and/or oxygen sources. In Solvents as Reagents in Organic Synthesis, Reactions and Applications; Wu, X.-F., Ed.; Wiley-VCH: Weinheim, Germany, 2017; pp. 199–314. [Google Scholar]
  5. Kolosov, M.A.; Shvets, E.H.; Manuenkov, D.A.; Vlasenko, S.A.; Omelchenko, I.V.; Shishkina, S.V.; Orlov, V.D. A synthesis of 6-functionalized 4,7-dihydro[1,2,4]triazolo[1,5-a]pyrimidines. Tetrahedron Lett. 2017, 58, 1207–1210. [Google Scholar] [CrossRef]
  6. Borah, A.; Goswami, L.; Neog, K.; Gogoi, P. DMF dimethyl acetal as carbon source for α-methylation of ketones: A hydrogenation-hydrogenolysis strategy of enaminones. J. Org. Chem. 2015, 80, 4722–4728. [Google Scholar] [CrossRef] [PubMed]
  7. Kumar, S.; Vanjari, R.; Guntreddi, T.; Singh, K.N. Sulfur promoted decarboxylative thioamidation of carboxylic acids using formamides as amine proxy. Tetrahedron 2016, 72, 2012–2017. [Google Scholar] [CrossRef]
  8. Chen, J.; Feng, J.-B.; Natte, K.; Wu, X.-F. Palladium-catalyzed carbonylative cyclization of arenes by C-H bond activation with DMF as the carbonyl source. Chem. Eur. J. 2015, 21, 16370–16373. [Google Scholar] [CrossRef] [PubMed]
  9. Guo, S.; Zhu, Z.; Lu, L.; Zhang, W.; Gong, J.; Cai, H. Metal-free Csp3-N bond cleavage of amides using tert-butyl hydroperoxide as oxidant. Synlett 2015, 26, 543–546. [Google Scholar] [CrossRef]
  10. Zhu, Z.; Wang, Y.; Yang, M.; Huang, L.; Gong, J.; Guo, S.; Cai, H. A metal-free cross-dehydrogenative coupling reaction of amides to access N-alkylazoles. Synlett 2016, 27, 2705–2708. [Google Scholar]
  11. Alanthadka, A.; Devi, E.S.; Selvi, A.T.; Nagarajan, S.; Sridharan, V.; Maheswari, C.U. N-Heterocyclic carbene-catalyzed Mannich reaction for the synthesis of β-amino ketones: N,N-dimethylformamide as carbon source. Adv. Synth. Catal. 2017, 359, 2369–2374. [Google Scholar] [CrossRef]
  12. Zewge, D.; Bu, X.; Sheng, H.; Liu, Y.; Liu, Z.; Harman, B.; Reibarkh, M.; Gong, X. Mechanistic insight into oxidized N,N-dimethylacetamide as a source of formaldehyde related derivatives. React. Chem. Eng. 2018, 3, 146–150. [Google Scholar] [CrossRef]
  13. Li, Y.; Tu, D.-H.; Wang, B.; Lu, J.-Y.; Wang, Y.-Y.; Liu, Z.-T.; Liu, Z.-W.; Lu, J. Nickel-catalyzed carbonylation of arylboronic acids with DMF as CO source. Org. Chem. Front. 2017, 4, 569–572. [Google Scholar] [CrossRef]
  14. Lu, K.; Han, X.-W.; Yao, W.-W.; Luan, Y.-X.; Wang, Y.-X.; Chen, H.; Xu, X.-T.; Zhang, K.; Ye, M. DMF-promoted redox-neutral Ni-catalyzed intramolecular hydroarylation of alkene with simple arene. ACS Catal. 2018, 8, 3913–3917. [Google Scholar] [CrossRef]
  15. Bannwart, L.; Abele, S.; Tortoioli, S. Metal-free amidation of acids with formamides and T3P®. Synthesis 2016, 48, 2069–2078. [Google Scholar]
  16. Zhu, Y.-P.; Gao, Q.-H.; Lian, M.; Yuan, J.-J.; Liu, M.-C.; Zhao, Q.; Yang, Y.; Wu, A.-X. A sustainable byproduct catalyzed domino strategy: Facile synthesis of α-formyloxy and acetoxy ketones via iodination/nucleophilic substitution/hydrolyzation/oxidation sequences. Chem. Commun. 2011, 47, 12700–12702. [Google Scholar] [CrossRef] [PubMed]
  17. Mata, L.; Avenoza, A.; Busto, J.H.; Peregrina, J.M. Chemoselectivity control in the reactions of 1,2-cyclic sulfamidates with amines. Chem. Eur. J. 2013, 19, 6831–6839. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, W.; Chen, C.; Zhou, P. N,N-Dimethylformamide (DMF) as a source of oxygen to access α-hydroxy arones via the α-hydroxylation of arones. J. Org. Chem. 2017, 82, 2219–2222. [Google Scholar] [CrossRef] [PubMed]
  19. Liu, F.; Yang, H.; Hu, X.; Jiang, G. Metal-free synthesis of ortho-CHO diaryl ethers by a three-component sequential coupling. Org. Lett. 2014, 16, 6408–6411. [Google Scholar] [CrossRef] [PubMed]
  20. Yoshioka, E.; Kohtani, S.; Miyabe, H. Three-component coupling reactions of arynes for the synthesis of benzofurans and coumarins. Molecules 2014, 19, 863–880. [Google Scholar] [CrossRef] [PubMed]
  21. Yoshioka, E.; Kohtani, S.; Miyabe, H. 2,3,4,9-Tetrahydro-9-(3-hydroxy-1,4-dioxo-1H-dihydronaphthalen-2-yl)-8-methoxy-3,3-dimethyl-1H-xanthen-1-one. Molbank 2015, 2015, M841. [Google Scholar] [CrossRef]
  22. Yoshioka, E.; Tanaka, H.; Kohtani, S.; Miyabe, H. Straightforward synthesis of dihydrobenzofurans and benzofurans from arynes. Org. Lett. 2013, 15, 3938–3941. [Google Scholar] [CrossRef] [PubMed]
  23. Zhou, C.; Wang, J.; Jin, J.; Lu, P.; Wang, Y. Three-component synthesis of α-amino-α-aryl carbonitriles from arynes, aroyl cyanides, and N,N-dimethylformamide. Eur. J. Org. Chem. 2014, 2014, 1832–1835. [Google Scholar] [CrossRef]
  24. Yoshioka, E.; Tamenaga, H.; Miyabe, H. [4+2] Cycloaddition of intermediates generated from arynes and DMF. Tetrahedron Lett. 2014, 55, 1402–1405. [Google Scholar] [CrossRef]
  25. Kodimuthali, A.; Mungara, A.; Prasunamba, P.L.; Pal, M. A simple synthesis of aminopyridines: Use of amides as amine source. J. Braz. Chem. Soc. 2010, 21, 1439–1445. [Google Scholar] [CrossRef]
  26. Zhang, Q.; Song, C.; Huang, H.; Zhang, K.; Chang, J. Cesium carbonate promoted cascade reaction involving DMF as a reactant for the synthesis of dihydropyrrolizino[3,2-b]indol-10-ones. Org. Chem. Front. 2018, 5, 80–87. [Google Scholar] [CrossRef]
  27. Chen, W.-X.; Shao, L.-X. N-Heterocyclic carbene–palladium(II)-1-methylimidazole complex catalyzed amination between aryl chlorides and amides. J. Org. Chem. 2012, 77, 9236–9239. [Google Scholar] [CrossRef] [PubMed]
  28. Yao, C.-J.; Sun, Q.; Rastogi, N.; Koenig, B. Intermolecular formyloxyarylation of alkenes by photoredox Meerwein reaction. ACS Catal. 2015, 5, 2935–2938. [Google Scholar] [CrossRef]
  29. Yan, H.; Yang, H.; Lu, L.; Liu, D.; Rong, G.; Mao, J. Copper-catalyzed synthesis of α,β-unsaturated acylamides via direct amidation from cinnamic acids and N-substituted formamides. Tetrahedron 2013, 69, 7258–7263. [Google Scholar] [CrossRef]
  30. Salamone, M.; Milan, M.; DiLabio, G.A.; Bietti, M. Reactions of the cumyloxyl and benzyloxyl radicals with tertiary amides. Hydrogen abstraction selectivity and the role of specific substrate-radical hydrogen bonding. J. Org. Chem. 2013, 78, 5909–5917. [Google Scholar] [CrossRef] [PubMed]
  31. Du, B.; Sun, P. Syntheses of amides via iodine-catalyzed multiple sp3 C-H bonds oxidation of methylarenes and sequential coupling with N,N-dialkylformamides. Sci. China Chem. 2014, 57, 1176–1182. [Google Scholar] [CrossRef]
  32. Du, B.; Jin, B.; Sun, P. The syntheses of α-ketoamides via n-Bu4NI-catalyzed multiple sp3C-H bond oxidation of ethylarenes and sequential coupling with dialkylformamides. Org. Biomol. Chem. 2014, 12, 4586–4589. [Google Scholar] [CrossRef] [PubMed]
  33. Bai, C.; Yao, X.; Li, Y. Easy access to amides through aldehydic C-H bond functionalization catalyzed by heterogeneous Co-based catalysts. ACS Catal. 2015, 5, 884–891. [Google Scholar] [CrossRef]
  34. Wu, J.-J.; Li, Y.; Zhou, H.-Y.; Wen, A.-H.; Lun, C.-C.; Yao, S.-Y.; Ke, Z.; Ye, B.-H. Copper-catalyzed carbamoylation of terminal alkynes with formamides via cross-dehydrogenative coupling. ACS Catal. 2016, 6, 1263–1267. [Google Scholar] [CrossRef]
  35. Zhang, C.; Yue, Q.; Xiao, Z.; Wang, X.; Zhang, Q.; Li, D. Synthesis of O-aroyl-N,N-dimethylhydroxylamines through hypervalent iodine-mediated amination of carboxylic acids with N,N-dimethylformamide. Synthesis 2017, 49, 4303–4308. [Google Scholar]
  36. Wu, X.; Zhao, Y.; Ge, H. Direct aerobic carbonylation of C(sp2)-H and C(sp3)-H bonds through Ni/Cu synergistic catalysis with DMF as the carbonyl source. J. Am. Chem. Soc. 2015, 137, 4924–4927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Rao, D.N.; Rasheed, S.; Das, P. Palladium/silver synergistic catalysis in direct aerobic carbonylation of C(sp2)–H bonds using DMF as a carbon source: Synthesis of pyrido-fused quinazolinones and phenanthridinones. Org. Lett. 2016, 18, 3142–3145. [Google Scholar]
  38. Boosa, V.; Bilakanti, V.; Gutta, N.; Velisoju, V.K.; Medak, S.; Ramineni, K.; Jorge, B.; Muxina, K.; Akula, V. C–H bond cyanation of arenes using N,N-dimethylformamide and NH4HCO3 as a CN source over a hydroxyapatite supported copper catalyst. Catal. Sci. Technol. 2016, 6, 8055–8062. [Google Scholar]
  39. Zhang, L.; Lu, P.; Wang, Y. Copper-mediated cyanation of indoles and electron-rich arenes using DMF as a single surrogate. Org. Biomol. Chem. 2015, 13, 8322–8329, Correction in Org. Biomol. Chem. 2016, 14, 1840. [Google Scholar] [CrossRef] [PubMed]
  40. Xiao, J.; Li, Q.; Chen, T.; Han, L.-B. Copper-mediated selective aerobic oxidative C3-cyanation of indoles with DMF. Tetrahedron Lett. 2015, 56, 5937–5940. [Google Scholar] [CrossRef]
  41. Zhao, M.-N.; Hui, R.-R.; Ren, Z.-H.; Wang, Y.-Y.; Guan, Z.-H. Ruthenium-catalyzed cyclization of ketoxime acetates with DMF for synthesis of symmetrical pyridines. Org. Lett. 2014, 16, 3082–3085. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, W.; Tan, H.; Chen, C.; Pan, Y. A method to access symmetrical tetrasubstituted pyridines via iodine and ammonium persulfate mediated [2+2+1+1]-cycloaddition reaction. Adv. Synth. Catal. 2017, 359, 1594–1598. [Google Scholar] [CrossRef]
  43. Bai, Y.; Tang, L.; Huang, H.; Deng, G.-J. Synthesis of 2,4-diarylsubstituted-pyridines through a Ru-catalyzed four component reaction. Org. Biomol. Chem. 2015, 13, 4404–4407. [Google Scholar] [CrossRef] [PubMed]
  44. Nandwana, N.K.; Dhiman, S.; Saini, H.K.; Kumar, I.; Kumar, A. Synthesis of quinazolinones, imidazo[1,2-c]quinazolines and imidazo[4,5-c]quinolines through tandem reductive amination of aryl halides and oxidative amination of C(sp3)–H bonds. Eur. J. Org. Chem. 2017, 2017, 514–522. [Google Scholar] [CrossRef]
  45. Weng, Y.; Zhou, H.; Sun, C.; Xie, Y.; Su, W. Copper-catalyzed cyclization for access to 6H-chromeno[4,3-b]quinolin-6-ones employing DMF as the carbon source. J. Org. Chem. 2017, 82, 9047–9053. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, Y.; Nie, G.; Zhou, Z.; Jia, L.; Chen, Y. Copper-catalyzed oxidative cross-dehydrogenative coupling/oxidative cycloaddition: Synthesis of 4-acyl-1,2,3-triazoles. J. Org. Chem. 2017, 82, 9198–9203. [Google Scholar] [CrossRef] [PubMed]
  47. Zheng, L.-Y.; Guo, W.; Fan, X.-L. Metal-free, TBHP-mediated, [3+2+1]-type intermolecular cycloaddition reaction: Synthesis of pyrimidines from amidines, ketones, and DMF through C(sp3)–H activation. Asian J. Org. Chem. 2017, 6, 837–840. [Google Scholar] [CrossRef]
  48. Wang, F.; Wu, W.; Xu, X.; Shao, X.; Li, Z. Synthesis of substituted phenols via 1,1-dichloro-2-nitroethene promoted condensation of carbonyl compounds with DMF. Tetrahedron Lett. 2018, 59, 2506–2510. [Google Scholar] [CrossRef]
  49. Wang, J.-B.; Li, Y.-L.; Deng, J. Metal-free activation of DMF by dioxygen: A cascade multiple-bond-formation reaction to synthesize 3-acylindoles from 2-alkenylanilines. Adv. Synth. Catal. 2017, 359, 3460–3467. [Google Scholar] [CrossRef]
  50. Rasal, K.B.; Yadav, G.D. Carbon dioxide mediated novel synthesis of quinazoline-2,4(1H,3H)-dione in water. Org. Process Res. Dev. 2016, 20, 2067–2073. [Google Scholar] [CrossRef]
  51. Pu, F.; Li, Y.; Song, Y.-H.; Xiao, J.; Liu, Z.-W.; Wang, C.; Liu, Z.-T.; Chen, J.-G.; Lu, J. Copper-catalyzed coupling of indoles with dimethylformamide as a methylenating reagent. Adv. Synth. Catal. 2016, 358, 539–542. [Google Scholar] [CrossRef]
  52. Mondal, S.; Samanta, S.; Santra, S.; Bagdi, A.K.; Hajra, A. N,N-Dimethylformamide as a methylenating reagent: Synthesis of heterodiarylmethanes via copper-catalyzed coupling between imidazo[1,2-a]pyridines and indoles/N,N-dimethylaniline. Adv. Synth. Catal. 2016, 358, 3633–3641. [Google Scholar] [CrossRef]
  53. Lee, S.; Jung, H.J.; Choi, H.C.; Hwang, Y.S.; Choi, M.Y. Solvent acting as a precursor: Synthesis of AgCN from AgNO3 in N,N-DMF solvent by laser ablation. Bull. Korean Chem. Soc. 2017, 38, 136–139. [Google Scholar] [CrossRef]
  54. Zhou, X.; Zhang, G.; Gao, B.; Huang, H. Palladium-catalyzed hydrocarbonylative C-N coupling of alkenes with amides. Org. Lett. 2018, 20, 2208–2212. [Google Scholar] [CrossRef] [PubMed]
  55. Bi, X.; Li, J.; Shi, E.; Wang, H.; Gao, R.; Xiao, J. Ru-catalyzed direct amidation of carboxylic acids with N-substituted formamides. Tetrahedron 2016, 72, 8210–8214. [Google Scholar] [CrossRef]
  56. Sun, Y.-J.; Li, P.; Huang, Q.-Q.; Zhang, J.-J.; Itoh, S. Dioxygenation of flavonol catalyzed by copper(II) complexes supported by carboxylate-containing ligands: Structural and functional models of quercetin 2,4-dioxygenase. Eur. J. Inorg. Chem. 2017, 2017, 1845–1854. [Google Scholar] [CrossRef]
  57. Chen, X.; Yang, Q.; Zhou, Y.; Deng, Z.; Mao, X.; Peng, Y. Synthesis of 4-(dimethylamino)quinazoline via direct amination of quinazolin-4(3H)-one using N,N-dimethylformamide as a nitrogen source at room temperature. Synthesis 2015, 47, 2055–2062. [Google Scholar] [CrossRef]
  58. Gupta, S.S.R.; Nakhate, A.V.; Rasal, K.B.; Deshmukh, G.P.; Mannepalli, L.K. Oxidative amidation of benzaldehydes and benzylamines with N-substituted formamides over a Co/Al hydrotalcite-derived catalyst. New J. Chem. 2017, 41, 15268–15276. [Google Scholar] [CrossRef]
  59. Liu, W.; Chen, C.; Zhou, P. Concise access to α-arylketothioamides by redox reaction between acetophenones, elemental sulfur and DMF. ChemistrySelect 2017, 2, 5532–5535. [Google Scholar] [CrossRef]
  60. Liu, W.; Chen, C.; Liu, H. Dimethylamine as the key intermediate generated in situ from dimethylformamide (DMF) for the synthesis of thioamides. Beilstein J. Org. Chem. 2015, 11, 1721–1726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Amupitan, J.O. An extension of the Willgerodt-Kindler reaction. Synthesis 1983, 1983, 730. [Google Scholar] [CrossRef]
  62. Nooshabadi, M.; Aghapoor, K.; Darabi, H.R.; Mojtahedi, M.M. The rapid synthesis of thiomorpholides by Willgerodt-Kindler reaction under microwave heating. Tetrahedron Lett. 1999, 40, 7549–7552. [Google Scholar] [CrossRef]
  63. Wei, E.; Liu, B.; Lin, S.; Liang, F. Multicomponent reaction of chalcones, malononitrile and DMF leading to γ-ketoamides. Org. Biomol. Chem. 2014, 12, 6389–6392. [Google Scholar] [CrossRef] [PubMed]
  64. Huang, X.; Wang, J.; Ni, Z.; Wang, S.; Pan, Y. Copper-mediated S–N formation via an oxygen-activated radical process: A new synthesis method for sulfonamides. Chem. Commun. 2014, 50, 4582–4584. [Google Scholar] [CrossRef] [PubMed]
  65. Bao, X.; Rong, X.; Liu, Z.; Gu, Y.; Liang, G.; Xia, Q. Potassium tert-butoxide-mediated metal-free synthesis of sulfonamides from sodium sulfinates and N,N-disubstituted formamides. Tetrahedron Lett. 2018, 59, 2853–2858. [Google Scholar] [CrossRef]
  66. Zeng, W.; Wang, E.Y.; Qiu, R.; Sohail, M.; Wu, S.; Chen, F.X. Oxygen-atom insertion of NHC-copper complex: The source of oxygen from N,N-dimethylformamide. J. Organomet. Chem. 2013, 743, 44–48. [Google Scholar] [CrossRef]
  67. Chen, J.; Natte, K.; Wu, X.-F. Pd/C-catalyzed carbonylative C-H activation with DMF as the CO source. Tetrahedron Lett. 2015, 56, 6413–6416. [Google Scholar] [CrossRef]
  68. Feng, X.; Li, Z. Photocatalytic promoting dimethylformamide (DMF) decomposition to in-situ generation of self-supplied CO for carbonylative Suzuki reaction. J. Photochem. Photobiol. A Chem. 2017, 337, 19–24. [Google Scholar] [CrossRef]
  69. Krishnakumar, V.; Chatterjee, B.; Gunanathan, C. Ruthenium-catalyzed urea synthesis by N-H activation of amines. Inorg. Chem. 2017, 56, 7278–7284. [Google Scholar] [CrossRef] [PubMed]
  70. Yang, Y.; Zhang, X.; Zeng, W.; Huang, H.; Liang, Y. Copper catalyzed three-component synthesis of benzothiazolones from o-iodoanilines, DMF, and potassium sulfide. RSC Adv. 2014, 4, 6090–6093. [Google Scholar] [CrossRef]
  71. Mete, T.B.; Singh, A.; Bhat, R.G. Transition-metal-free synthesis of primary to tertiary carboxamides: A quick access to prodrug-pyrazinecarboxamide. Tetrahedron Lett. 2017, 58, 4709–4712. [Google Scholar] [CrossRef]
  72. Yao, B.; Deng, C.-L.; Liu, Y.; Tang, R.-Y.; Zhang, X.-G.; Li, J.-H. Palladium-catalyzed oxidative carbamoylation of isoquinoline N-oxides with formylamides by means of dual C-H oxidative coupling. Chem. Commun. 2015, 51, 4097–4100. [Google Scholar] [CrossRef] [PubMed]
  73. Wang, H.; Guo, L.-N.; Wang, S.; Duan, X.-H. Decarboxylative alkynylation of α-keto acids and oxamic acids in aqueous media. Org. Lett. 2015, 17, 3054–3057. [Google Scholar] [CrossRef] [PubMed]
  74. Singh, S.J.; Mir, B.A.; Patel, B.K. A TBPB-mediated C-3 cycloalkylation and formamidation of 4-arylcoumarin. Eur. J. Org. Chem. 2018, 2018, 1026–1033. [Google Scholar] [CrossRef]
  75. Fan, W.; Shi, D.; Feng, B. TBAI-catalyzed synthesis of α-ketoamides via sp3 C-H radical/radical cross-coupling and domino aerobic oxidation. Tetrahedron Lett. 2015, 56, 4638–4641. [Google Scholar] [CrossRef]
  76. Chen, W.-T.; Bao, W.-H.; Ying, W.-W.; Zhu, W.M.; Liang, H.; Wei, W.-T. Copper-promoted tandem radical reaction of 2-oxindoles with formamides: Facile synthesis of unsymmetrical urea derivatives. Asian J. Org. Chem. 2018, 7, 1057–1060. [Google Scholar] [CrossRef]
  77. Li, J.; Hua, R. Stereodivergent ruthenium-catalyzed transfer semihydrogenation of diaryl alkynes. Chem. Eur. J. 2011, 17, 8462–8465. [Google Scholar] [CrossRef] [PubMed]
  78. Li, J.; Hua, R.; Liu, T. Highly chemo- and stereoselective palladium-catalyzed transfer semihydrogenation of internal alkynes affording cis-alkenes. J. Org. Chem. 2010, 75, 2966–2970. [Google Scholar] [CrossRef] [PubMed]
  79. Tam, C.M.; To, C.T.; Chan, K.S. Carbon-carbon σ-bond transfer hydrogenation with DMF catalyzed by cobalt porphyrins. Organometallics 2016, 35, 2174–2177. [Google Scholar] [CrossRef]
  80. Nale, D.B.; Bhanage, B.M. N-substituted formamides as C1-sources for the synthesis of benzimidazole and benzothiazole derivatives by using zinc catalysts. Synlett 2015, 26, 2835–2842. [Google Scholar] [CrossRef]
  81. Gao, X.; Yu, B.; Mei, Q.; Yang, Z.; Zhao, Y.; Zhang, H.; Hao, L.; Liu, Z. Atmospheric CO2 promoted synthesis of N-containing heterocycles over B(C6F5)3 catalyst. New J. Chem. 2016, 40, 8282–8287. [Google Scholar] [CrossRef]
  82. Zhu, J.; Zhang, Z.; Miao, C.; Liu, W.; Sun, W. Synthesis of benzimidazoles from o-phenylenediamines and DMF derivatives in the presence of PhSiH3. Tetrahedron 2017, 73, 3458–3462. [Google Scholar] [CrossRef]
  83. Simion, C.; Gherase, D.; Sima, S.; Simion, A.M. Serendipitous synthesis of a cyclic formamidine. Application to the synthesis of polyazamacrocycles. C. R. Chim. 2015, 18, 611–613. [Google Scholar] [CrossRef]
  84. Kobayashi, E.; Togo, H. Facile one-pot preparation of 5-aryltetrazoles and 3-arylisoxazoles from aryl bromides. Tetrahedron 2018, 74, 4226–4235. [Google Scholar] [CrossRef]
  85. Ochiai, H.; Niwa, T.; Hosoya, T. Stereoinversion of stereocongested carbocyclic alcohols via triflylation and subsequent treatment with aqueous N,N-dimethylformamide. Org. Lett. 2016, 18, 5982–5985. [Google Scholar] [CrossRef] [PubMed]
  86. Yu, M.; Zhang, Q.; Wang, J.; Huang, P.; Yan, P.; Zhang, R.; Dong, D. Triflic anhydride mediated ring-opening/recyclization reaction of α-carbamoyl α-oximyl cyclopropanes with DMF: Synthetic route to 5-aminoisoxazoles. Synthesis 2016, 48, 1934–1938. [Google Scholar] [CrossRef]
  87. Liu, L.; Zhang-Negrerie, D.; Du, F.; Zhao, K. PhICl2 and wet DMF: An efficient system for regioselective chloroformyloxylation/α-chlorination of alkenes/α,β-unsaturated compounds. Org. Lett. 2014, 16, 436–439. [Google Scholar] [CrossRef] [PubMed]
  88. Wang, Y.; Zhang, J.; Liu, J.; Zhang, C.; Zhang, Z.; Xu, J.; Xu, S.; Wang, F.; Wang, F. C-N and N-H bond metathesis reactions mediated by carbon dioxide. ChemSusChem 2015, 8, 2066–2072. [Google Scholar] [CrossRef] [PubMed]
  89. Wang, Y.; Wang, F.; Zhang, C.; Zhang, J.; Li, M.; Xu, J. Transformylating amine with DMF to formamide over CeO2 catalyst. Chem. Commun. 2014, 50, 2438–2441. [Google Scholar] [CrossRef] [PubMed]
  90. Sheng, H.; Zeng, R.; Wang, W.; Luo, S.; Feng, Y.; Liu, J.; Chen, W.; Zhu, M.; Guo, Q. An efficient heterobimetallic lanthanide alkoxide catalyst for transamidation of amides under solvent-free conditions. Adv. Synth. Catal. 2017, 359, 302–313. [Google Scholar] [CrossRef]
  91. Sonawane, R.B.; Rasal, N.K.; Jagtap, S.V. Nickel-(II)-catalyzed N-formylation and N-acylation of amines. Org. Lett. 2017, 19, 2078–2081. [Google Scholar] [CrossRef] [PubMed]
  92. Lindsay-Scott, P.J.; Charlesworth, N.G.; Grozavu, A. A flexible strategy for the regiocontrolled synthesis of pyrazolo[1,5-a]pyrazines. J. Org. Chem. 2017, 82, 11295–11303. [Google Scholar] [CrossRef] [PubMed]
  93. Nakagawa-Goto, K.; Kobayashi, T.; Kawasaki, T.; Somei, M. Lithiation of 1-alkoxyindole derivatives. Heterocycles 2018. [Google Scholar] [CrossRef]
  94. Ramírez-Jiménez, R.; Franco, M.; Rodrigo, E.; Sainz, R.; Ferritto, R.; Lamsabhi, A.M.; Aceña, J.L.; Cid, M.B. Unexpected reactivity of graphene oxide with DBU and DMF. J. Mater. Chem. A 2018, 6, 12637–12646. [Google Scholar] [CrossRef]
  95. Deng, X.; Lei, X.; Nie, G.; Jia, L.; Li, Y.; Chen, Y. Copper-catalyzed cross-dehydrogenative N2-coupling of NH-1,2,3-triazoles with N,N-dialkylamides: N-amidoalkylation of NH-1,2,3-triazoles. J. Org. Chem. 2017, 82, 6163–6171. [Google Scholar] [CrossRef] [PubMed]
  96. Gonzalez, P.E.; Sharma, H.K.; Chakrabarty, S.; Metta-Magaña, A.; Pannell, K.H. Triethylsiloxymethyl-N,N-dimethylamine, Et3SiOCH2NMe2: A dimethylaminomethylation (Mannich) reagent for O–H, S–H, P–H and aromatic C–H systems. Eur. J. Org. Chem. 2017, 2017, 5610–5616. [Google Scholar] [CrossRef]
  97. Hu, Y.; Hu, Y.; Hu, Q.; Ma, J.; Lv, S.; Liu, B.; Wang, S. Direct access to fused salicylaldehydes and salicylketones from tetraynes. Chem. Eur. J. 2017, 23, 4065–4072. [Google Scholar] [CrossRef] [PubMed]
  98. Zhu, R.; Liu, Z.; Chen, J.; Xiong, X.; Wang, Y.; Huang, L.; Bai, J.; Dang, Y.; Huang, J. Preparation of thioanisole biscarbanion and C–H lithiation/annulation reactions for the access of five-membered heterocycles. Org. Lett. 2018, 20, 3161–3165. [Google Scholar] [CrossRef] [PubMed]
  99. Zeng, R.; Sheng, H.; Rao, B.; Feng, Y.; Wang, H.; Sun, Y.; Chen, M.; Zhu, M. An efficient and green approach to synthesizing enamines by intermolecular hydroamination of activated alkynes. Chem. Res. Chin. Univ. 2015, 31, 212–217. [Google Scholar] [CrossRef]
  100. Liu, D.; Mao, J.; Rong, G.; Yan, H.; Zheng, Y.; Chen, J. Concise synthesis of semicarbazides and formylhydrazines via direct addition reaction between aromatic azoarenes and N-substituted formamides. RSC Adv. 2015, 5, 19301–19305. [Google Scholar] [CrossRef]
  101. Mori, S.; Shibuya, M.; Yamamoto, Y. Ruthenium-catalyzed hydrocarbamoylative cyclization of 1,6-diynes with formamides. Chem. Lett. 2017, 46, 207–210. [Google Scholar] [CrossRef]
  102. Yamamoto, Y.; Okude, Y.; Mori, S.; Shibuya, M. Combined experimental and computational study on ruthenium(II)-catalyzed reactions of diynes with aldehydes and N,N-dimethylformamide. J. Org. Chem. 2017, 82, 7964–7973. [Google Scholar] [CrossRef] [PubMed]
  103. Yamamoto, Y. Theoretical study on ruthenium-catalyzed hydrocarbamoylative cyclization of 1,6-diyne with dimethylformamide. Organometallics 2017, 36, 1154–1163. [Google Scholar] [CrossRef]
  104. Adams, R.D.; Dhull, P. Formyl C-H activation in N,N-dimethylformamide by a dirhenium carbonyl complex. J. Organomet. Chem. 2017, 849–850, 228–232. [Google Scholar] [CrossRef]
  105. Guo, W.; Liao, J.; Liu, D.; Li, J.; Ji, F.; Wu, W.; Jiang, H. A four-component reaction strategy for pyrimidine carboxamide synthesis. Angew. Chem. Int. Ed. 2017, 56, 1289–1293. [Google Scholar] [CrossRef] [PubMed]
  106. Pastoriza-Santos, I.; Liz-Marzán, L.M. N,N-Dimethylformamide as a reaction medium for metal nanoparticle synthesis. Adv. Funct. Mater. 2009, 19, 679–688. [Google Scholar] [CrossRef]
  107. Akbari, S.; Mokhtari, J.; Mirjafari, Z. Solvent-free and melt aerobic oxidation of benzyl alcohols using Pd/Cu2(BDC)2DABCO–MOF prepared by one-step and through reduction by dimethylformamide. RSC Adv. 2017, 7, 40881–40886. [Google Scholar] [CrossRef]
  108. Tahmasebi, S.; Mokhtari, J.; Naimi-Jamal, M.R.; Khosravi, A.; Panahi, L. One-step synthesis of Pd-NPs@Cu2(BDC)2DABCO as efficient heterogeneous catalyst for the Suzuki-Miyaura cross-coupling reaction. J. Organomet. Chem. 2017, 853, 35–41. [Google Scholar] [CrossRef]
  109. Azuma, R.; Nakamichi, S.; Kimura, J.; Yano, H.; Kawasaki, H.; Suzuki, T.; Kondo, R.; Kanda, Y.; Shimizu, K.-I.; Kato, K.; et al. Solution synthesis of N,N-dimethylformamide-stabilized iron-oxide nanoparticles as an efficient and recyclable catalyst for alkene hydrosilylation. ChemCatChem 2018, 10, 2378–2382. [Google Scholar] [CrossRef]
  110. Onishi, K.; Oikawa, K.; Yano, H.; Suzuki, T.; Obora, Y. N,N-Dimethylformamide-stabilized palladium nanoclusters as a catalyst for Larock indole synthesis. RSC Adv. 2018, 8, 11324–11329. [Google Scholar] [CrossRef]
  111. Oka, H.; Kitai, K.; Suzuki, T.; Obora, Y. N,N-Dimethylformamide-stabilized copper nanoparticles as a catalyst precursor for Sonogashira–Hagihara cross coupling. RSC Adv. 2017, 7, 22869–22874. [Google Scholar] [CrossRef]
  112. Oikawa, K.; Itoh, S.; Yano, H.; Kawasaki, H.; Obora, Y. Preparation and use of DMF-stabilized iridium nanoclusters as methylation catalysts using methanol as the C1 source. Chem. Commun. 2017, 53, 1080–1083. [Google Scholar] [CrossRef] [PubMed]
  113. Rasal, K.B.; Yadav, G.D. One-pot synthesis of benzimidazole using DMF as a multitasking reagent in presence CuFe2O4 as catalyst. Catal. Today 2018, 309, 51–60. [Google Scholar] [CrossRef]
Scheme 1. Fragments from DM (R = H or Me) used in synthesis.
Scheme 1. Fragments from DM (R = H or Me) used in synthesis.
Molecules 23 01939 sch001
Scheme 2. Reactions of DM (R = H or Me).
Scheme 2. Reactions of DM (R = H or Me).
Molecules 23 01939 sch002
Scheme 3. Aryne formation and reaction with DM.
Scheme 3. Aryne formation and reaction with DM.
Molecules 23 01939 sch003

Share and Cite

MDPI and ACS Style

Le Bras, J.; Muzart, J. Recent Uses of N,N-Dimethylformamide and N,N-Dimethylacetamide as Reagents. Molecules 2018, 23, 1939. https://doi.org/10.3390/molecules23081939

AMA Style

Le Bras J, Muzart J. Recent Uses of N,N-Dimethylformamide and N,N-Dimethylacetamide as Reagents. Molecules. 2018; 23(8):1939. https://doi.org/10.3390/molecules23081939

Chicago/Turabian Style

Le Bras, Jean, and Jacques Muzart. 2018. "Recent Uses of N,N-Dimethylformamide and N,N-Dimethylacetamide as Reagents" Molecules 23, no. 8: 1939. https://doi.org/10.3390/molecules23081939

Article Metrics

Back to TopTop