Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

ORGANIC CHEMISTRY

FRONTIERS

View Article Online


REVIEW View Journal
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Recent advances in the synthesis and reactivity of


Cite this: DOI: 10.1039/d0qo01575j
quinoxaline
Gauravi Yashwantrao and Satyajit Saha *

Quinoxaline has become a subject of extensive research due to its emergence as an important chemical
moiety, demonstrating a wide range of physicochemical and biological activities. The last few decades
have witnessed several publications utilizing quinoxaline scaffolds for the design and development of
numerous bioactive molecules, dyes, fluorescent materials, electroluminescent materials and organic
sensitizers for solar cell applications and polymeric optoelectronic materials. Therefore, to fulfill the need
of the scientific community, tremendous effort has been observed in the development of newer synthetic
strategies as well as novel methodologies to decorate the quinoxaline scaffold with proper functional
groups. Hence, to provide an updated comprehensive account of the diverse synthetic routes to access
Received 16th December 2020, quinoxaline as well as approaches for structural diversifications, we have attempted to document the syn-
Accepted 10th March 2021
thetic strategies and their functionalization with possible mechanistic rationalization. This will no doubt be
DOI: 10.1039/d0qo01575j helpful for the readers who are anticipating a comprehensive overview on quinoxaline as well as benefi-
rsc.li/frontiers-organic tting researchers for future development.

Quinoxaline belongs to a class of aromatic heterocycles quinoxaline are analogous to pyrazine. The positions 2, 4, 5, 7,
formed by the fusion of two six-membered aromatic rings, one and 8a of quinoxaline are electron deficient, as can be
of which consists of two nitrogen atoms symmetrically placed observed from the resonating structures depicted in Fig. 1, and
at the 1 and 4 positions. In general, the physical properties of hence, are susceptible to nucleophilic attack.
Quinoxaline has received a great deal of attention over the
Department of Speciality Chemicals Technology, Institute of Chemical Technology,
years. This molecule has become a subject of extensive
Mumbai-400019, India. E-mail: ss.saha@ictmumbai.edu.in research because of the unique disposition of the two nitrogen

Ms Gauravi B. Yashwantrao Satyajit Saha earned his Ph.D.


completed her SSC from B.P.V., (organic chemistry) in the year
Ambernath, and HSC from 2011 from IIT Kanpur, India.
B. K. Birla College, Kalyan in Thereafter, as a FWO Visiting
2010 and 2012, respectively. Postdoctoral Fellow, he visited
Thereafter, she pursued a Belgium for his postdoctoral
Bachelor of Science with research at the University of
Chemistry as a major in 2015 Antwerpen, Belgium. In 2013 he
from B.K. Birla College and M.Sc moved to Universitat Leipzig,
in Organic Chemistry in 2017 Germany for his second post-doc-
from Ruia College, Matunga. She toral research. He returned to
then worked as an ‘Assistant India in 2015 to join the
Gauravi Yashwantrao Research Associate’ at Oriental Satyajit Saha Institute of Chemical
Aromatics Ltd for one year. In Technology, Mumbai, India as
October 2018, she joined the Institute of Chemical Technology, an Assistant Professor at the Department of Dyestuff Technology
Mumbai to pursue her Ph.D. under the guidance of Dr Satyajit under the UGC-Faculty Recharge Programme. Currently, he is
Saha. Her research interest is in the field of synthesis and process associated with the Department of Speciality Chemicals
intensification of N-heterocycles using greener approaches and Technology of ICT Mumbai. His research interests are focused on
their applications thereafter as functional materials. organic synthesis, catalysis, synthesis of functional materials,
stereoselective transformations, and green chemistry.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

zation and insights about their mechanistic pathways. The


review has been subdivided into two parts. The first part of
this review will focus on the several synthetic approaches to
access quinoxaline, while the second part will discuss the
functionalization routes of quinoxaline and its derivatives with
possible mechanistic interpretation. We believe that this
update will be beneficial for the advancement of quinoxaline
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

chemistry as well as equip the researchers for future


development.

Fig. 1 Resonating structures of quinoxaline.


Synthesis of quinoxalines
The significance of the quinoxaline scaffold in the areas of
atoms in one of the rings, which is abundant in many natural medicinal chemistry, electroluminescent materials, DSSC,
and man-made compounds. The 1,4 disposition of the two molecular probes, and sensors has resulted in an upsurge of
imine nitrogen atoms in one of the rings of the quinoxaline the synthetic strategies to access quinoxaline sustainably
bestows it with the electron-withdrawing property. In addition under milder conditions in higher yields. The recent develop-
to that, their easy synthetic accessibility has been the major ment in synthesizing quinoxaline derivatives has witnessed
reason for the extensive applications of quinoxaline derivatives several strategies that avoid hazardous chemicals, release fewer
as dyes, or as a building block in the synthesis of organic semi- chemical wastes and rely on sustainable technology.
conductors, electroluminescent material, dye-sensitized solar The synthetic methodologies to access quinoxalines as
cells (DSSC), etc. Even quinoxaline derivatives possess a wide highlighted in Fig. 4 can be classified into four major cat-
range of physicochemical and biological activities such as anti- egories and are listed below.
bacterial, antiviral, anticancer activity, etc., Fig. 2.1,2 1. Condensation reaction
The last two decades have witnessed the appearance of 2. Oxidative condensation
several publications demonstrating the utility of quinoxaline 3. Synthesis via Schmidt reaction
scaffolds in the design and development of numerous bio- 4. Oxidative dehydrogenation
active molecules, dyes, fluorescent materials, electrolumines- Methods B, E, F, G, and H are categorized as condensation
cent materials, organic sensitizers for solar cell applications reactions, while method C proceeds via the Schmidt reaction,
and polymeric optoelectronic materials, Fig. 3. Therefore, to method D represents oxidative condensation pathways and A
fulfill the need of the scientific community, there has been a represents oxidative dehydration reactions. Each method will
persistent effort in the development of newer synthetic strat- be discussed separately emphasizing the approaches under-
egies as well as novel methodologies to decorate the quinoxa- taken, as well as highlighting the different conditions, catalysts
line scaffold with proper functional groups. The necessity to used and strategies that will project the advancement in the
build the scaffold has resulted in a huge volume of publi- synthetic routes.
cations that have appeared since the first synthesis of quinoxa-
line in 1950.3 Several synthetic protocols that are mild, Condensation reaction
efficient, and sustainable have been developed over the last The condensation of ortho-phenylenediamines (o-PD)’s 2 with
two decades, so it will be a Herculean task to summarize the aryl or alkyl-substituted 1,2-dicarbonyl compounds 3 in the
developments in terms of synthesis and reactivity unless done presence of either a Lewis acid or a Brønsted acid catalyst is
systematically (Fig. 4). one the most widely employed protocols to access quinoxaline
Two consecutive reviews by Zhukova et al. that appeared in 4, Scheme 1.
2012 and 2013 highlighted the synthesis and reactivities of The ease of synthesis along with the commercial availability
quinoxaline.4,5 Thereafter Vieira et al. in 20156 and Menezes of the starting materials makes this cyclocondensation reac-
et al. in 20207 reported two reviews describing the applications tion a widely explored route to access quinoxaline. Primarily,
of 2,3-dichloroquinoxaline as a building block and the devel- the Lewis acids or Brønsted acids employed are associated
opment of bioactive quinoxaline derivatives, respectively. with the activation of the carbonyls which facilitate the nucleo-
Several publications have appeared since then on the synthesis philic addition followed by the cyclocondensation reaction, as
and functionalization of quinoxaline which need to be highlighted in Scheme 2.
addressed systematically. Therefore, the need for an update is Over the years, a variety of Lewis acids, as well as Brønsted
obvious considering the number of publications that have acids, have been employed by several research groups to syn-
appeared since the last review on quinoxaline scaffolds, high- thesize quinoxaline 4 from o-PDs 2 and 1,2-diketones 3.
lighting their synthesis and reactivity.5 Therefore, Table 1 provides a glimpse of the variety of Lewis
This review will provide an update on the synthetic strat- and Brønsted acid catalysts that were used to access quinoxa-
egies of quinoxalines, their possible routes for functionali- line via the condensation pathway.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Fig. 2 Representative examples of molecules containing a quinoxaline scaffold.

The use of aqueous hydrofluoric acid acting both as a reac- 4 from 1,2-dicarbonyls 3. Apart from the recyclability of the
tion medium as well as a catalyst to access quinoxaline 4 in a catalyst, mild reaction conditions to obtain the products in
relatively milder condition in good yield was demonstrated nearly quantitative yields within a short reaction time offered
by Shekhar et al. The absence of coupling agents and easy obvious advantages to access temperature or acid-sensitive
workup procedures highlights the advantages of this protocol substrates (entry 2, Table 1).9 Later, More10 and Bhosale11 syn-
(entry 1, Table 1).8 Sulfamic acid in MeOH was employed as a thesized quinoxaline 4 in excellent yields using a catalytic
catalyst by the group of Darabi for the synthesis of quinoxaline amount of iodine in MeCN and DMSO solvent, respectively

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 1 Condensation of 1,2-diamines 2 with 1,2-dicarbonyls 3 with


Brønsted and Lewis acid catalysts.

Fig. 3 Schematic representation of the areas of quinoxaline


applications.

(entries 3 and 4, Table 1). The high yield and short duration of
the reaction coupled with the use of non-toxic and inexpensive
iodine as a catalyst featured in the usefulness of the reaction.
In 2007, Heravi demonstrated the use of CuSO4·5H2O as an in-
expensive and mild Lewis acid catalyst to access quinoxaline 4
Scheme 2 Mechanistic pathway to access quinoxaline by the conden-
at room temperature in a solvent mixture of EtOH and water
sation of (o-PD)s 2 with 1,2-dicarbonyls 3 employing Brønsted and
(entry 5, Table 1).12 The group of Heravi further described a Lewis acid catalysts.
simple, one-pot efficient method for the synthesis of quinoxa-
lines via MnCl2-catalyzed condensation of a variety of o-PDs
and α-diketone derivatives (entry 6, Table 1).13 mechanochemical grinding of o-PDs 2 with 1,2-dicarbonyls 3.
However, many of these methods were not sustainable and The operational simplicity of running the reaction in solvent-
do not feature sufficient environmental benignancy in their free conditions which avoids the work-up and purification
reaction parameters. So, with the intention of avoiding volatile step was the key advantage of this process (entry 7, Table 1).14
organic solvents, Bhutia and the group established a solvent- Recently, Sajjadifar et al. explored silica-supported 1-(2-
free synthetic protocol to access quinoxaline by liquid-assisted (sulfoxy)ethyl)pyridine-1-ium chloride as an efficient recyclable

Fig. 4 Available methods for the synthesis of quinoxaline.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Table 1 Reaction conditions and catalysts employed in the conden- Polyaniline-sulfate salt 5, an easily synthesizable Brønsted
sation reaction of o-PD 2 with 1,2-diketones 3 acid catalyst, is known for its excellent acidic properties along
with easy recoverability, reusability, and stability. In 2007
the research group of Srinivas demonstrated the synthesis of
quinoxaline in excellent yields by the condensation reaction
using a catalytic amount of polyaniline-sulfate salt (entry 1,
Table 2).16
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Harsha et al. developed a magnetically retrievable hetero-


Entry Reaction conditions % yield geneous nano-γ-Fe2O3-SO3H catalytic system for the cyclocon-
densation of 1,2-diketones or α-bromoketones with o-PDs to
1 Aq. HF/rt (10 min–90 min) >958
GdCl3/EtOH, reflux (1–10 h) >95 access quinoxaline under solvent-free conditions (entry 2,
2 Sulfamic acid/MeOH (5 mol%), rt., 5 min up to 959 Table 2).17 Easy work-up and purification processes, along
3 I2 (10 mol%), MeCN, 25 °C, 3–30 min, >9010 with high product yield obtained in less time, are the obvious
4 I2 (10 mol%), DMSO, rt., >8511
5 CuSO4·5H2O (10 mol%), EtOH, rt., 10–15 min >9512 advantages of this environmentally benign protocol. The cata-
CuSO4·5H2O (10 mol%), water, rt., 15–30 min lyst can be recovered magnetically and can be reused up to 5
6 MnCl2 (10 mol%), rt., 15–20 min >9013 times. The electron-rich o-PDs afforded excellent yields of the
7 Mechanochemical grinding, EtOH, rt, 10–30 min 85–9514
8 Silica supported 1-(2- (sulfoxy)ethyl)pyridin-1-ium up to desired quinoxalines compared with the o-PDs attached with
chloride (SiO2/[SEP]Cl), EtOH, 65 °C, 15–20 min 9715 electron-withdrawing groups.17
Sulfated polyborate 6 belongs to a class of promising cata-
lytic systems displaying both Lewis and Brønsted acidic pro-
perties that were used to synthesize quinoxaline in high yields
catalyst for the synthesis of quinoxaline in the presence of
within a short reaction time.23 The easy workup procedure
ethanol as a green solvent at 65 °C (entry 8, Table 1).15
along with broad substrate scope provided both economical as
There were relentless efforts, which are still ongoing, to
well as ecological rewards to this protocol (entry 3, Table 2).18
look for alternative strategies to access quinoxaline sustain-
ably. One way to achieve environmental benignancy is by cata-
lyst immobilization, which has become one of the prominent
approaches to address the issue of sustainability. Moreover,
catalyst immobilization on a solid support seems to be crucial
in terms of stability, catalytic site availability, catalyst recover-
ability, and reusability. Based on this rationale, several
researchers have come up with different catalytic systems for
the efficient synthesis of quinoxaline as featured in Table 2.

Brønsted acidic ionic liquids (BAILs) belong to a set of ILs dis-


playing dual properties of Brønsted acids and ionic liquids (IL)
Table 2 Solid supported catalysts used for the synthesis of quinoxaline
from o-PD 2 and 1,2-diketones 3 simultaneously. In addition to that, one ionic center is co-
valently grafted to the polymeric backbone which features the
unique properties of ILs as well as the macromolecular poly-
meric architectures. Tamami prepared two silica-supported
polyvinyl imidazole based Brønsted acidic ionic liquids and
demonstrated their application in the sustainable synthesis of
quinoxaline derivatives in high yields (entry 4, Table 2).19

Entry Reaction conditions % yield

1 Polyaniline sulfate salt catalyst (5 wt%), DCE, rt, >9016


10 min–2 h
2 Nano-γ-Fe2O3-SO3H (0.1 g), solvent-free, 120 °C, 1 h 85–9017
(1,2-diketone)
Nano-γ-Fe2O3-SO3H (0.1 g), solvent-free, 120 °C, 1 h
(α-bromoketone)
3 Sulfated polyborate (10 wt%), solvent-free, 100 °C, >9518
3–10 min
4 Silica-supported polymeric Brønsted acidic ionic 85–9019
liquid catalyst (0.5 mol%), EtOH, rt
5 Epoxidized novolac phenol-formaldehyde resin- 80–8520
modified using sulphanilic acid (ENPFSA) (5 mol%), Tarpada and co-workers designed a polymer-supported hetero-
EtOH, rt
6 PS/AlCl3 (10 mol%), EtOH, reflux 84–9721 geneous catalyst based on sulphanilic acid for the one-pot syn-
7 BiCl3/SiO2 (5 mol%), MeOH, rt 85–9522 thesis of various quinoxaline derivatives 4. The developed

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

methodology features the advantages of operational simplicity, Table 3 Synthetic strategies for the preparation of quinoxaline 4 in
easy workup, and recyclability to furnish the desired quinoxa- water from o-PD 2 and 1,2-diketones 3
lines with excellent purity and high yields. The reaction was
Entry Conditions % yield
very much substrate-dependent. Diamines with electron-with-
drawing substituents afforded desired products in high yields 1 Water, rt, 30 min 70–9025
2 In powder (5 mol%), MWI (260 W, 60 °C, up to 9726
in comparison with the diamines with electron-donating
24–50 psi), water, <3 min
attachments (entry 5, Table 2).20 3 Itaconic acid (15 mol%), water, rt, 1 h up to 9627
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Like the polymer grafted Brønsted acids, heterogenizing


the homogeneous metal species by supporting it on in-
soluble supports is also an attractive strategy in terms of Table 4 Solvent-free strategies for the synthesis of quinoxalines 4 from
catalyst recovery and recyclability. In 2012, Rahmatpour and o-PD 2 and 1,2-diketones 3
co-workers used aluminum chloride supported on cross-
linked polystyrene (PS/AlCl3) as a recyclable heterogeneous Entry Conditions % yield
Lewis acid catalyst for the synthesis of quinoxaline via a con- 1 Selectfluor (1 mol%), EtOH, rt., 5 min 85–9028
densation reaction (entry 6, Table 2).21 Although AlCl3 is a Selectfluor (1 mol%), solvent-free condition, 3 min 90–9528
water-sensitive, corrosive, and environmentally harmful com- 2 Nano-TiO2 (0.12 mol), solvent-free, rt. <9629
pound, the PS–AlCl3 complex has sufficient stability and can
be easily recycled and reused without appreciable loss of its
activity. agent made that protocol an appealing sustainable method-
The concept of heterogenization of the metal species was ology compared with the conventional thermal processes invol-
further extended by the group of Aghapoor in the acid-cata- ving acid–base catalysts (entry 1, Table 3).25
lyzed condensation of various o-PDs 2 with 1,2-diketone 3. To Microwave technology has emerged as a rapid synthetic tool
design catalysts with high selectivity and easy separability, to access diverse molecules sustainably. To access quinoxaline
bismuth has been integrated into the framework of silica. The sustainably, Bandyopadhyay reported a microwave-assisted
catalyst (silica-supported bismuth(III) chloride (BiCl3/SiO2)) convenient and expeditious synthesis of quinoxaline in water
exhibited remarkable catalytic performance and can be reused, using a trace amount of indium catalyst. The use of indium
unlike the homogeneous BiCl3 (entry 7, Table 2).22 circumvented the toxicity, cost, and air and moisture sensi-
Catalyst recoverability is an appealing strategy in terms of tivity associated with the organometallic reagents used as cata-
achieving sustainability. Paul investigated the KF-alumina pro- lysts (entry 2, Table 3).26 The activation by indium followed by
moted direct conversion of α-hydroxy ketones 10a as well as cyclocondensation leads to the desired product. In the search
α-halo ketones 10b into quinoxalines 4. It was observed that for an environmentally benign sustainable approach to synthe-
functionalized quinoxalines as a condensation product of sizing the quinoxaline scaffold in water, Tamuli et al. devel-
o-PDs 2 and α-hydroxy ketones 10a or α-bromo ketones 10b oped a synthetic water route using o-PDs 2 and 1,2 diketone as
can be obtained at ambient temperature in good yield, as reactants and biodegradable itaconic acid as a mild acid pro-
shown in Scheme 3.24 motor. The proposed method is advantageous as it is energy
Performing organic transformations in water is always a efficient, requires less reaction time, and the catalyst is reusa-
challenging and demanding task. The reactions often take ble (entry 3, Table 3).27
place as an emulsion and exhibit unusual rate enhancement. The drive for the development of solvent-free reactions to
The advantages of replacing organic solvents with water rely eliminate volatile organic solvents as well as to expedite the
on the non-flammability and non-toxicity of the water exhibit- product recovery process without involving tedious extraction
ing strong polar character due to its intense hydrogen and evaporation steps has escalated tremendously to a new
bonding. Few efforts are evident in the literature where qui- horizon. Similar efforts were also noticed where quinoxaline
noxaline has been synthesized in an aqueous environment was synthesized under solvent-free conditions using catalysts
(Table 3). like selectfluor and nanoTiO2 in good to excellent yields
Delpivo reported a catalyst-free efficient synthesis of a series (Table 4).
of quinoxaline in water by reacting 1,2-dicarbonyl compounds Selectfluor, a commercially available, inexpensive, non-
with o-PDs. The high yield and the absence of any condensing toxic, non-volatile fluorinating agent, known to have efficient
Lewis acidic properties, was used as a catalyst for condensation
reactions by Hojati and co-workers (entry 1, Table 4).28
Mirjalili and co-workers further reported the synthesis of qui-
noxaline in the presence of nano-TiO2 under solvent-free con-
ditions with yields ranging from 88–90% in just 30–35 min
(entry 2, Table 4).29
A. Oxidative coupling. Another prominent strategy to access
Scheme 3 Tandem oxidation-condensation or condensation reactions quinoxaline relies on the redox condensation of 1,2-diols 12
using KF-alumina catalyst. with o-PDs 2 (Scheme 4). The use of renewable feedstocks in

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 4 General schematic representation of oxidative coupling


reaction to access quinoxaline.

chemical transformations is one of the many drives towards


sustainability. Carbohydrates, as biomass, were routinely used
as raw materials for the synthesis of a diverse set of chemicals.
Clement et al. have utilized carbohydrate feedstocks to develop
an efficient one-pot two-step methodology for the synthesis of Scheme 6 Synthesis of quinoxaline 4 using o-nitroaniline and the
probable mechanistic pathway.
quinoxalines by oxidative coupling of glycols 14 derived from
biomass with o-PD derivatives 2. The reaction was catalyzed by
gold nanoparticles supported on nanoparticulated ceria (Au/
CeO2) or hydrotalcite (Au/HT). The reaction proceeds utilizing Roy and co-workers developed a one-pot and metal-free
air as the oxidant in the absence of any homogeneous base via tandem synthesis of quinoxalines from ortho-nitroaniline 16 or
a cascade process with yields ranging between 75–80%. The dinitrobenzene 13 using graphene oxide (GO) as the ‘carboca-
rate-limiting step for the reaction is the alcohol oxidation step talyst’ in conjugation with hydrazine hydrate in 85–90% yield
to generate the 1,2-diketones (Scheme 5).30 (Scheme 6).32
Nguyen et al. reported a method employing a sodium Reduced graphene oxide (rGO) in combination with a large
sulfide (40 mol%) and iron(III) chloride hexahydrate (1 mol%) excess of hydrazine hydrate were competent to reduce nitro-
combination for the redox condensation of o-nitro aniline 16 benzene to aniline. In this reaction, the nitro group is in situ
with alcohols to synthesize quinoxaline 4 with up to 85% yield. being reduced by graphene oxide or reduced graphene oxide to
Besides being a precursor for the iron-sulfur (Fe/S) catalyst, the corresponding amine group followed by cyclocondensation
hydrated sodium sulfide also behaved as an excellent non- as depicted in Scheme 6.32
competitive, multi-electron reducing agent (Scheme 6).31 DMSO-propyl phosphonic anhydride is a well-known oxidiz-
ing and cyclodehydrating agent. Harsha and co-workers
reported the use of propyl phosphonic anhydride–DMSO for
the first time as an oxidative condensing reagent for the syn-
thesis of quinoxalines 4 in excellent yields under simple and
mild reaction conditions in one step (Scheme 7).33 The
mechanistic details are highlighted in Scheme 7.
Bera and co-workers reported a (NiCl2/1,10-phenanthroline)
system for the synthesis of quinoxaline via dehydrogenative
condensation of primary alcohols and ethylene glycol 18 with
o-PDs 2 (Scheme 8). The catalytic system enabled the synthesis
of quinoxaline by releasing water and hydrogen gas as bypro-
ducts. The methodology has wide functional group tolerance
and tolerated challenging functional groups, like trifluoro-
methyl, as well as heterocyclic rings including unsaturated
alcohols (Scheme 8).34
An improvement in the yield of the above reaction was done
by Daw and co-workers in 2018. They have presented an acri-
dine-based pincer complex of earth-abundant manganese that
Scheme 5 Mechanistic pathway for the synthesis of quinoxaline from catalyzed the formation of quinoxalines 4 by dehydrogenative
ortho-dinitrobenzene 13 and vicinal diols 14. coupling of o-PD 2 with 1,2-vicinal diols 14 in good to excellent

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 9 Mechanistic pathway for the dehydrogenative coupling


reaction for the formation of quinoxaline.

Scheme 7 Phosphorus acid-mediated tandem oxidative condensation


reaction to access quinoxaline and the probable mechanistic pathway
for the synthesis.

Scheme 8 Nickel catalyzed tandem oxidative condensation protocol


for the synthesis of quinoxaline.

Scheme 10 Oxidative cycloaddition of o-diamine 2 with various diols


35
yields (Scheme 9). The mechanistic pathway discloses that 17 or amino alcohols 20 using [Ru] catalyst.

the terminal alcoholic group of 1,2 diol system 14 gets oxidized


to the corresponding carbonyl 14a under the influence of a
manganese catalyst. Subsequently, diamine 2 condenses with Samsonov by heating benzofurazans 25 with ethanolamine 26,
carbonyl moiety 14a followed by a proton shift and tautomeri- which acts as a reducing agent (Scheme 12).38 The mechanistic
zation furnishing the intermediate 14f. The second conden- pathway highlights the formation of enamine 26a from
sation reaction of 14f with amine, and on subsequent aromati- ethanolamine 26 on treatment with acid. Subsequent attack of
zation, delivered the final product in good yield 15 (Scheme 9). enamine on N atoms opened the furazan ring 25 and the ring
In 2015, Tang and co-workers developed a diruthenium closure furnished quinoxaline-N-oxide 26d. Finally, the oxygen
complex with 2,7-bis(di-2-pyridinyl)-1,8-naphthyridine to act as atom of N-oxide gets eliminated upon treatment using ethanol-
a catalyst for the oxidative coupling of 1,2-diols/1,2-amino amine 26 to furnish the substituted quinoxaline 27.
alcohol 17/20 with o-PDs 2 leading to quinoxalines 4 in yields B. Schmidt reaction. Although o-PD is known to be the most
ranging up to 75% (Scheme 10).36 widely used precursor for the synthesis of quinoxaline, ortho-
Dowlati et al. developed an electrochemical method for the halo carboxylic acid was also used as a starting precursor to
synthesis of quinoxaline 24 in an aqueous solution. The room access quinoxaline. An altogether different approach has been
temperature synthesis of quinoxaline featured low energy con- designed where the conversion of carboxylic acid to an amine
sumption electrochemical synthesis which avoided toxic by the expulsion of nitrogen was the key step for this trans-
reagents making the entire process environmentally benign formation via Schmidt’s reaction.
(Scheme 11).37 Saha reported an unconventional but efficient strategy for
An altogether different approach to accessing quinoxaline the synthesis of quinoxaline 4 employing 2-iodo benzoic acid
employing a catalytic quantity of p-TSA was demonstrated by 28a and sodium azide in the presence of organo-Cu(II) catalyst

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 13 Cu(II) mediated synthesis of quinoxaline.

of quinoxaline via the Schmidt reaction is highlighted in


Scheme 13.39 The conventional Schmidt reaction furnished
ortho-iodoaniline 32, which under the influence of Cu(II) cata-
lyst was transferred to o-PD 2 and subsequently reacted with
1,2-diketone 3 to furnish quinoxaline 4.
C. Oxidative dehydrogenation. The oxidative dehydrogena-
tion strategy is a one-step sustainable and atom-economical
route to access quinoxaline molecules from tetrahydroquinoxa-
Scheme 11 Electrooxidation of 3,4-dihydroxybenzoic acid 22 with N1, line 33 (Scheme 14).
N2-dibenzylethane-1,2-diamine 23, and its mechanistic pathway. Treatment of 1,2-dichloroethane with o-PD in TBACl at
60 °C is one of the simplest ways to access tetrahydroquinoxa-
line in a moderate yield of 56% (Scheme 15).40
A B(C6F5)3 catalyzed one-pot metal-free tandem sequential
procedure involving cyclization and hydrosilylation of imines
and amides towards the formation of 1,2,3,4-tetrahydroquinox-
alines directly from o-PDs and α-keto esters was reported by

Scheme 14 General schematic representation of oxidative dehydro-


genation reaction.

Scheme 12 Formation of quinoxaline from benzofurazans 25 and


ethanolamine 26 and the probable mechanistic pathway.

via Schmidt reaction followed by the nucleophilic substitution


reaction and subsequent cyclocondensation reaction with
65–70% yield (Scheme 13).39 The mechanism of the formation Scheme 15 Synthesis of tetrahydro quinoxaline.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

effective catalyst among the other screened catalysts like


H2WO4, Na2MoO4, (Mo)2(acac)2, VO(acac)2 and CeO2 which
furnished poor yields of hydroxamic acids. The oxidation
process was strongly affected by the solvent used and superior
yields were obtained in water (entry 1, Table 5).42
Vanadium pentoxide, being the most stable and frequently
used mixed-oxide catalyst, was shown to be effective in oxi-
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 16 B(C6F5)3 catalyzed synthesis of tetrahydroquinoxaline 36.


dative dehydrogenation (ODH) processes by Karki in 2013 for
the aromatization of tetrahydroquinoxaline in refluxing
toluene (entry 2, Table 5).43
Pan (Scheme 16).41 The scalability and non-requirement of Heterogeneous catalysts have mostly been used for the
anhydrous solvent or inert atmosphere are the advantages of dehydrogenation of N-heterocycles which require harsh con-
this method. ditions and show poor functional-group tolerance. The urge
Depending on the involvement of a hydrogen atom accep- for the development of easily available and inexpensive cata-
tor, the oxidative dehydrogenation of aza-heterocycles can be lysts for the oxidative dehydrogenation of N-heterocycles has
classified into two subclasses: acceptorless and acceptor- led to the development of iron-based catalyst 37 in 2014 which
assisted dehydrogenation. They can either be accessed by is inexpensive, abundant, and isoelectronic (d6 metal) with the
hydrogen acceptorless metal-mediated processes involving Ir(III) catalyst (entry 3, Table 5).44
metals like Fe, V, Co, Rh, Pd, etc., or via oxidative dehydrogena- Considering the disadvantages of nano-iron catalysts, like
tion involving oxidants leading to the complete aromatization poor stability and easy deactivation by aggregation and
of the moiety.42–68 leaching, led to the development of core–shell nanoparticles
A representative mechanism depicting the oxidative dehy- by Cui and co-workers in 2015. They developed nitrogen-
drogenation of tetrahydroquinoxaline to quinoxaline is rep- doped graphene shells immobilized on carbon supports
resented in Scheme 17.55 The electron transfer from the surrounding the nanostructured iron oxides for the
oxidant Cu(III)TPP to Cu(II)TPP resulted in the formation of an catalytic dehydrogenation of tetrahydroquinoxalines (entry 4,
amine radical species which then produced the imine inter- Table 5).45
mediate by α C–H abstraction, liberating water as the side In 2015, Iosub and co-workers initially employed a Pd-
product. After tautomerization, the final dehydrogenation based catalyst and got unsatisfactory results. Thereafter they
product was then obtained by the second dehydrogenation tested heterogeneous Co-oxide-based catalysts for the dehydro-
reaction (Scheme 17).55 genative aromatization of N-heterocycles. Low yield was
Several groups have demonstrated an oxidative dehydro- obtained for other heterogeneous Mn, Fe, and Ni oxide cata-
genation strategy to access quinoxaline from tetrahydroqui- lysts. However, the Co3O4-NGr/C catalyst exhibited superior
noxaline by employing diverse sets of catalysts. Therefore, performance in terms of product yield at 60 °C, lower than the
Table 5 highlights a list of several catalysts along with the temperatures typically required for dehydrogenation (entry 5,
corresponding yields for the oxidative dehydrogenation reac- Table 5).46
tion to access quinoxaline 1.42–68 By considering the same strategy, in 2015 Xu and coworkers
In 1990, Murahashi and colleagues employed sodium tung- employed the cobalt pincer catalyst 38 for dehydrogenation
state in 30% methanolic hydrogen peroxide solution under and hydrogenation of N-heterocycles (entry 6, Table 5).47
argon atm to synthesize quinoxaline in 37% yield by catalytic
oxidation of 1,2,3,4-tetrahydroquinoxaline at room tempera-
ture. The lower yield obtained was due to the formation of a
quinoline by-product. Na2WO4·2H2O was found to be the most

Although several groups have already developed excellent


aerobic oxidative systems, reactions under external ligand-
free conditions was considered challenging for the practical
and cost-efficient synthesis of N-heterocycles. Hence, in
2016 Wang and the group reported a Pd-catalyzed aerobic
oxidation reaction for the preparation of quinoxaline and
other N-heterocycles using potassium tert-butoxide. Other
screened inorganic bases gave poor yield and only starting
Scheme 17 Possible reaction pathway for oxidative dehydrogenation materials were recovered in most of the cases (entry 7,
reaction. Table 5).48

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Table 5 Reaction conditions for oxidative dehydrogenation of tetrahydroquinoxaline 33 to quinoxaline 1


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Entry Reaction conditions % yield

1 H2O2/Na2WO4, MeOH 3742


2 V2O5 (2.0 equiv.), SiO2, toluene, reflux 24 h 8843
V2O5 (0.2 equiv.), DTBP, SiO2, toluene, reflux 84 h 79
3 Fe-Pincer catalyst (3 mol%), xylene reflux, 30 h 10044
4 FeOx@NGr-C (0.18 mol% Fe), heptane, 15 bar air, 100 °C, 12 h 9345
5 Co3O4-NGr/C, K2CO3, O2, MeOH, 16 h, 60 °C 9746
6 Co-Pincer catalyst (20 mol%), p-xylene, 150 °C, 2 day 9347
7 Pd(CH3CN)2Cl2, KOtBu, O2, DCE, 120 °C, sealed tube 9848
8 FeCl2, DMSO, O2, 100 °C, p-xylene, 26 h 4649
9 Rose Bengal catalyst (1.0 mol%), DMA, visible light, R.T., air 7950
10 PdNPs/SBA-15 (5 mol%), toluene-d8, 130 °C, 23 h 9951
RhNPs/SBA-15 (5 mol%), toluene-d8, 130 °C, 23 h 99
PtNPs/SBA-15 (5 mol%), toluene-d8, 130 °C, 23 h 72
11 Pd-NPs (4% of substrate), TBHP(8.0 equiv.), water, 80 °C, 18 h 7552
12 GO(30 mg), Na2CO3(40 mol%), EtOAc, 120 °C, 48 h, air 8553
13 Meso MnOx (25 mg), DMF, 130 °C, 20 h, air >9954
14 CoTPP (10 mg), DMF, 110 °C, O2, 15 h 6855
15 NaClO2, NaOCl, and catalytic amounts of TEMPO, CH3CN, 0 °C 4856
16 Ru(II)-NNC pincer catalyst (2 mol%), o-xylene, 140 °C, 48 h 8457
17 ISAS-Co/OPNC (25 mg), mesitylene, 120 °C, 8 h 9958
18 Laccase(200U)/DTBC(20 mol%), O2, NaPBS, MeCN, 45 °C, 24 h 9559
Laccase (200 U)/TEMPO(20 mol%), O2, NaPBS, MeCN, 45 °C, 24 h 93
19 PMI catalyst (50 mg), water: MeOH, O2, 120 °C, 300 rps, 24 h 7360
20 Ni2Mn-LDH (80 mg), mesitylene, O2, 120 °C, 9 h 6961
21 Ni(II)/TiO2, 4-amino TEMPO, i-PrOH, O2, blue LED 453 nm 8362
22 2[PW]-OMS-2 (20 mg), DMC, O2, 80 °C, 6 h 9863
23 99% KOtBu (1.5 mmol), o-xylene, 0.1 MPa N2, 140 °C, 36 h 9464
24 Ir catalyst (0.2 mol%), H2, H2O, 120 °C, 17 h 9565
25 Co NCs/N–C (50 mg), MeOH, 50 °C, 12 h 9966
26 PDI-SN (10 mg), ambient air, DMA, blue-LED 400 nm 8667
27 NiBr2/1,10-phenanthroline, TEMPO, t-BuOK, t-AmOH, O2, 95 °C, 24–48 h up to 99%68

The necessity to design and develop a suitable catalytic junction with tert-butyl hydroperoxide (TBHP) as an oxidant
system that will be inexpensive as well as efficient led Zhou for the hydrogenation of quinoxalines in water under mild
and co-workers to develop a simple FeCl2/DMSO system that conditions (entry 11, Table 5).52
can effectively catalyze the oxidative dehydrogenation of
N-heterocycles employing molecular oxygen as the oxidant
(entry 8, Table 5).49
In 2017 Sahoo et al. reported a room temperature visible
light-mediated organo-photoredox catalyst for the oxidative
dehydrogenation of tetrahydroquinoxalines to quinoxaline for
the first time under base- and additive-free conditions with
ambient air (entry 9, Table 5).50
Nanoparticles of Pd, Pt, Rh stabilized by G4OH PAMAM
dendrimers and supported on SBA-15 (MNPs/SBA-15 with M = In 2017, Zhang and co-workers reported an environmentally
Pd, Pt, Rh) were shown to be an effective catalytic system by friendly method for the synthesis of quinoxaline by the oxi-
Deraedt and co-workers for the acceptor less oxidative dehy- dative dehydrogenation of tetrahydroquinoxalines employing
drogenation of tetrahydroquinoxaline to quinoxaline at 130 °C graphene oxide (GO) (entry 12, Table 5).53
in toluene. These developed catalysts were robust, active, air- Nevertheless, the search for a simple and efficient catalytic
stable and recyclable, and afforded products in very good system for oxidative dehydrogenation under aerobic conditions
yields (entry 10, Table 5).51 was highly desirable as most of the existing techniques
Sun and co-workers employed a palladium nanocatalyst required harsh reaction conditions, difficult catalyst prepa-
stabilized by carbon-metal covalent bonds (PdvNPs) in con- ration, and the use of high oxygen pressure and additives.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

Hence, Mullick and group54 developed mesoporous manga- store H2 employing either formic acid or external hydrogen as
nese oxide as an efficient catalytic system for environmentally the hydrogen source (entry 17, Table 5).58
benign oxidative dehydrogenation under aerobic, additive-free The use of the transition metals due to their cost as well as
conditions (entry 13, Table 5). the requirement of high reaction temperatures is often con-
Metalloporphyrins are typical biomimetic catalysts and sidered unsuitable for the thermally unstable substrates. This
were studied as a new catalytic system for the aerobic oxidative has led to the development of alternative non-metallic catalysts
dehydrogenation of N-heterocycles without any additives. that are sustainable for the aerobic oxidative synthesis of tetra-
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Hence, Zhou and coworkers developed an additive-free hydroquinoxalines. Hence Saadati came up with a new biocata-
method for the synthesis of quinoxaline by aerobic oxidative lytic approach employing Laccases ( p-benzenediol: oxygen
dehydrogenation of tetrahydroquinoxaline using cobalt por- oxidoreductase) for the oxidation of electron-rich aromatic
phyrin as a catalyst. The reaction mechanism suggested that substrates. The biocatalyst Laccases featured good thermal
the oxidative dehydrogenation probably proceeded with the stability, substrate selectivity, and the capability to catalyze the
imine formation and no high valent oxo-cobalt formation as reaction under mild reaction conditions (entry 18, Table 5).59
the active species (entry 14, Table 5).55 To bridge the gap between the homogeneous and hetero-
Chamorro-Arenas developed a protocol for the synthesis of geneous catalysts, Zhang and coworkers have prepared a single-
quinoxaline by using cheap and innocuous reagents, such as active site heterogeneous catalyst via the oxidative polymeriz-
NaClO2, NaOCl, and catalytic amounts of TEMPO (entry 15, ation of maleimide derivatives and used it for selective oxi-
Table 5), Scheme 18.56 The subsequent oxidative deallylation dation of heterocyclic compounds (entry 19, Table 5).60
furnished the quinoxaline albeit in moderate yield. During the development of efficient catalysts for aerobic
Acceptor less dehydrogenation is favored over the tra- oxidation, Zhou demonstrated the use of NiMn-hydrotalcite as
ditional method because such processes are atom-economical an effective catalyst for the oxidative dehydrogenation of tetra-
and environmentally benign and avoid the use of stoichio- hydroquinoxaline by molecular oxygen under mild conditions
metric oxidants. Wang and coworkers in the year 2018 prepared (entry 20, Table 5).61
a highly active transition metal complex of ruthenium(II) TiO2 has been known to be a widely used heterogeneous
hydride bearing a pyrazolyl-(2-indol-1-yl)-pyridine ligand photocatalyst due to its outstanding stability and nontoxicity
(ruthenium(II)-NNC pincer complex) 41 which exhibited high as well as its reusability. However, to improve the photo-
catalytic activity and broad substrate scope for the synthesis of catalytic activity of TiO2, Balayeva and co-workers employed
N-heterocycles and secondary alcohols via an acceptor less transition-metal nanoparticles or ions grafted onto the TiO2
dehydrogenation strategy (entry 16, Table 5).57 surface and demonstrated a visible-light-mediated dehydro-
genation of N-heterocycles (entry 21, Table 5).62
Bi and co-workers prepared recyclable OMS-2-based nano-
composites doped with Mn2O3 for the catalytic oxidative dehy-
drogenation of N-heterocycles in excellent yield (entry 22,
Table 5).63
Since many of the existing dehydrogenation protocols were
often affluent because of the involvement of expensive tran-
sition metals as well as additives for the reaction and were not
environmentally benign, Liu reported a transition metal-free,
acceptor less protocol for the dehydrogenation of
Single-atom catalysts (SACs) attracted extensive concern
N-heterocycles to access quinoxalines and allied heterocycles
because of their maximum atom utilization, superior reactivity
employing potassium tert-butoxide and molecular hydrogen
and selectivity and became potential substitutes for homo-
(entry 23, Table 5).64
geneous catalysts. Porous N-doped carbon catalysts with Co
Wang and co-workers developed an additive-free dehydro-
atomic sites (ISAS-Co/OPNC) were used by the group of Han
genation of N-heterocycles in water employing an iridium-
for the highly efficient dehydrogenation of tetrahydroquinoxa-
based catalyst bearing an electron-enriched dipyridylamine
line. The dehydrogenation proceeds with the release of H2 and
ligand 42, under mild conditions (entry 24, Table 5).65
the reverse transfer hydrogenation of tetrahydroquinoxaline to

Scheme 18 TEMPO-catalyzed selective dual C–H oxidation.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Since non-noble metals frequently get oxidated and leached


under harsh reaction conditions, the key concern to replace
the noble metals with non-noble metals to achieve the
superior catalytic activity, as well as sustainability, was towards
developing a robust catalytic system that will be sustainable as
well. Wu and co-workers developed a bifunctional catalyst,
namely Co NCs/N–C (cobalt nanocrystals stabilized by nitro-
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

gen-doped graphitized carbon), which can catalyze the oxi-


dative dehydrogenation with methanol as a solvent at 50 °C
within 4 h (entry 25, Table 5).66
Recently, Yu and coworkers developed a visible light-
mediated, additive-free, efficient reusable photocatalytic
system for the oxidative dehydration reaction utilizing
perylene diimide (PDI), Fig. 5.67 The formation of tiny
PDI semiconducting material because of the aggregation of
the PDI molecules on the SiO2 surface was responsible for the
high-efficiency photo-oxidative activity (entry 26, Table 5).67
Scheme 19 Selective C–H oxidative sulfonylation of tetrahydro-qui-
noxalines; and the mechanistic pathway.

dehydroaromatization furnished the final product 48,


Scheme 19.69

Using epoxides
Murata and coworkers demonstrated the use of epoxy
aldehydes or epoxy esters to access dihydroquinoxaline
derivatives.
They stated that o-PD 2 on reaction with 2,3-epoxy alde-
hyde 49 furnished 2-hydroxyalkyl-1,2-dihydro quinoxaline
50, whereas 2,3-epoxy esters 51 delivered either dihydro-
Fig. 5 PDIs in different phase states. quinoxaline derivatives 52 or uncyclized hydroxylamine
ester 51a. The reaction of epoxy aldehyde and o-PD initially
resulted in the formation of Schiff’s base which then
Bera et al. developed an efficient and selective nickel cata- underwent intramolecular nucleophilic ring-opening by
lyzed method for dehydrogenation of the six-membered epoxide resulting in the formation of dihydroquinoxaline
N-heterocycles via the application of molecular oxygen to derivatives 50. Whereas ring-opening reactions of epoxy
access quinoxaline and allied heterocycles (entry 27, Table 5).68 ester proceeded only in the presence of LiOH catalysts to
The selective introduction of the sulfonyl group into the activate epoxide oxygen to furnish the desired product,
organic molecules enhances the hydrophilicity and biologi- Scheme 20.70
cal affinity of the entire molecule. Therefore, there is a great
deal of interest in developing sulfonylated heterocycles.
Oxidative sulfonation of tetrahydroquinoxaline was demon-
strated by Xie and co-workers employing highly dispersed
cobalt catalysts using sodium sulfinate. This methodology
allowed access to a variety of sulfonyl quinoxalines with
high regio- and chemoselectivity with very good functional
group tolerance, Scheme 19.69 The mechanistic detail dis-
closes that the coupling between tetrahydroquinoxaline and
sodium sulfide occurs in the presence of iodine which
worked as a radical initiator to generate the N-radical via
single-electron oxidation (SEO). Subsequent deprotonation
followed by isomerization furnished the intermediate aryl
radical 33b. Simultaneously, an iodide radical initiated the
formation of the electrophilic sulfonyl radical 47b via the
SEO of sulfinate. Thereafter, radical coupling followed by Scheme 20 Synthesis of 1,2-dihydro quinoxaline using epoxides.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

In 2002, Antoniotti and co-workers employed a catalytic


amount of Bi(0) powder and triflic acid or copper triflate to
prepare substituted quinoxalines by reacting epoxides with
o-PD derivatives under molecular oxygen in DMSO at 100 °C.
For mono-substituted epoxides Cu(OTf )2 was used as an addi-
tive, whereas for di-substituted epoxide a strong Brønsted acid
such as triflic acid was used as an additive. This reaction pro-
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

ceeded via the oxidative ring opening of oxirane by DMSO in


the presence of an acid catalyst to generate α-hydroxy ketone.
Second oxidation of the alpha-hydroxy ketone in the presence Scheme 23 Palladium-catalyzed synthesis of 3,4-dihydro-1,4-benzox-
of Bi(0)/O2 system or a Bi(III)/Bi(0) redox process resulted in azine-1,2,3,4-tetrahydroquinoxaline derivatives from epoxides.

the formation of α-diketone which then underwent a double


condensation reaction with the diamine to furnish the
required quinoxaline, Scheme 21.71 ceeding via ring-opening of oxirane. The use of Cs2CO3 with a
In 1980, Taylor and co-workers utilized α-substituted epox- simple operational method and the use of inexpensive starting
ides as bifunctional two-carbon synthons to synthesize qui- materials, avoiding the use of oxidizing agents, are the advan-
noxaline by condensing with o-PDs 2 in DMF solvent. tages of this procedure. The authors described the mechanistic
The heterocyclization of α-cyano epoxide 54 with o-PD 2 pathway where the reaction proceeded via the formation
resulted in achieving quinoxaline in up to 94% yield. of imine 59a. Subsequent nucleophilic addition, epoxide
Mechanistic pathways suggested that nucleophilic attack on ring-opening, and dehydration resulted in the formation of
α-cyano epoxides 54 occurred at the C-1 carbon atom rather a six-membered intermediate 59b. A 1,3-prototropic shift
than C-2. Dehydration of the benzylic alcohol followed by finally afforded the desired quinoxaline 60 in good yield,
intramolecular conjugate addition resulted in the desired qui- Scheme 24.74
noxaline 55 in good yields, Scheme 22.72 Nitroepoxides were also used as precursors for the synthesis
Vinyl epoxides are known to furnish 1,4 addition products of quinoxalines as they can be easily accessed by the one-step
through a nucleophilic opening of a π-allyl palladium complex. epoxidation of nitroalkenes. Vidal-Albalat et al. employed such
However, altering the reagents can lead to the formation of a strategy to synthesize quinoxalines from nitroepoxides using
1,2-addition products intramolecularly. Capitalizing on this ethanol as an environmentally benign solvent, Scheme 25.75
concept, Srikanth and co-workers demonstrated the synthesis A mechanistic pathway suggested the formation of
of quinoxaline via a palladium-catalyzed stereoselective α-iminoamine intermediate 61a which then afforded the qui-
opening of cis and trans-vinyl epoxides 56 with an amine noxalines 62 on treatment with air, whereas the in situ
nucleophile in the presence of boron reagent 56a. The role of addition of reducing agents resulted in the formation of tetra-
the boron was to form a stable adduct with the o-PD and assist hydroquinoxaline 62a.
the intramolecular Michael addition to access densely functio- In 2013, Ibrahim et al. developed a synthetic route to access
nalized chiral 1,2,3,4-tetrahydroquinoxaline derivatives in high functionalized quinoxaline using α-nitroepoxides. The one-pot
yield, Scheme 23.73 approach involving an environmentally friendly metal-free
Tu and co-workers developed a domino strategy to syn- strategy employing easily accessible starting compounds made
thesize quinoxaline derivatives by [4 + 2] heterocyclization pro-

Scheme 21 Bi(0) catalyzed synthesis of substituted quinoxaline.

Scheme 24 [4 + 2] heterocyclization towards the synthesis of quinoxa-


Scheme 22 Synthesis of 2-aryl-3-cyano-1,2,3,4-tetrahydroquinoxaline. line derivatives and the probable mechanistic pathway.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 27 Formation of fluorinated quinoxaline from


perfluoroepoxyoxolanes.

Scheme 28 The reaction of epoxide derivatives with o-PDs.

Scheme 25 Synthesis of quinoxaline from α-nitro epoxide and


mechanistic pathway for the formation of quinoxaline derivatives.
23–67% yield in the presence of aprotic solvent (dioxane),
Scheme 28.78

this protocol highly attractive and advantageous over the exist- Reactivity and reactions
ing methods. Mechanistic studies revealed that the attack of Electrophilic attack on quinoxaline is less likely to occur
the diamine 58 on the α-nitroepoxide 63 on the less sterically because of the presence of N-atoms making it electron
hindered side leads to the epoxide ring-opening followed by deficient. However, the nucleophilic attack can occur at the 2,
intramolecular nucleophilic attack by the second amine group. 4, 5, and 8a positions. Over the years, quinoxaline moiety has
Subsequent ring closure accompanied by the release of nitrous gained tremendous importance due to its potential application
acid, along with dehydration and aromatization, furnished the in dyes, pharmaceuticals, and advanced functional materials
desired quinoxaline 64, Scheme 26.76 and as a result, a rational design approach is very much essen-
Perfluoroolefin oxides, being an important and versatile tial. So here we will be documenting the different aspects of
synthon, were used for the synthesis of polyfunctional quinox- the functionalization of quinoxaline moiety.
alines containing fluorine atoms and perfluoroalkyl groups,
Scheme 27.77 Halogenation and dehalogenation
In 2006, Saloutina and group employed epoxy derivatives of Catalytic condensation of an aromatic o-PD with α-dialdehydes
internal perfluoro olefine 68 to synthesize quinoxaline 4 in (or α-diketones) in the presence of either a Lewis or a Brønsted
acid has been one of the most practiced methods for the syn-
thesis of quinoxalines because of the utilization of commer-
cially available reagents under ambient experimental con-
ditions. However, the synthetic route lacks diversity as the
existing methodologies do not leave us with many choices to
diversify the groups attached at the C2 and C3 positions.
Moreover, the challenges associated with bringing diketone
derivatives with structural complexity can be overcome by late-
stage functionalization of the quinoxaline scaffold.
Therefore, to overcome such limitations, halogenations to
access mono or di-halo substituted quinoxalines appear to be
a viable solution to access diverse quinoxaline derivatives. The
π electron-deficient character of the quinoxaline allowed facile
palladium-mediated oxidative addition reactions without invol-
ving any expensive ligands. Halo-substituted quinoxaline 70a/b
offered a perfect template for a wide range of cross-coupling
reactions, Scheme 29, and Fig. 6, and provided the
Scheme 26 Formation of quinoxaline from α-nitro epoxides and molecule with a perfect handle for structural and functional
mechanism for the formation of quinoxaline from α-nitroepoxide. diversification.6

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Fig. 7 Brominated products of quinoxaline under different conditions.

Table 6. However, C2 and C3 brominated quinoxalines can


Scheme 29 Schematic representation of several coupling reactions for
the functionalization of the quinoxaline scaffold.
only be accessed when the bromination of the quinoxaline
N-oxides is carried out. Quinoxaline N-oxides which can be
accessed from quinoxaline provided a unique platform for the
bromination at the electron-deficient ring of quinoxaline, as
evident from the different conditions highlighted in
Scheme 30.79 Bromination of N-oxides furnishes the bromi-
nated product 89 and 92 with bromination occurring at the
nitrogenous ring of the quinoxaline, Scheme 31.79
Directed lithiation of aromatics and heteroaromatics consti-
tuted an important strategy for ring functionalization.
However, the competitive nucleophilic addition reaction
makes the metalation of diazines even more challenging. A
highly active and soluble magnesium base (TMP)2Mg·2LiCl
Fig. 6 Scope of nucleophilic aromatic substitution of quinoxaline.
(TMP = 2,2,6,6-tetramethylpiperamidyl) 93 smoothly allows the
metallations of quinoxaline with excellent functional group
compatibility. Dong demonstrated the use of (TMP)2Mg·2LiCl
The importance of 2,3-dichloroquinoxaline (DCQX) 79 as a 93 for the synthesis of zincated quinoxaline 94b which was
synthetic precursor has been documented recently in 2020 by further converted to iodo-substituted quinoxaline 94 in excel-
Menezes and colleagues highlighting the application of this lent yields on treatment with nucleophilic iodine,
scaffold towards the synthesis of diverse sets of functionally Scheme 32.80
and medicinally important molecules.7 In 2011, Blair and co-workers further synthesized 2-iodo
The halo substituted quinoxaline 70a/b featuring C2 and C3 quinoxaline 94 demonstrating the chemoselective deprotona-
labile protons are exact matches for the metallation employing tive zincation reactions of a range of quinoxaline using lithium
lithium or magnesium. The corresponding organometallic inter- zincate base [Li(TMP)Zn(tBu)2] 93 followed by treatment with
mediate 71 can be easily converted into a functionalized deriva- nucleophilic iodine in 50% yield, Scheme 32.81
tive by reaction with an appropriate electrophile, Scheme 29.7 The mechanistic details are highlighted in Scheme 33. In
The bromination of quinoxalines often resulted in the for- pathway a, the formation of manganese-containing hetero-
mation of N-bromo complexes. However, precise control over cyclic species 94a from the reaction between quinoxaline 1 and
the reaction parameters like solvent, temperature and light, as base (TMP)2Mg·2LiCl was proposed which then undergoes fast
well as the equivalents of brominating agents, will allow the transmetallation with zinc reagent to provide corresponding
formation of several brominated derivatives as highlighted in zinc-containing heterocycle 94b. In pathway b, the base
Fig. 7 and Table 6.79 The conditions under which different (TMP)2Mg·2LiCl first reacts with zinc reagent and the formed
bromo derivatives of quinoxaline are obtained are illustrated species further reacted with quinoxaline to yield the zinc-con-
in Table 6.79 taining heterocycle 94b. In pathway c, the heterocycle first
The direct bromination of quinoxaline results in the bromi- reacted with zinc reagent forming an intermediate zinc
nation on the benzene ring of the quinoxaline, Fig. 7, and complex 94c which then underwent fast transmetallation in

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Table 6 Bromination of quinoxaline under different conditions with brominating agents

Entry Reaction conditions % yield

1 Br2 (6.6 equiv.), AcOH, 118 °C, 24 h B = 46%, E = 11%


2 Br2 (12 equiv.), H2SO4, 200 °C, 40 h B = 28%, E = 9%
3 Br2 (8 equiv.), MeCN, 82 °C, 25 h D = 75%
4 Br2 (8 equiv.)/BaCO3 (2 eq.), MeCN, 82 °C, 25 h B = 31%, C = 14%, D = 5%, E = 6%, F = 28%
5 Br2 (3 equiv.)/BaCO3 (2 eq.), MeCN, 82 °C, 72 h B = 65%
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

6 Br2 (8 equiv.), CCl4, hν (150 W projector lamp), 18 h D = 30%, F = 25%


7 Br2 (2 equiv.), CCl4, hν (150 W projector lamp), 45 h A = 70%, C = 3%, D = 10%
8 NBS (6 equiv.)/BPO (benzoyl peroxide), AcOH, 118 °C, 20 h B = 50%
9 NBS (6 equiv.)/BPO (benzoyl peroxide), DMF, 153 °C, 18 h B = 51%
10 NBS (7 equiv.)/BPO (benzoyl peroxide), chloroform, 62 °C, 45 h B = 29%, C = 29%, D = 3%, E = 10%
11 NBS (6 equiv.)/BPO (benzoyl peroxide), MeCN, 82 °C, 20 h B = 55%, C = 15%, E = 5%

Scheme 30 Synthesis of brominated quinoxaline from quinoxaline N-oxide.

the presence of a base to yield the corresponding zinc deriva-


tive, Scheme 33.
Fluorinated heterocycles are prevalent in pharmaceuticals,
agrochemicals, and materials.82 The selective introduction of
fluorine into small molecules often results in the alteration of
the acid dissociation constant and molecular conformation
due to the origin of non-covalent interactions resulting in
effects on the solubility, as well as the stability, compared with
the non-fluorinated counterpart.
In 1999, the research group of Chambers synthesized
2-fluoro- or 2,3-difluoro-quinoxaline 96 by treating quinoxaline
1 with a mixture of fluorine and iodine in the presence of tri-
ethylamine. The formation of poly-fluorinated products was
Scheme 31 Bromination of quinoxaline N-oxide under different dependent on the quantity of fluorine used in the reaction,
conditions. Scheme 34.83 Formation of iodonium and fluoride ion sources

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

Table 7 Reaction conditions for dehalogenation of quinoxaline

Entry Reaction conditions % yield


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

1 Indium powder, water, reflux 4485,


5686
2 Pd(OAc)2 (5 mol%), PPh3 (20 mol%), NaBH4 (1.7 7287
equiv.), TMEDA (1.7 equiv.), THF, 25 °C, 1 h
PdCl2(dppf) (5 mol%), NaBH4 (1.7 equiv.), TMEDA 5287
(1.7 equiv.), THF, 25 °C, 1 h
3 Si-325 mesh, Co(OAc)2 (20 mol%), PPh3 (22 mol%), 8288
NaOH (130 mol%), TBAB (4 equiv.), H2O, 100 °C,
24 h

Scheme 32 Iodination of quinoxaline via the alkali–metal mediated


zincation.
Similarly, Fier developed site-selective fluorination of a
single carbon–hydrogen bond in diazines using commercially
available silver(II) fluoride, Scheme 34.84

Dehalogenation
Several dehalogenative methods to access quinoxaline from
2-halo substituted quinoxalines 1 were reported by several
research groups and are highlighted in Table 7.85–88 The deha-
logenation step constitutes an important chemical transform-
ation in terms of accessing selective functionalization at either
C2 or C3 positions of quinoxaline.
Environmentally benign, facile, and safe protocols avoiding
standard dehalogenative reagents like Pd–C, RANEY® Ni, or
halogen–metal exchange were reported by Hirasawa85 and
Fukuda.86 They have employed indium for the dehalogenation
of aromatic halides in water (entry 1, Table 7).85,86 Indium
metal in water was reported to furnish quinoxaline from
2-haloquinoxaline in moderate yield.
Chelucci and co-workers employed a combination of
Scheme 33 Postulated intermediates of the metalation of quinoxaline NaBH4-TMEDA together with a palladium catalyst for the dehy-
with (TMP)2Mg·2LiCl in the presence of ZnCl2. drohalogenation of halo substituted quinoxalines, as well as
other allied heterocycles. NaBH4-TMEDA was used as a hydride
source. Pd(OAc)2-PPh3 rapidly hydrodehalogenated the bromo-
heteroarenes whereas PdCl2(dppf ) and PdCl2(tbpf ) reduced
chlorinated heteroarenes. The reaction was compatible with a
wide variety of functional groups and offered high chemo-
selectivity in the presence of functional groups like ester,
alkyne, alkene, and nitrile (entry 2, Table 7).87
The Pd-based conditions were not effective for hydrodehalo-
genation of aryl halides because of the formation of biaryls
along with the dehydrohalogenation products making the sep-
aration tedious. A novel catalytic system comprising Co(OAc)2/
Scheme 34 Selective fluorination of quinoxaline.
PPh3 was reported by Gevorgyan and coworkers which can
effectively catalyze the dehydrohalogenation with no formation
of side products and unreacted starting materials. The authors
reported that the utilization of 130% of the base to neutralize
was observed in situ when iodine was reacted with fluorine fol- the acid generated during the reaction was beneficial to
lowed by the fluorination of intermediate N-iodo-heterocycle improve the yield of the reaction. The reaction was substrate-
by fluoride ions. dependent and found to work efficiently with electron-

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

deficient hetero-arenes or arenes attached with electron-with- quinoxaline was possible by precisely controlling the basicity
drawing groups rather than with electron-rich species (entry 3, and lipophilicity balance between the quinoxaline and the
Table 7).88 monoacylation products in a bi-phasic system. Decreasing the
basicity of heteroarenes was the key to improving the yield of
C–C bond formation via Minisci reaction (acylation) the monoacylated product 100. The mechanistic pathway is
highlighted in Scheme 36 and the reaction follows the redox
Other than the halogenations followed by coupling reactions,
pathway forming the acyl radical as illustrated below.89
direct functionalization of the C2 and C3 centers is a well-
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Chaubey developed a green protocol for the acylation of het-


explored strategy to get diversely substituted quinoxalines.
erocycles by using water as a green solvent for the reaction.
Acylation of electron-deficient heteroarenes is highly chal-
The protocol avoided the use of hazardous organic solvents as
lenging knowing the fact that conventional Friedel–Crafts acy-
well as performing under metal and additive-free conditions,
lation will be highly improbable with electron-deficient
Scheme 37.90 The reaction of N-heterocycles 102 and α-keto
species. Direct acylation of electron-deficient quinoxalines,
although challenging, has been widely explored because it pro-
vides a handle for further synthetic modifications.
The Minisci reaction is the insertion of C-centered radicals
to aromatic heterocycles followed by the rearomatization of the
ring with the loss of hydrogen atoms. The reaction has gained
extensive importance since its inception due to its ability to
directly functionalize the heteroaromatics rapidly.
The popularity of the Minisci reaction due to simple reac-
tion conditions has resulted in several applications in hetero-
cyclic chemistry including the acylation of quinoxalines. The
direct acylation of quinoxaline employing α-keto acids has
been carried out under different reaction conditions as illus-
trated in Scheme 35.89–92
Selective acylation of quinoxaline 1 by the silver catalyzed
decarboxylation of keto acids by persulfate was described by
Fontana in 1991, Scheme 35.89
Although multiple positions of high nucleophilic activity
are available in the quinoxaline, selective monoacylation of Scheme 36 Redox chain process of acylation reaction.

Scheme 35 Direct acylation of quinoxaline 1, with α-ketoacid.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

of radicals from the light absorption of the α–keto acid 99.


Second, by the formation of the electron donor–acceptor (EDA)
complex A, resulting from the photodecarboxylation process of
absorbing the visible light, and third, by the generation of the
EDA complex B resulting from the homolytic cleavage of
S2O82− to produce SO4•− which triggered the decarboxylation
of the α-keto acid. Formation of the acyl radical was followed
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

by the addition of the protonated heteroarene and subsequent


hydrogen atom transfer (HAT) and re-aromatization delivered
the acylated N-heterocycles, Scheme 39.92

Scheme 37 Metal-free acylation of quinoxaline. Cross-dehydrogenative coupling


A cross dehydrogenative coupling strategy has become one of
the most effective approaches for constructing C–C bonds
acids 99 was catalyzed by K2S2O8 and followed the radical directly from two unmodified C–H, C–C, or C–N bonds utiliz-
pathway as demonstrated in Scheme 37.90 ing oxidizing agents like TBHP, Scheme 40.93 The advantage of
Wang and co-workers used FeSO4 catalyst instead of expen- the CDC approach is that direct tandem oxidation of simple
sive Ag(I) salt as a catalyst for the Minisci type acylation reac-
tion, Scheme 37.91 A quinoxaline-bearing electron-donating
group furnishes a good yield of an acylated product compared
with the electron-withdrawing group, which resulted in a very
low yield along with unreacted starting material. The reaction
followed a pathway where persulfate oxidizes Fe(II) to Fe(III)
resulting in the decarboxylation of α-keto acid 99 to generate
acyl radicals. An alternative pathway stating the generation of
methyl radicals by Fe(II)/DMSO/S2O82− oxidation systems was
also proposed which subsequently afforded the acyl radical by
the abstraction of the hydrogen atom. The acyl radical 99a
then underwent radical addition with the protonated heteroar-
enes 102, and further oxidation by Fe(III) or persulphate
resulted in the desired acylated product 100, Scheme 38.91
Visible-light mediated reactions are known to occur under
extremely mild conditions and avoid the use of photosensiti-
zers. As a result, the photoinduced reaction pathway provided
a new direction towards sustainable development. Therefore, a
photo-induced pathway was explained by Guillemard as a sus-
tainable approach for the acylation of quinoxaline,
Scheme 35.92
Three different mechanistic pathways were suggested for
the generation of an acyl radical. First, by the direct generation

Scheme 38 Possible mechanistic pathways for the Minisci acylation Scheme 39 Mechanistic pathway for the acylation of quinoxaline by
reaction using Fe(II)/DMSO/S2O82− oxidation system. radical functionalization.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Scheme 40 General schematic representation for the cross dehydro-


genation coupling reaction.
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

C–H and N–H bonds allows the use of simple and readily avail-
able reagents and reduces the reaction steps.
The construction of carbon–carbon bonds between two
hydrocarbons via oxidative cross-coupling is considered one of Scheme 42 Proposed mechanisms for the oxidative CDC reaction.
the most cherished methodologies in organic chemistry.
However, some challenges like low reactivity of some C–H
bonds, overoxidation, selective C–H bond functionalization,
dimerization, etc. are still underexplored. A series of CDC strat-
egies for the construction of acylated quinoxaline has been
reported by the groups of Liu, Prabhu and Lei and are high-
lighted in Scheme 41.94–96
An efficient metal-free oxidative cross-dehydrogenative
coupling (CDC) of quinoxaline, as well as allied N-heterocycles
with a diverse series of aromatic aldehydes, was demonstrated
by Liu and co-workers in the presence of TBHP/TFA. Desired
mono substituted product 107 was obtained as the major com-
ponent with yields up to 70%, Scheme 41.94 The mechanistic
pathway illustrated in Scheme 42 shows that the benzaldehyde
gets converted to phenacyl radical 99d in the presence of
TBHP at 110 °C followed by the addition of the iminium ion to
generate the intermediate 109b. Subsequently, this generates Scheme 43 Visible-light mediated oxidative C–H acylation of
the desired product by a single electron transfer mechanism, quinoxaline.
Scheme 42.94
Prabhu and co-workers replaced the TFA used in the earlier
reported reaction by Liu with NCS (N-chlorosuccinamide) as
reactions between quinoxaline and an aldehyde avoiding the
an additive and received 53% overall yield comprising 42%
use of any photocatalyst and furnished mono acylated deriva-
mono substituted derivative 107 and 11% di substituted
tive 100 in 48% yield, Scheme 41.96 The probable mechanistic
derivative 108, Scheme 43.95
pathway is highlighted in Scheme 43 which indicates the
Visible light-mediated photoredox catalysis has evolved as a
initial formation of a complex of heteroarene and t-BuOOH
powerful tool for creating new chemical bonds under mild
102a. The resulting complex 102a promoted the light-mediated
reaction conditions. Zhang developed a direct cross-coupling
hemolytic cleavage pathway to produce the tert-butoxy radical
method for acylation of quinoxaline by visible-light mediated
and hydroxyl radical. The tert-butoxy radical or hydroxyl
radical then undergoes a hydrogen atom transfer with benz-
aldehyde to generate the phenacyl radical 99d. The resulting
phenacyl radical 99d undergoes nucleophilic addition on the
protonated heteroarene to generate the corresponding radical
cation 102d. Subsequent single electron transfer from the
intermediate afforded the desired product 107.96
The hunt for a sustainable CDC pathway has led several
research groups to try out innovative approaches for the acyla-
tion of electron-withdrawing heteroarenes. Such is the case for
the research groups of Sharma and Sultan who have used ionic
liquid and CFL light respectively to access acylated quinoxaline
113, Scheme 44.
Sharma and co-workers developed an ionic liquid-mediated
efficient strategy to access C-benzoylated quinoxalines in
Scheme 41 CDC strategy for the construction of acylated quinoxaline. 75–80% yield via cross-dehydrogenative coupling (CDC) of

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 44 Acylation of quinoxaline.

benzylamines 111 with quinoxalines 1. The benzoylation of


N-heterocycles occurred via (NH4)2S2O8 catalyzed benzoyl Scheme 46 The mechanism for photoredox-catalyzed acylation of
radical formation, Scheme 44.97 The ionic liquid can be heteroarenes.
recycled up to 5 times and the reaction has broad substrate
scope encompassing both electron-donating as well as elec-
tron-withdrawing groups. However, the yields reported for the
electron-donating groups were slightly lower than those of mation of acyl radicals. Stoichiometric use of acyl precursors
electron-withdrawing groups.97 and shorter reaction times are some of the advantages of this
The mechanism highlighted in Scheme 45 describes that at protocol. The use of blue LEDs resulted in lower yields and
the initiation of the reaction, benzylamine 111 transforms to longer reaction times compared with the CFL, probably
benzaldehyde 109 via aerial oxidation. Subsequently, because of the faster oxidative cleavage of the phenylacetylene
ammonium persulfate generates the benzoyl radical 109a 112 to acyl radicals and ultimately leading to benzaldehyde as
which in turn attacks the C2 position of quinoxaline BF4− the by-product.
complex 102 and finally, the aromatization results in the final Therefore, the reaction was carried out in the presence of
product 107. The [Bmim]BF4 helped in the radical formation, CFL. The photoexcitation of Ru(bpy)3Cl2 leads to the single
Scheme 45.97 electron transfer in the presence of persulfate salt leading to
Sultan et al. utilized terminal alkynes 112 as an acyl source the formation of the radical anion (SO4•−) and Ru(III).
for the photoredox-mediated acylation reaction of electron- Meanwhile, the decarboxylation of the phenylglyoxalic acid 99
deficient quinoxaline and achieved 78% yield of the monoacy- formed in situ from phenylacetylene 112 followed by hydrogen
lated product 107, Scheme 46.98 The method relied on oxi- atom transfer (HAT) with sulfate radical anion (SO4•−) furn-
dative cleavage of phenylacetylenes as the key step for the for- ished the acyl radical, which subsequently coupled with the
protonated heteroarene 102 to provide the intermediate 102a.
A SET reaction between the intermediate and Ru(III) completed
the photoredox catalytic cycle and resulted in the formation of
the acylated quinoxaline 107 as shown in the mechanistic
pathway, Scheme 46.98
A nickel catalyzed electrochemical method for the direct
acylation of quinoxaline employing α-keto acids 99 was
reported by Ding and co-workers in 2020. The reaction was per-
formed under constant current conditions and avoided the use
of excess oxidant as well as silver catalysts, Scheme 47.99
The mechanistic pathway illustrated the oxidation of Ni
(acac)2 at the anode in the presence of α-keto carboxylic acid
99 to generate a nickel(III) complex 114. Subsequent ligand-to-
metal electron transfer resulted in the generation of ArCOCOO
radicals along with the regeneration of Ni(acac)2 which, being
Scheme 45 Probable mechanistic pathway for direct metal-free ben- quite unstable, extruded CO2 to generate the acyl radical 109b.
zoylation of the quinoxaline. The radical addition to the protonated quinoxaline afforded

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Scheme 47 Nickel-catalyzed electrochemical Minisci acylation


reaction.
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

the adduct 114c. Subsequent oxidation and deprotonation


furnished the acylated product 100 followed by the cathodic
reduction of the proton to afford hydrogen, Scheme 48.

Alkylation Scheme 49 Direct alkylation of quinoxaline.


Alkylation of quinoxaline is yet another way of constructing C–
C bonds at the C2 or C3 positions. Alkyl groups normally get
substituted or added to the molecules as carbocations, carba-
nions, radicals, or carbenes. However, alkyl radicals have been
the prime source for the alkylation of quinoxaline.
Due to the electron-donating nature of the phenylalkynyl
attached at the 2-position of the quinoxaline moiety, a nucleo-
philic substitution reaction is unlikely to occur. The highly
delocalized pi-electrons from the phenyl ring and the acetyl
moiety make the –N=C–H site less electrophilic. Therefore,
very strong nucleophiles (n-BuLi has been used for the alkyl-
ation of quinoxaline) can be the solution, as depicted in
Scheme 49.100
The Minisci reaction, due to its high versatility, has become
a more and more attractive strategy for medicinal chemists
and appears as a frequent synthetic tool in the synthesis of
biologically active molecules, Scheme 50.92
The alkylation of quinoxaline via the cross dehydrogenative
coupling turned out to be one of the straightforward routes to
access alkyl-substituted quinoxalines. Although challenging,
diverse substrates like acid anhydrides 125, alpha-keto acids Scheme 50 General mechanism of Minisci reaction.
119 S-tert-butyl o-ethyl xanthates 126, aliphatic acids 119, and
aryldifluoro acetic acid 127 have efficiently reacted under
different reaction conditions to furnish alkylated quinoxalines internal oxidant is the main advantage of this methodology.
4, 128–129 in good to excellent yields as illustrated in The alkyl-substituted quinoxaline 4 was obtained in yields up
Scheme 51.101–103 to 75%, Scheme 51.101 Zeng et al. proposed a reaction mecha-
The direct C–H alkylation strategy of quinoxalines under nism portrayed in Scheme 52. Initial protonation of the qui-
Brønsted acid-catalyzed conditions (TfOH) employing alkyl noxaline followed by the attack of the alkyl radical on the pro-
peroxide 125 acting both as an alkylating reagent as well as an tonated quinoxaline 102 leads to the amino-radical cation
102a. Subsequent oxidation of the cation radical intermediate
102a generates the alkylated product 4 after workup.
Revil-Baudard and coworkers demonstrated the use of
S-tert-butyl o-ethyl xanthate 126 as a source of the tert-butyl
group for the easy functionalization of quinoxaline as well as
other allied N-heterocycles, Scheme 51.102 The presence of
dilauroyl peroxide (DLP) became instrumental in the gene-
ration of tertiary alkyl radicals involved in the alkylation
process. The detailed mechanistic pathway is highlighted in
Scheme 53.102
Yet another mild methodology for the direct alkylation of
Scheme 48 Plausible mechanistic pathway for the nickel catalyzed quinoxaline 1 employing carboxylic acid 119 as the alkylating
electrochemical acylation of quinoxaline. source was reported by Galloway and co-workers in 2017,

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

Owing to the unique stability, isosteric properties, and lipo-


philic hydrogen-bond donor ability of the fluoride atoms, the
CF2 unit has gained importance in medicinal chemistry.
However, the installation of a difluoromethyl unit into a het-
eroaromatic scaffold is quite challenging. Therefore, to access
CF2 appended heteroarenes, Xie developed a mild silver-
promoted decarboxylative difluoro methyl arylation employing
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

aryldifluoro acetic acid 127 as difluoromethyl sources,


Scheme 51.104 An excellent yield of difluoromethylated qui-
noxaline 129 was obtained under the operational conditions.
Interestingly, another method for the difluoromethylation
of quinoxalines and other related heterocycles has been
reported by Lytkina and coworkers, employing cheap and
readily accessible sodium difluoromethane sulfinate,
Scheme 54.105 Sodium difluoromethane sulfinate is con-
sidered a cheap and promising alternative to zinc(II) difluoro-
methane sulfinate for the difluoromethylation of quinoxaline.
The authors proposed a mechanistic pathway quite analogous
to the Minisci reaction; however, quinoxaline turned out to be
a weakly reactive substrate toward sodium difluoromethane
sulfinate furnishing a moderate yield of the product even with
Scheme 51 C–H alkylation reaction of quinoxaline.
6–7 equivalents of catalyst loading.
Convenient access to 2 or 3-substituted quinoxaline by
C-nucleophiles via substitution reactions was reported by
Ndlovu and co-workers.100
The group has investigated the effect of the substituent at
the 2-position of mono-substituted quinoxalines in the syn-
thesis of di-substituted quinoxaline derivatives via Negishi
reaction, Scheme 55.100

Arylation using haloarenes


The direct arylation of electron-withdrawing heteroarenes,
although appealing, is challenging as it requires harsh reac-
Scheme 52 Metal-free approach for Brønsted acid promoted C–H- tion conditions. Nevertheless, several efficient strategies have
alkylation of quinoxaline. been reported for the direct arylation of electron-deficient het-
eroarenes involving halogenated aryls 135, aryl boronic acids
136, and aryls 137 to access C-arylated heterocycles, as illus-
trated in Scheme 56.106–118
The direct arylation strategies for synthesizing arylated qui-
noxaline by reacting quinoxaline with iodo arenes under
different reaction conditions are illustrated in Scheme 57.106–108
A rapid microwave-assisted direct arylation of quinoxaline 1
with ortho-iodoarene 135a in the presence of potassium t-but-
oxide was reported by Yanagisawa and co-workers,
Scheme 57.106 The transition metal-free protocol furnishing
products 133 and 140 in a very short time and with good yields
are some of the advantages of this methodology; however, the
Scheme 53 The mechanism for the xanthate-based heteroaromatic poor product regioselectivity is certainly the major setback in
alkylation. its wide-scale applicability.

Scheme 51.103 The silver catalyzed Minisci reaction using


selectfluor as a mild oxidant and carboxylic acid as the alkylat-
ing source furnished the 2-alkylated quinoxaline 4 in moderate
to good yield within 24 h. The role of TFA in the reaction is to
enhance the electrophilicity of heterocycles by protonation. Scheme 54 Difluoro methylation of quinoxaline.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

and Li/Mg/Zn amides prepared easily from the secondary


amines.107
A similar kind of metalation of quinoxaline followed by a
cross-coupling strategy to access arylated quinoxaline was
reported by Wunderlich and co-workers. A directed ortho meta-
lation employing tmp2Zn·2MgCl2·2LiCl resulted in the corres-
ponding organometallic intermediate of quinoxaline which
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

was then subjected to a Pd-catalyzed cross-coupling reaction to


furnish the arylated quinoxaline 142. Other amide bases like
Scheme 55 The Negishi cross-coupling reaction. tmpMgCl·LiCl (tmp), 2,2,6,6-tetramethylpiperidyl, tmp2Mg·
2LiCl were also found to be equally competent. Highly reactive
tmp2Mg·2LiCl was found to be an effective agent for metala-
tion of less activated aromatic substrates, Scheme 57.108
Hyodo and co-workers developed a highly regioselective
nickel catalyzed oxidative nucleophilic arylation strategy for
the arylation of electron-deficient heteroarenes with good
yield. The regioselectivity, as well as substrate scope, comple-
mented the well-known direct arylations strategy as well as the
CMD pathway. The methodology avoided the use of any exter-
nal oxidants as the organozinc reagent functions both as an
Scheme 56 Schematic routes for the direct arylation of quinoxaline. aryl donor and an oxidant. Yields obtained from this reaction
at 80 °C were 66% with both mono and diaryl products,
whereas at 100 °C only diaryl product was obtained in 74%,
Scheme 58.109
Sakurai developed a method for decarbonylative direct ary-
lation using acyl fluorides 144 to furnish the arylated product
146 in 65% yield. An iridium(I) complex together with a
BrettPhos ligand served as an effective catalytic system for this
reaction. Therefore, the methodology only worked with acyl
fluorides 144, indicating the key role of fluorine in the acti-
vation of the C–H bond. However, the applicability of the pro-
tocol to acyl fluorides and not to other halo-fluorides is the
limitation of this methodology, Scheme 59.110
An electrochemical arylation protocol through reductive
activation of electro-deficient arenes and reacting with aryl dia-
zonium tetrafluoroborate 147 acting as an aryl source was
developed by Wang and co-workers. Cyclic voltammetry
studies revealed that both quinoxaline and aryl diazonium salt
had relatively low reduction potential, thereby proposing that
their activation could be via their reduction in the reaction,
Scheme 60.111
The mechanism for the electrochemical pathway is high-
lighted in Scheme 60. The reaction was initiated by the electro-

Scheme 57 Direct arylation of quinoxaline involving aryl halide as the


coupling component.

Later, a chemoselective direct arylation of quinoxaline was


reported by Rohbogner and co-workers, Scheme 57.107 The
one-pot chemoselective metalation (zincation or magnesiation)
of quinoxaline and subsequent cross-coupling with aryl iodide
in the presence of Pd-catalysts furnished quinoxaline 141 in
good yield and excellent selectivity, Scheme 57.107 The key for
the observed chemoselectivity was the directed magnesiation
or zincation of quinoxalines using a new kind of mixed Li/Mg Scheme 58 Sequential direct arylations of quinoxaline scaffold.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 59 Iridium-catalyzed decarbonylative coupling of acyl


fluorides with quinoxaline.

Scheme 61 Arylation using a boronic acid reagent.

coupling partner together with the functional group compatibil-


ity are the additional features of this methodology.
The direct arylation of quinoxaline reported by Deb and co-
workers in 2013 with aryl boronic acid employing inexpensive
Fe(NO3)2 as a catalyst in presence of K2S2O8 as an oxidant furn-
ished the desired arylated product 148 in very good yield
Scheme 61.112 The one-pot operationally simple strategy
avoided the pre-functionalization of the heterocycles and toler-
ated a wide variety of functional groups.
A room temperature open flask arylation of quinoxaline 1
was reported by Singh and coworkers involving the cross-coup-
ling reaction of electron-deficient quinoxaline with aryl
boronic acid in the presence of iron(II) acetylacetonate, oxidant
K2S2O8, and phase-transfer catalyst (TBAB). The mono-arylated
Scheme 60 Electrochemical method for the arylation of quinoxaline. quinoxaline was obtained in good yield within 12 h at room
temperature, Scheme 61.113
The enhanced stability of the potassium organotrifluorobo-
chemical reduction of the quinoxaline to generate the corres- rates over their corresponding boronic acids is because of the
ponding radical cation species. Simultaneously, 4-methyl lack of empty p-orbitals, thereby making them robust, easy to
phenyl radical 147a was also generated by the cathodic handle, and stable under harsh reaction conditions. Their
reduction of 4-methylbenzenediazonium tetrafluoroborate 147. operational ease and enhanced stability have prompted a few
Subsequent radical–radical coupling generated the 2-( p-tolyl)- of the research groups to employ potassium organotrifluorobo-
1,2-dihydro quinoxaline 147b which on oxidation at the anode rates as the coupling partners for the direct arylation of qui-
generated the aryl radical. A second single-electron-transfer noxalines as highlighted in Scheme 61.
(SET) oxidation and deprotonation furnished the desired Molander et al. reported the direct alkylation of various het-
product 148, Scheme 60.111 eroaryls using a stoichiometric amount of potassium alkyl-
and alkoxymethyl trifluoroborates as nucleophilic radical pre-
Arylation using boronic reagent cursors, Scheme 62.114
Direct arylation of electron-deficient quinoxaline with aryl Matsui and group employed mesityl acridinium perchlorate
boronic acid allows rapid access to aryl functionalized quinoxa- as a photocatalyst for the synthesis of 2-heteroaryl-substituted
lines. Owing to the stability, nontoxicity, and compatibility with 4-chromanones 154. The reaction involved a photoredox cata-
various reaction conditions, boronic acids have been considered lyzed coupling between 2-trifluoroborato-4-chromanones and
as the most versatile coupling component. The age-old strategy various heteroarenes through a mechanistic pathway like the
of using aryl boronic acid as the coupling partner has resulted Minisci reaction. The reaction offered a feasible route for
in the development of several novel protocols to gain access to the radical-induced C–H functionalization of quinoxalines
the arylated quinoxalines in good yields as highlighted in leading to the arylation in good chemo- and regioselectivity,
Scheme 61.112 The substrate scope concerning the boronic acid Scheme 62.115

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 64 The mechanism for palladium-catalyzed C-2 selective


Scheme 62 Direct arylation of quinoxalines using arylation.
organotrifluoroborates.

proposed that the formation of the palladium(II) complex with


Arylation via C–H activation strategy the quinoxaline will lead to the palladium migration to the C2
Direct functionalization of heteroarenes through C–H bond center of the quinoxaline 158a. Subsequent palladation will
activation is a highly attractive strategy. Moreover, C–H acti- generate the intermediate 158b. Simultaneously, electrophilic
vation has been a promising approach for functionalization of palladation of the dichlorobenzene through the C–H activation
the quinoxaline core, expanding the chemical space and pro- will generate the intermediate 158c which will then undergo
ducing innovative compounds. The reported C–H activation reductive elimination to afford the desired product 158. The
strategies to functionalize the C2 and C3 positions of quinoxa- subsequent oxidation of the Pd(0) to Pd(II) by the Ag(I)/PivOH/
lines are highlighted in Scheme 63.116–118 O2 completes the catalytic cycle, Scheme 64.116
Ren and co-workers have developed a Pd-catalyzed strategy Chen demonstrated a novel method to access heteroarenes
for the regioselective C2-arylation of unmodified quinoxaline 1 by the direct coupling of quinoxaline with pyrrole and its
with 1,2-dichlorobenzene 155 using Ag2CO3 and O2 oxidants, derivatives taking advantage of the cooperative actions of
albeit in low yields of 24%, Scheme 63.116 The authors have iridium and NaOTf, Scheme 63.117 The high regioselectivity in
suggested a mechanistic pathway for the direct arylation of terms of achieving the mono-arylated product along with high
unmodified quinoxaline with an arene, Scheme 64.116 It was yield and atom economy were claimed as the beneficial fea-
tures of this methodology. The proposed mechanism high-
lighted in Scheme 65 117 discloses the course of the reaction
along with the origin of high selectivity. The proposed mecha-
nism suggested the formation of a complex between the het-
eroarene and iridium catalyst via the β-metalation pathway.
Simultaneously, the coordination of NaOTf to the quinoxaline
afforded the corresponding triflate salt of the quinoxaline 1a.
The insertion of the carbon-iridium bond into the imino bond

Scheme 63 Direct C–H activation methods for the synthesis of Scheme 65 The general mechanism for iridium catalyzed cross-coup-
quinoxaline. ling reaction for the arylation of heteroarenes.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

and subsequent β-hydride elimination generated the desired The mechanistic pathway is detailed in Scheme 66,118
product 159. Finally, the interaction of HX with [HIrIIIX2L3] where the authors proposed a homolytic cleavage of the C–H
regenerated the catalyst with the release of the H2 gas. bond initiated by the excited TBADT and subsequent oxidation
In 2017, Quattrini and group developed facile sunlight- of the photocatalyst by K2S2O8, liberating an equivalent of acid
induced derivatization of heteroaromatics via photocatalyzed along with a strong oxidant (SO4•−). The generated cyclohexyl
C–H functionalization of the alkylating agent using tetra- radical was trapped by the protonated quinoxaline 1a leading
butylammonium decatungstate (TBADT) photocatalyst. to the formation of the desired derivative following the oxi-
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Cyclohexane was used as a hydrogen donor as it has the stron- dation by the strong oxidant (SO4•–), Scheme 66.
gest C–H bond (95 kcal mol−1). The drawback of the proposed
method is that it gives a mixture of mono and disubstituted Reactions using N-oxide
products instead of giving monosubstituted products alone, Quinoxaline N-oxide belongs to a notable class of intermedi-
Scheme 63.118 ates which provide regioselective access to several synthetic
intermediates that otherwise cannot be accessed by the direct
functionalization of quinoxaline. Several regioselective trans-
formations to selectively functionalize the C2-positions of qui-
noxaline have been achieved and are highlighted in
Scheme 67.119–123
Leclerc and co-workers reported a cross-coupling strategy
that involved the use of quinoxaline N-oxide 90 as an in-
expensive, bench stable reagent for the palladium-catalyzed
direct arylation with halo arenes involving Pd(OAc)2·tBu3PHBF4
in very good yields, Scheme 67.119
Transition metal-catalyzed Cross Dehydrogenative Coupling
(CDC) reactions of quinoxaline N-oxide and allied heterocycles
are known to construct C–C bonds. However, the construction
of a C–N bond via a CDC strategy was considered challenging
until the report by Li and co-workers in 2013. Li achieved C–H,
N–H dehydrogenative coupling of quinoxaline N-oxide with
lactams/cyclic amines in the presence of the Cu(OAc)2 catalyst
Scheme 66 Mechanistic pathway for the photocatalytic cross-dehy- to afford the dehydrogenative amidation of quinoxaline
drogenative coupling of quinoxalines. N-oxide in excellent yield, Scheme 67.120

Scheme 67 C–C, C–N, C–P bond formation strategies of quinoxaline N-oxide.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Li and co-workers reported a copper acetate mediated direct


amination of quinoxaline N-oxide in the presence of Ag2CO3
and tBuOH via a C–H activation strategy employing O-benzoyl
hydroxylamine 166 as an electrophilic amination reagent,
Scheme 67.122 The electrophilic amidation initiates with the for-
mation of catalytically active species by reducing Cu(II) to Cu(I)
followed by complexation with the solvent. After that, metalliza-
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

tion leads to the generation of aryl copper complex 90e which is


then added to O-benzoyl hydroxylamine 166 to form an inter-
mediate 90f. Subsequently, this intermediate then undergoes
reductive elimination to release Cu(I)OBz and the final product
167. Finally, ligand exchanges the solvent and regenerates the
catalytic species to repeat the catalytic cycle, Scheme 70.120
A phosphonate group has been successfully introduced into
the C2 positions of quinoxaline in a very good yield of 85% by
the process described by Chen and co-workers employing
Scheme 68 Mechanistic pathway for the amidation of quinoxaline in situ activation of P(OR)3 with bromotrichloromethane under
N-oxide. mild conditions in the absence of solvent and metal catalysts,
Scheme 67.123 Excellent compatibility with a wide range of
functional groups, low operational cost, and scalability of the
The mechanistic pathway is highlighted in Scheme 68.120 process along with the easy availability of reagents made that
Metalation of heterocyclic N-oxide with Cu(OAc)2 generates an methodology a prominent complement to the well-known
intermediate 90a that reacts with amide followed by oxidation Hirao coupling (synthesis of aryl phosphonates from haloar-
using Ag2CO3 to form a metal complex 90c. The metal complex enes ArX (X = Br, I) by palladium catalyst).
then undergoes reductive elimination to release CuIOAc and trans-Styrylquinoxalines can be considered as isosteres of
forms the final product 90d. Subsequently, CuIOAc gets oxi- stilbenes, which are widely present in natural compounds exhi-
dized by Ag2CO3 to generate the active catalyst which then biting a wide range of biological activities. Appropriate chemical
initiates the catalytic cycle, Scheme 68.120 modification strategies for this core could be useful for medic-
Realizing the fact that metal-free reactions have a great inal chemistry applications. Franco achieved the efficient alke-
impact on environmental remediation as well as cost nylation of quinoxaline N-oxide via Pd-catalyzed C–H activation,
reduction, Bering and co-workers reported oxidant- and metal- using the assistance of a mono-N-protected amino acid.
free regioselective cross-coupling of quinoxaline N-oxides with Quinoxaline N-oxide proved to be an ideal substrate for the
aryl boronic acids to access 2-arylated quinoxaline 133 in a low Fujiwara-Moritani coupling and offered enhanced reactivity at
yield of 38%, Scheme 67.121 The authors suggested a plausible the center ortho to the N-oxide functionality, Scheme 71.124
mechanistic pathway where the coordination of the aryl The mechanistic pathway highlights that the palladium(II)
boronic acid with the quinoxaline nitrogen leads to the inter- acetate forms a complex with the N-protected amino acid at
mediate 136a. Subsequent aryl migration from the intermedi-
ate 136a via a nucleophilic attack on C2 position of quin-
oxaline results in the formation of 2-arylated quinoxaline,
Scheme 69.121

Scheme 69 Mechanistic pathway for the C-2 arylation of quinoxaline Scheme 70 Mechanistic pathway for the electrophilic amination reac-
N-oxide with boronic acid. tion of quinoxaline N-oxide.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers


Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Scheme 72 Deoxygenative C–H amination reaction and the mechanis-


tic pathway highlighting C–H amination via [3 + 2] cycloaddition.

line or its 6,7-difluoro derivative with С-nucleophiles in pres-


ence of conc. HCl under nitrogen atmosphere. Along with the
substituted product 174, the formation of bisquinoxaline 175
was also observed in a low yield of 11%, Scheme 73.126
An alternative route to access N,N’-diquaternised-3,3′-biqui-
noxalinium ions was reported by Leblanc and co-workers via
the oxidative radical coupling of dithionite reduced quinoxa-
line quaternary salts. Although the methodology furnished the
product in moderate yield as well as poor regioselectivity, the
scalability of the reaction and the involvement of inexpensive
Scheme 71 Palladium-catalyzed C–H alkenylation of quinoxaline
N-oxide and proposed mechanism for quinoxaline N-oxide alkenylation reagents with no purification requirements were claimed to be
by the Fujiwara-Moritani reaction. the advantages of this reported process, Scheme 74.127
The route to the electrophilic nitration of quinoxaline was
disclosed by Juarez-Ornelas and co-workers employing iodosyl-
the beginning utilizing both the –NH and carboxylate groups benzene, which behaved as a nontoxic and iodine(III)-based
to coordinate with the Pd center in a bidentate fashion. organocatalyst in conjunction with Al(NO3)3 acting as the
Subsequently, this Pd-complex coordinates with the oxygen at nitrate source, Scheme 75.128 The mechanistic pathway
the N-oxide, and reversible ligand exchange occurs by a con- depicted in Scheme 75 proposes the formation of Al-complex
certed metalation/deprotonation (CMD) pathway. The complex obtained by the coordination of iodosylbenzene to aluminum
formation is relatively facile because of the hydrogen-bond nitrate followed by [3,3] cycloaddition to yield an intermediate
between the acetic acid ligand and the anionic oxygen. Finally, having iodine(III)-nitrate bonds.
coordination of Pd to the styrene, followed by carbopalladation Thereafter, decomposition of the intermediate to release
leads to the quinoxaline, Scheme 71.124 the NO2+ was accompanied by the regeneration of the NO2+
Accessing 2-substituted aminoquinoxaline 173 via a conven- which readily attacked the naphthol 179 or quinoxaline 1 and
tional route will be a two-step process involving halogenation
followed by amination. However, a one-pot metal-free reductive
amination of quinoxaline N-oxides with acyl azides 172 to
access 2-substituted amino quinoxalines 173 was reported by
Kim and co-workers. The key step is the Curtius rearrangement
which leads to the formation of isocyanate intermediate 172a
from acyl azides thereafter undergoing a [3 + 2] dipolar cyclo-
addition reaction with quinoxaline N-oxides to afford the
2-substituted amino quinoxalines 173, Scheme 72.125
Apart from the direct C-2 arylation/alkylation of quinoxa-
line, nucleophilic substitution with C-centered nucleophiles is
also an interesting strategy to functionalize the quinoxaline at
the C-2 position. Substitution of quinoxalines with C-centered Scheme 73 Direct nucleophilic substitution and dimerization of
nucleophiles was reported by Azev simply by treating quinoxa- quinoxaline.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

offers the advantage of an easy setup and mild reaction con-


ditions contrary to the direct hydrogenation with hydrogen gas.
In 1981, quinoxaline was reduced to corresponding 1,2,3,4-
tetrahydro quinoxaline using bis(trifluoroacetoxy) borane-tetra-
hydrofuran in TFA/THF. The important feature of this catalyst
is its stability in TFA to allow its employment as a reducing
agent (entry 1, Table 8).129
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Ranu and group used catalytic amounts of N,N-dimethyl


aniline, and zinc borohydride in DME under sonication for the
reduction of quinoxaline (entry 2, Table 8).130
McKinney used borane in THF solution for the reduction of
mono and di-substituted alkyl and aryl quinoxaline. It was
observed that the di-substituted compounds exclusively
yielded cis-isomer (entry 3, Table 8).131 Reduction of alkyl and
aryl quinoxalines was also achieved using sodium borohydride
Scheme 74 Synthesis of N,N’-diquaternised biquinoxen dications and in acetic acid; however, this resulted in diminished product
the mechanism of oxidative radical homocoupling for the synthesis of
yields along with side products. Borane-THF complex was
N,N‘-diquaternised biquinoxen dications.
found to be a superior reagent to render products in high
yields in a short duration.
Tan and co-workers developed a new catalytic system by
combining [Cp*IrCl2]2 with N-(2-aminoethyl)-4-(trifluoro-
methyl)benzenesulfonamide 185 in situ for regio- and chemo-
selective transfer hydrogenation of a variety of quinoxalines in
water employing HCOONa as the hydrogen source (entry 4,
Table 8). The catalyst was found to be air-stable and reduction
was performed without the protection of the nitrogen in an
acidic medium. The use of the HOAc/NaOAc buffer solution
was found to be essential for maintaining a stable pH during
the reaction, Scheme 76.132
Wu and coworkers developed iodine promoted transfer
hydrogenation in the presence of [(Cp*RhCl2)2] using KI as a
promotor and formic acid ( pKa 3.6) as the reductant by
heating in an azeotropic mixture of HCO2H/NEt3 at 40 °C. This
methodology could help to reduce the metal loading without
affecting the reaction conversion (entry 5, Table 8).133
Scheme 75 Nitration of quinoxaline and the mechanism for PhIO cata- A catalytic system comprising (pentamethylcyclopentadienyl)
lyzed electrophilic nitration. rhodium dichloride dimer [Cp*RhCl2]2 in conjunction with 2,2′-
bipyridine (bpy) was reported by Zhang and co-workers for
transfer hydrogenation of quinoxalines under aqueous con-
transferred a proton to the OAl(NO3)2 anion to afford nitrated ditions employing formate as the hydrogen source (entry 6,
product 180 or 178, respectively. Table 8). The choice of the ligand was found to be critical and
acidic pH is critical for optimal aqueous phase reduction,
Reduction Scheme 77.134
Tetrahydroquinoxaline 184 derivatives are abundant in natural The mechanistic pathway highlighted in Scheme 77 indi-
products as well as in medicinal chemistry. Therefore, the cates the in situ formation of the formate complex with [Cp*Rh
direct reduction of quinoxalines represents one of the most (bpy)Cl]Cl 187, followed by the rhodium hydride generation
frequently adopted strategies for the preparation of tetrahydro- from the decarboxylation of formate complex and subsequent
quinoxalines in one step. Several methodologies for this one- transfer to the protonated substrate 15a to furnish the target
step transformation have been developed in the last two tetrahydroquinoxaline 186 in good yield.
decades employing several catalytic systems which are high- The solvent has an important role to play in influencing the
lighted in Table 8. reaction rate and selectivity of the catalytic hydrogenation reac-
A quick perusal of the several conditions listed in Table 8 tion. Sun used a new catalyst containing metallic ruthenium
confirms that most of the protocols relied on the transfer nanoparticles intercalated in hectorite (nano-Ru′@hectorite) to
hydrogenation strategy where the dihydrogen equivalent is catalyze the reduction by NaBH4 in an aqueous solution under
transferred from an alcoholic source to quinoxaline. Transfer self-generated pressure. The catalyst can be recycled by cen-
hydrogenation, therefore, is a useful industrial strategy that trifugation (entry 7, Table 8).135

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

Table 8 List of reaction conditions for the reduction of quinoxaline to access tetrahydroquinoxaline
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Entry Reaction conditions % yield

1 BH[OC(O)CF3]2·THF, TFA/THF 86129


2 Zn(BH4)2/DME, N,N-dimethyl aniline, sonication, 10 h 90130
3 BH3·THF, N2 85–90131
4 [Cp*IrCl2]2 (2.5 µmol), N-(2-aminoethyl)-4-(trifluoromethyl)benzenesulfonamide (6.0 µmol), HCOONa, HOAc/NaOAc buffer, >91132
80 °C
5 [(Cp*RhCl2)2] (0.05–0.5 µmol), HCOOH/NEt3 azeotrope, KI (50 mol%), 40 °C, 12–24 h 99133
6 [(Cp*RhCl2)2] (0.025 µmol)/bpy (0.06 µmol), HCOOH/NaCOOH, H2O, 80 °C, pH 4.4, 3–12 h >90134
7 nanoRu’@hectorite, NaBH4, water, 60 °C, 4 h 99135
8 Pd(OAc)2 (10 mol%), B2pin2 (3.0 equiv.), water, N2, rt 75–80136
Pd(OAc)2 (10 mol%), B2pin2 (3.0 equiv.), Cs2CO3 (0.5 equiv.) water, N2, rt 90136
9 CoOx@CN (10 mg), MeOH, 3MPa H2, 120 °C, 10 h 77137
10 IrO2 NPs (0.04 mmol), H2 balloon, rt, 24 h 91–96138
11 B2(OH)4 (10 equiv.), 80 °C, 10 min 85139
12 PdNPs (1 mol%), water, H2 balloon, rt, 24 h 80–85140
13 Co(BF4)2·6H2O (5 mol%), tris(2-(diphenylphosphino) 88141
phenyl)phosphine (5 mol%), THF, H2 (10 bar), 60 °C, 15 h
14 Co(BF4)2·6H2O (5 mol%), tris(2-(diphenylphosphino) 90142
phenyl)phosphine (5 mol%), HCOOH (10 equiv.), dry i-PrOH, 100 °C, 20 h
15 B(C6H5)3 (5 mol%), NH3·BH3, toluene, 80 °C, 8 h 80143
16 B(C6H5)3 (5 mol%), Ph2SiH2 (5 equiv.), toluene, 110 °C, 24 h, N2 atm 78144
17 B(C6H5)3 (5 mol%), Et2SiH2 (3.5 equiv.), CHCl3, 25 °C–65 °C, 12 h, 0.25 M HCl/Et2O, Na2CO3, MeOH 80145
18 Co-NPs (5 mol%), 10 bar H2, 60 °C, 24 h 95146
19 Ru catalyst (1 mol%), t-BuOK (10 mol%), propanol, 80 °C, 24 h 70147
20 Mn catalyst (0.5 mol%), t-BuOK (5 mol%), propanol, 80 °C, 12 h 84148
21 HMPA (20 mol%), HSiCl3 (6.0 equiv.), DCM, 25 °C, 24 h 91–92149
22 Rh catalyst (1 mol%), 36 atm H2, neat, 100 °C, 48 h 92150
23 Co catalyst (0.1–3 mol%), KOH (10 mol%), i-PrOH, 30 atm H2, 100 °C, 48 h 92–95151
24 Ir catalyst (0.2 mol%), H2 (1 bar), water, 120 °C, 17 h 95152
25 Metallic Na, ammonia, t-BuOH 6579
Metallic Li, ammonia, t-BuOH 10079

Scheme 77 Possible mechanistic pathway for the transfer hydrogen-


ation of quinoxaline in water under acidic conditions.
Scheme 76 Transfer hydrogenation of 2-methylquinoxaline.

With this idea, Xuan and co-workers carried out Pd-catalyzed


transfer hydrogenation of various N-heteroaromatic compounds
Water being the most abundant, environmentally benign, employing air-stable and inexpensive borylating agent B2pin2 in
and inexpensive hydrogen-containing source in the world water which acts as a hydrogen donor (entry 8, Table 8).136
could be an ideal source of hydrogen for the transfer hydro- Often the presence of the N-heterocycle causes catalyst poi-
genation reaction. From the viewpoint of sustainable develop- soning due to the coordination of nitrogen atoms with the
ment, water is considered the most interesting solvent because metal nanoparticles, thereby making the catalyst recyclability
it is not only a green solvent but also plays a promising role in difficult. Therefore, the need to design metal nanoparticles
enhancing the catalyst activity. encapsulated with a protective shield to minimize the catalyst

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

leaching was highly desirable. The encapsulation of metal stability achieved from the aromaticity, Adam and Cabrero-
nanoparticles within graphene layers opens an avenue for the Antonino reported a general and efficient method for the
design of highly active and reusable heterogeneous catalysts nonprecious metal homogeneous cobalt-catalyzed hydrogen-
for more challenging molecules. ation of N-heterocycles to access quinoxaline in 88% yield,
With that idea, Wei designed a nanomagnetic material featur- under mild reaction conditions using the tetradentate ligand
ing nitrogen-doped carbon-coated cobalt NPs for the efficient tris(2-(diphenylphosphino) phenyl)phosphine and H2 balloon
and selective catalytic hydrogenation of quinoxaline. The cata- (entry 13, Table 8).141 Even the authors have demonstrated that
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

lytic performance was attributed to the enhanced recyclability of formic acid can also be used as an inexpensive and convenient
the catalyst due to the protective layer between the quinoxaline source of hydrogen (entry 14, Table 8).142
and the active metal NP along with the large surface area. The A transition-metal-free method for the hydrogenation of
mesoporous pores of the catalyst allow easy accessibility at the N-heterocycles to access tetrahydroquinoxaline was developed
catalytically active sites thereby allowing the catalyst to have by Ding and co-workers employing B(C6F5)3 along with
better contact with the reactants (entry 9, Table 8).137 ammonia borane under mild reaction conditions. Contrary to
Naked nanocatalysts usually exhibit higher catalytic the frustrated Lewis pairs (FLP)-catalyzed hydrogenations
efficiency because of their large surface-to-volume ratio. But at employing H2, the described method avoided the use of high-
the same time, they suffer from aggregation and poor re- pressure H2, thereby provided an alternative approach for the
usability, which impedes their applicability. However, in com- hydrogenation of unsaturated nitrogen-containing compounds
parison with the stabilized transition metal nanocatalysts, the (entry 15, Table 8).143
naked/unsupported nanocatalysts offer a better opportunity to Liu and coworkers employed B(C6F5)3 as a reducing agent
comprehend the intrinsic catalytic activity more easily by the for the efficient reduction of electron-deficient N-heteroarenes
elimination of the support effect. with hydrosilanes (or hydroboranes) and amines in good
Thereafter, Ji and co-workers prepared iridium(IV) oxide yields (entry 16, Table 8).144
nanoparticles by reacting iridium trichloride hydrate with Further, B(C6F5)3 was utilized by the group of Gandhamsetty
sodium hydroxide by ball-milling at room temperature and to develop a metal-free procedure for the reductive hydrogen-
employed them for the hydrogenation of a series of nitrogen ation of substituted N-heteroaromatics by using hydrosilanes as
heterocycles in high yield (entry 10, Table 8). The reaction reducing agents and heating quinoxaline in 0.25 M HCl/Et2O
demonstrated high efficiency and extraordinary recyclability mixture in MeOH at 65 °C (entry 17, Table 8).145
under mild reaction conditions.138 In 2018, Buschelberger and co-workers employed chemi-
Boron atoms possess a strong affinity for O and N atoms cally stable Co(0) nanoparticles and developed a straight-
and have the propensity to activate H2O and N-heterocycles. forward method for the catalytic hydrogenations of hetero-
Encouraged by this information, in 2016 Xia and co-workers cycles in a short time to access tetrahydroquinoxaline in good
reported diboronic acid-mediated hydrogenation of yields. The easy catalyst recovery process together with recycl-
N-heterocycles in water to access the tetrahydroquinoxaline ability to perform multiple hydrogenations without any obser-
within 10 min in good yields, Scheme 78. Mechanistic studies vable catalytic loss were the advantages of the reported proto-
revealed the proposed reaction pathway followed an initial 1,2- col. The ease of preparing cobalt catalyst without the use of
addition, where diboronic acid synergistically activated the any complex ligands, wide catalytic applicability, recyclability,
substrates and water via a six-membered ring transition state and long-term stability were claimed to be the other advan-
188a followed by the transfer of hydrogen from H2O (entry 11, tages of this method (entry 18, Table 8).146
Table 8).139 Alshakova reported the preparation of new ruthenium(II)
Zhang and co-workers demonstrated a mild protocol complexes supported by a pyrazole-phosphine ligand 189 and
employing carbon-metal covalent bond stabilized palladium its application in the synthesis via transfer hydrogenation of
nanoparticles for the hydrogenation of quinoxaline in water. various heterocycles (entry 19, Table 8).147
The role of water is to accelerate substrate absorption and
synergistic activation of molecular hydrogen. The Pd catalyst
can be easily recovered and reused for 5 runs on the Pd nano-
particles surface (entry 12, Table 8).140
Although the homogeneous hydrogenation of arenes to
access tetrahydroarenes is difficult because of the superior

Dubey and coworkers developed a simple bipyridine ligated


MnI complex 190 as an efficient catalytic system for transfer
hydrogenation of quinoxalines to furnish tetrahydroquinoxa-
lines in good yield. Among the several substituted bipyridines
screened, the complex supported by dihydroxy-substituted
Scheme 78 Metal-free hydrogen atom transfer from water. bipyridine was superior in catalytic activity, and the OH group

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

in bipyridine ligand 190 plays an important role in the transfer tetrahydroquinoxaline from quinoxaline in good yield display-
hydrogenation step (entry 20, Table 8).148 ing both high efficiency (up to 24 000 TON and 12 000 h−1
TOF) and excellent chemoselectivities (entry 23, Table 8).151
An iridium-based catalytic system bearing an electron-
enriched dipyridylamine ligand 193 was developed by
Wang and co-workers for the reversible hydrogenation of
N-heterocycle under mild conditions in water in the absence of
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

any co-catalyst or stoichiometric additives (entry 24, Table 8).152

Fu developed HMPA-mediated transfer hydrogenation for the


reduction of N-heteroarenes (entry 21, Table 8).149
Rhodium is known to be one of the most widely employed
metals in both homogeneous and heterogeneous transition
metals known to catalytically hydrogenate unsaturated organic
compounds.
Kim discovered additive-free hydrogenation of N-heteroarenes
using a homogeneous precatalyst [(η5-C5Me5)Rh(N–C)H] 191
(entry 22, Table 8).150 Since one of the well-known and trivial methods for the controlled
The need for the development of earth-abundant catalysts reduction of aromatic rings is via Birch reduction, Uçar and co-
to replace the noble metals led to the development of a highly workers carried out the reduction of quinoxaline with metallic
general and efficient method for homogeneous hydrogenation sodium in liquid ammonia and with metallic Li and obtained
of N-heteroarenes by Duan and co-workers in 2016. The group reasonably good yields of the product (entry 25, Table 8).79
has employed cobalt/tetraphosphine complex 192 to access the Troian-Gautier reported efficient oxidation of various elec-
tron-poor quinoxaline-core-containing compounds using [bis
(trifluoroacetoxy)iodo]benzene and acetonitrile/water solvent
mixture. [Bis(trifluoroacetoxy)iodo]benzene (BTI) is an in-
expensive, hypervalent iodine system that is commercially avail-
able, easy to synthesize, and was used as an oxidant in the reac-
tion. Highly electron-poor heterocycles are normally resistant to
oxidation. However, the use of BTI made the oxidation process
facile, Scheme 79.153 The mechanistic pathway depicted in
Scheme 79 proposes a nucleophilic attack by the quinoxaline
nitrogen to the iodine atom followed by the addition of water,
leading to the intermediate 194b. The electron density on the
carbon atom is reduced compared with that of the pyridinic
core due to the presence of the second nitrogen atom, thereby
favoring the second nucleophilic attack by water, ultimately fur-
nishing the desired product 197 in good yield.153
Kabadwal et al. discovered the cost-efficient iron-catalyzed
system for the alkylation of 2-methyl quinoxalines 198 with
alcohols 199 with moderate to good yield (entry 1, Table 9).154
Scheme 79 BTI catalyzed synthesis of quinoxaline core containing Recently, similar work has been reported by Jana et al.
molecules and the possible mechanistic pathway. where they explored manganese complex 201 as an efficient

Table 9 Metal catalyzed alkylation of 2-methylquinoxaline

Entry Conditions % yield

1 Fe(OAc)2 (5 mol%), 1,10-phenanthroline (6 mol%), t-BuOK, 1,4-dioxane, 140 °C, 24 h up to 82%154


2 Mn catalyst 201 (2 mol%), t-BuOK, t-AmOH, 140 °C, 24 h up to 81%155

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

catalyst for the functionalization of 2-methylquinoxaline (entry Conclusions


2, Table 9).155
The prominence of the quinoxaline scaffold in the areas of
medicinal chemistry, electroluminescent materials, DSSC,
molecular probes, and sensors has resulted in the upsurge of
the synthetic strategies to access quinoxaline, mildly and sus-
tainably in high yields. The necessity to decorate the scaffold
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

with proper functional groups to design quinoxaline-based


materials for diverse applications has reached a celestial
height. This account has presented an overview of the several
synthetic strategies to gain access to the quinoxaline scaffold
which having promising applications in the field of medicinal
chemistry, dyes, and optoelectronic materials. Routes to access
quinoxalines as well as functionalizing them have evolved
Additionally, Zilla et al. reported a method for the functionali- from the traditional routes to more sustainable routes invol-
zation of quinoxaline where the authors have synthesized ving greener catalytic systems, generating less wastes and less
2-phenoxyquinoxaline 204a and 1-phenylquinoxalin-2(1H)-one energies. Despite this remarkable advancement, it is still
204b in the ratio of 7 : 3 from quinoxaline-2(1H)-one. The reac- highly desirable to explore newer, greener, and cleaner routes
tion proceeds via the formation of an aryne intermediate to access and decorate quinoxalines. We conclude this review
which was generated in situ from 2-(trimethylsilyl)phenyl tri- by hoping that it will benefit the readers in endowing a com-
fluoromethane sulfonate and cesium fluoride, Scheme 80.156 prehensive overview about the quinoxaline chemistry and
The COVID pandemic situation has shaken the entire world stimulate interests in developing new catalytic methodologies
and the world has witnessed several publications highlighting to gain access to the quinoxaline scaffold and employ those
the design and development of molecules for the treatment of scaffolds in designing functional molecules and materials for
COVID-19 disease. 2-Substituted pyrrolo[2,3-b] quinoxalines diverse applications.
206 was found to be a promising candidate against COVID-19.
Therefore, Chemboli et al. explored the synthesis of 2-substi-
tuted pyrrolo[2,3-b]quinoxalines 206 via the Cu-catalyzed Conflicts of interest
sequential coupling-cyclization-desulfinylation of 3-alkynyl-2-
chloroquinoxalines 205 with t-butyl sulfinamide under ultra- The authors declare no conflicts of interest.
sound irradiation, Scheme 81.157

Acknowledgements
G.Y. is thankful to the Sarthi, Govt. of Maharashtra, India, for
her fellowship. S.S. acknowledges DST-SERB, CSIR ((02(0253)/
16/EMR-II)), UGC-Start-Up grant (FRP (F.4-5(128)/2014(BSR))),
ICT Golden Jubilee research grant (2018) and CoE-PI-TEQIP-III
for the financial assistance.

References
Scheme 80 Reaction of different substituted (E)-3-styrylquinoxalin-2
(1H)-one derivatives with 2-(trimethylsilyl)phenyltrifluoro methane 1 (a) Y. B. Kim, Y. H. Kim, J. Y. Park and S. K. Kim,
sulfonate.
Synthesis and biological activity of new quinoxaline anti-
biotics of echinomycin analogues, Bioorg. Med. Chem.
Lett., 2004, 14, 541–544, DOI: 10.1016/j.bmcl.2003.09.086;
(b) A. Jaso, B. Zarranz, I. Aldana and A. Monge, Synthesis
of New Quinoxaline-2-carboxylate 1,4-Dioxide Derivatives
as Anti-Mycobacterium tuberculosis Agents, J. Med.
Chem., 2005, 48, 2019–2025, DOI: 10.1021/jm049952w;
(c) M. S. F. Franco, M. H. d. Paula, P. C. Glowacka,
F. Fumagalli, G. C. Clososki and F. d. S. Emery, Palladium-
catalyzed C–H alkenylation of quinoxaline N-oxide
enabled by a mono-N-protected amino acid, Tetrahedron
Scheme 81 Ultrasound-assisted synthesis of pyrrolo[2,3-b ]quinoxaline Lett., 2018, 59, 2562–2566, DOI: 10.1016/j.
derivatives. tetlet.2018.05.054; (d) L. A. Raphoko, K. Lekgau,

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

C. M. Lebepe, T. C. Leboho, T. M. Matsebatlela and the Art review, Eur. J. Med. Chem., 2015, 97, 664–672, DOI:
W. Nxumalo, Synthesis of novel quinoxaline-alkynyl 10.1016/j.ejmech.2014.06.058.
derivatives and their anti-Mycobacterium tuberculosis, 7 J. M. Neri, L. N. Cavalcanti, R. M. Araujo and
activity, Bioorg. Med. Chem. Lett., 2021, 35, 127784, DOI: F. G. Menezes, 2,3-Dichloroquinoxaline as a versatile
10.1016/j.bmcl.2021.127784. building block for heteroaromatic nucleophilic substi-
2 (a) D. W. Chang, H. J. Lee, J. H. Kim, S. Y. Park, tution: a review of the last decade, Arabian J. Chem., 2020,
S.-M. Park, L. Dai and J.-B. Baek, Novel Quinoxaline-Based 13, 721–739, DOI: 10.1016/j.arabjc.2017.07.012.
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Organic Sensitizers for DyeSensitized Solar Cells, Org. 8 A. C. Shekhar, A. R. Kumar, G. Sathaiah, K. Raju,
Lett., 2011, 13(5), 3880–3883, DOI: 10.1021/ol2012378; P. V. S. S. Srinivas, P. S. Rao and B. Narsaiah, Aqueous
(b) K. Pei, Y. Wu, A. Islam, S. Zhu, L. Han, Z. Geng and Hydrofluoric Acid Catalyzed Facile Synthesis of 2,3,6-
W. Zhu, Dye-Sensitized Solar Cells Based on Quinoxaline Substituted Quinoxalines, J. Heterocycl. Chem., 2014, 51(5),
Dyes: Effect of π–Linker on Absorption, Energy Levels, and 1504–1508, DOI: 10.1002/jhet.1753.
Photovoltaic Performances, J. Phys. Chem. C, 2014, 9 H. R. Darabi, S. Mohandessi, K. Aghapoor and
118(30), 16552–16561, DOI: 10.1021/jp412259t; F. Mohsenzadeh, A recyclable and highly effective sulfa-
(c) D. N. Kanekar, S. Chacko and R. M. Kamble, mic acid/MeOH catalytic system for the synthesis of qui-
Quinoxaline based amines as Blue-orange emitters: Effect noxalines at room temperature, Catal. Commun., 2007, 8,
of modulating donor system on Optoelectrochemical and 389–392, DOI: 10.1016/j.catcom.2006.06.033.
theoretical properties, Dyes Pigm., 2019, 167, 36–50, DOI: 10 S. V. More, M. N. V. Sastry, C. C. Wang and C. F. Yao,
10.1016/j.dyepig.2019.04.005; (d) S. K. Dey, M. A. Kobaisi Molecular iodine: a powerful catalyst for the easy and
and S. V. Bhosale, Functionalized Quinoxaline for efficient synthesis of quinoxalines, Tetrahedron Lett., 2005,
Chromogenic and Fluorogenic Anion Sensing, 46, 6345–6348, DOI: 10.1016/j.tetlet.2005.07.026.
ChemistryOpen, 2018, 7, 934–952, DOI: 10.1002/ 11 R. S. Bhosale, S. R. Sarda, S. S. Ardhapure, W. N. Jadhav,
open.201800163; (e) A. A. Kalinin, L. N. Islamova, S. R. Bhusare and R. P. Pawar, An efficient protocol for the
A. G. Shmelev, G. M. Fazleeva, O. D. Fominykh, synthesis of quinoxaline derivatives at room temperature
Y. B. Dudkina, T. A. Vakhonina, A. I. Levitskaya, using molecular iodine as the catalyst, Tetrahedron Lett.,
A. V. Sharipova, A. S. Mukhtarov, A. R. Khamatgalimov, 2005, 46, 7183–7186, DOI: 10.1016/j.tetlet.2005.08.080.
I. R. Nizameev, Y. H. Budnikova and M. Y. Balakina, D–π-A 12 M. M. Heravi, S. Taheri, K. Bakhtiari and H. A. Oskooie,
chromophores with a quinoxaline core in the π-bridge and On Water: A practical and efficient synthesis of quinoxa-
bulky aryl groups in the acceptor: Synthesis, properties, line derivatives catalyzed by CuSO4.5H2O, Catal. Commun.,
and femtosecond nonlinear optical activity of the chromo- 2007, 8, 211–214, DOI: 10.1016/j.catcom.2006.06.013.
phore/PMMA guest-host materials, Dyes Pigm., 2021, 184, 13 M. M. Heravi, K. Bakhtiari, H. A. Oskooie and S. Taheri,
108801, DOI: 10.1016/j.dyepig.2020.108801; (f ) H. Yue, MnCl2-Promoted Synthesis of Quinoxaline Derivatives at
X. Guo, Y. Du, Y. Zhang, H. Du, J. Zhao and J. Zhang, Room Temperature, Heteroat. Chem., 2008, 19(2), 218–220,
Synthesis and characterization of donor–acceptor type DOI: 10.1002/hc.20401.
quinoxaline-based polymers and the corresponding elec- 14 Z. T. Bhutia, G. Prasannakumar, A. Das, M. Biswas,
trochromic devices with satisfactory open circuit memory, A. Chatterjee and M. Banerjee, A Facile, Catalyst-Free
Synth. Met., 2021, 271, 116619, DOI: 10.1016/j.synth- Mechano-Synthesis of Quinoxalines and their In vitro
met.2020.116619; (g) H. Wang, A. Q. Dayo, J. Wang, Antibacterial Activity Study, ChemistrySelect, 2017, 2,
J.-y. Wang and W.-b. Liu, Trifunctional quinoxaline-based 1183–1187, DOI: 10.1002/slct.201601672.
maleimide and its polymer alloys with benzoxazine: 15 S. Sajjadifar, G. Mansouri, I. Amini and M. Yari, Silica
Synthesis, characterization, and properties, J. Appl. Polym. Supported 1-(2-(Sulfooxy)ethyl)pyridin-1-ium Chloride
Sci., 2021, 138, 49694, DOI: 10.1002/app.49694. (SiO2/[SEP](Cl) as an Efficient and Solid Acid Catalyst for
3 G. W. H. Cheeseman and E. S. G. Werstiuk, Quinoxaline the Synthesis of Quinoxaline Derivatives, J. Med. Chem.
Chemistry: Developments 1963-1975, Adv. Heterocycl. Sci., 2021, 4, 8–16.
Chem., 1978, 22, 367–431, DOI: 10.1016/S0065-2725(08) 16 C. Srinivas, C. N. S. S. P. Kumar, V. J. Rao and
60107-5. S. Palaniappan, Efficient, convenient and reusable polya-
4 V. A. Mamedov and N. A. Zhukova, Progress in niline-sulfate salt catalyst for the synthesis of quinoxaline
Quinoxaline Synthesis (Part 1), Prog. Heterocycl. Chem., derivatives, J. Mol. Catal. A: Chem., 2007, 265, 227–230,
2012, 24, 55–88, DOI: 10.1016/B978-0-08-096807-0.00002- DOI: 10.1016/j.molcata.2006.10.018.
6. 17 K. B. Harsha, S. Rangappa, H. D. Preetham,
5 V. A. Mamedov and N. A. Zhukova, Progress in T. R. Swaroop, M. Gilandoust, K. S. Rakesh and
Quinoxaline Synthesis (Part 2), Prog. Heterocycl. Chem., K. S. Rangappa, An Easy and Efficient Method for the
2013, 25, 1–45, DOI: 10.1016/B978-0-08-099406-2.00001-7. Synthesis of Quinoxalines Using Recyclable and
6 J. A. Pereira, A. M. Pessoa, M. N. D. S. Cordeiro, Heterogeneous Nanomagnetic-Supported Acid Catalyst
R. Fernandes, C. Prudencio, J. P. Noronha and M. Vieira, under Solvent-Free Condition, ChemistrySelect, 2018, 3,
Quinoxaline, its derivatives and applications: A State of 5228–5232, DOI: 10.1002/slct.201800053.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

18 K. S. Indalkar, C. K. Khatri and G. U. Chaturbhuj, 29 B. B. F. Mirjalili and A. Akbari, Nano-TiO2: An eco-friendly


Rapid, efficient, and eco-friendly procedure for the syn- alternative for the synthesis of quinoxalines, Chin. Chem.
thesis of quinoxalines under solvent-free conditions Lett., 2011, 22, 753–756, DOI: 10.1016/j.cclet.2010.12.016.
using sulfated polyborate as a recyclable catalyst, J. Chem. 30 M. J. Climent, A. Corma, J. C. Hernández, A. B. Hungría,
Sci., 2017, 129(2), 141–148, DOI: 10.1007/s12039-017- S. Iborra and S. Martínez-Silvestre, Biomass into chemi-
1235-0. cals: One-pot two- and three-step synthesis of quinoxa-
19 B. Tamami, A. Sardarian and E. Ataollahi, Synthesis and lines from biomass-derived glycols and 1,2-dinitroben-
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

application of polyvinylimidazole-based Brønsted acidic zene derivatives using supported gold nanoparticles as
ionic liquid grafted silica as an efficient heterogeneous catalysts, J. Catal., 2012, 292, 118–129, DOI: 10.1016/j.
catalyst in the preparation of quinoxaline derivatives, jcat.2012.05.002.
Turk. J. Chem., 2016, 40, 422–433, DOI: 10.3906/kim-1504- 31 T. B. Nguyen, L. Ermolenko and A. Al-Mourabit, Sodium
40. Sulfide: A Sustainable Solution for Unbalanced Redox
20 U. P. Tarpada, B. B. Thummar and D. K. Raval, A green Condensation Reaction between o-Nitroanilines and
protocol for the synthesis of quinoxaline derivatives cata- Alcohols Catalyzed by an Iron–Sulfur System, Synthesis,
lyzed by polymer supported sulphanilic acid, Arabian J. 2015, 47, 1741–1748, DOI: 10.1055/s-0034-1380134.
Chem., 2017, 10, S2902–S2907, DOI: 10.1016/j.arabjc. 32 B. Roy, S. Ghosh, P. Ghosh and B. Basu, Graphene oxide
2013.11.021. (GO) or reduced graphene oxide (rGO): Efficient catalysts
21 A. Rahmatpour, Polystyrene-Supported AlCl3 as a Highly for onepot metal-free synthesis of quinoxalines from
Active and Reusable Heterogeneous Lewis Acid Catalyst 2-nitroaniline, Tetrahedron Lett., 2015, 56, 6762–6767,
for the One-Pot Synthesis of Quinoxalines, Heteroat. DOI: 10.1016/j.tetlet.2015.10.065.
Chem., 2012, 23(5), 472–477, DOI: 10.1002/hc.21039. 33 K. B. Harsha and K. S. Rangappa, One-step Approach for
22 K. Aghapoor, F. Mohsenzadeh, A. Shakeri, H. R. Darabi, Synthesis of Functionalized Quinoxalines Mediated by
M. Ghassemzadeh and B. Neumüller, Catalytic application T3P® - DMSO or T3P® via Tandem Oxidation –
of recyclable silica-supported bismuth(III) chloride in the Condensation or Condensation Reaction, RSC Adv., 2016,
benzo[N,N]-heterocyclic condensation, J. Organomet. 6, 57154–57162, DOI: 10.1039/C6RA03078E.
Chem., 2013, 743, 170–178, DOI: 10.1016/j. 34 A. Bera, M. Sk, K. Singh and D. Banerjee, Nickel-Catalysed
jorganchem.2013.06.037. Dehydrogenative Coupling of Aromatic Diamines with
23 V. P. Jejurkar, C. K. Khatri, G. U. Chaturbhuj and Alcohols: Selective Synthesis of Substituted
S. Saha, Environmentally Benign, Highly Efficient and Benzimidazoles and Quinoxalines, Chem. Commun., 2019,
Expeditious Solvent-Free Synthesis of Trisubstituted 55, 5958–5961, DOI: 10.1039/C9CC02319D.
Methanes Catalyzed by Sulfated Polyborate, 35 P. Daw, A. Kumar, N. A. Espinosa-Jalapa, Y. Diskin-Posner,
ChemistrySelect, 2017, 2, 11693–11696, DOI: 10.1002/slct. Y. Ben-David and D. Milstein, Synthesis of Pyrazines and
201702610. Quinoxalines via Acceptorless Dehydrogenative Coupling
24 S. Paul and B. Basu, Synthesis of libraries of quinoxalines Routes Catalyzed by Manganese Pincer Complexes, ACS
through eco-friendly tandem oxidation–condensation or Catal., 2018, 8, 7734–7741, DOI: 10.1021/acscatal.8b02208.
condensation reactions, Tetrahedron Lett., 2011, 52, 6597– 36 W. Tang, Y. Liu, S. Peng and S. Liu, Ruthenium(II) η6
6602, DOI: 10.1016/j.tetlet.2011.09.141. -arene complexes containing a dinucleating ligand based
25 C. Delpivo, G. Micheletti and C. Boga, paper A Green on 1,8-naphthyridine, J. Organomet. Chem., 2015, 775, 94–
Synthesis of Quinoxalines and 2,3-Dihydropyrazines, 100, DOI: 10.1016/j.jorganchem.2014.10.028.
Synthesis, 2013, 45, 1546–1552, DOI: 10.1055/s-0033- 37 B. Dowlati, D. Nematollahi and M. R. Othman, An
1338441. Efficient Electrochemical Method for the Synthesis of
26 D. Bandyopadhyay, S. Mukherjee and B. K. Banik, A Quinoxaline-dione Derivatives from Oxidation of
Selective, Expeditious and Sustainable Entry en Route to Catechols in the Presence of N1, N2-dibenzylethane-1,2-
Benzopyrazines and bis-Benzopyrazines, Comb. Chem. diamine, J. Electrochem. Soc., 2013, 160(1), G32–G36, DOI:
High Throughput Screening, 2015, 18, 53–62, DOI: 10.2174/ 10.1149/2.061301jes.
1386207318666150131125053. 38 V. A. Samsonov, Furazan ring opening upon treatment of
27 K. J. Tamuli, S. Nath and M. Bordoloi, In Water Organic benzofurazan with ethanolаmine to yield quinoxalines,
Synthesis: Introducing Itaconic Acid as a Recyclable Russ. Chem. Bull., 2007, 56(12), 2510–2512.
Acidic Promoter for Efficient and Scalable Synthesis of 39 B. Saha, B. Mitra, D. Brahmin, B. Sinha and P. Ghosh, 2-
Quinoxaline Derivatives at Room Temperature, Iodo benzoic acid: an unconventional precursor for the
J. Heterocycl. Chem., 2021, 58, 983–1002, DOI: 10.1002/ one pot multi-component synthesis of Quinoxaline using
jhet.4231. organo Cu(II) catalyst, Tetrahedron Lett., 2018, 59(41),
28 S. F. Hojati, Z. Nematdoust and T. Zeinali, The preparation 3657–3663, DOI: 10.1016/j.tetlet.2018.08.051.
of quinoxaline and 2,3-dihydropyrazine derivatives using 40 A. Monopoli, P. Cotugno, M. Cortese, C. D. Calvano,
selectfluor as an efficient and reusable catalyst, Iran. F. Ciminale and A. Nacci, Selective N-Alkylation of
Chem. Commun., 2015, 3, 6–15. Arylamines with Alkyl Chloride in Ionic Liquids: Scope

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

and Applications, Eur. J. Org. Chem., 2012, 3105–3111, Heterogeneous Catalyst for Oxidative Dehydrogenation of
DOI: 10.1002/ejoc.201200202. N-Heterocycles, ChemCatChem, 2017, 9(13), 2463–2466,
41 Y. Pan, C. Chen, X. Xu, H. Zhao, J. Han, H. Li, L. Xu, DOI: 10.1002/cctc.201700370.
Q. Fan and J. Xiao, Metal-free Tandem Cyclization/ 53 J. Zhang, S. Chen, F. Chen, W. Xu, G. Deng and H. Gong,
Hydrosilylation to Construct Tetrahydroquinoxalines, Dehydrogenation of N-Heterocycles Using Graphene
Green Chem., 2018, 20, 403–411, DOI: 10.1039/ Oxide as a Versatile Metal-free Catalyst under Air, Adv.
C7GC03095A. Synth. Catal., 2017, 359, 2358–2363, DOI: 10.1002/
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

42 S. Murahashi, T. Oda, T. Sugahara and Y. Masui, adsc.201700178.


Tungstate-Catalyzed Oxidation of Tetrahydroquinolines 54 K. Mullick, S. Biswas, A. M. Angeles-Boza and S. L. Suib,
with Hydrogen Peroxide: A Novel Method for the Heterogeneous mesoporous manganese oxide catalyst for
Synthesis of Cyclic Hydroxamic Acids, J. Org. Chem., 1990, aerobic and additive-free oxidative aromatization of
55(6), 1744–1749. N-heterocycles, Chem. Commun., 2017, 53, 2256–2259,
43 D. M. Karki, H. C. Araujo and J. Magolan, DOI: 10.1039/c6cc09095h.
Dehydroaromatization with V2O5, Synlett, 2013, 24, 1675– 55 W. Zhou, D. Chen, F. Sun, J. Qian, M. He and
1678, DOI: 10.1055/s-0033-1339277. Q. Chen, Aerobic oxidative dehydrogenation of
44 S. Chakraborty, W. W. Brennessel and W. D. Jones, A N-heterocycles catalyzed by cobalt porphyrin, Tetrahedron
Molecular Iron Catalyst for the Acceptorless Lett., 2018, 59, 949–953, DOI: 10.1016/j.tetlet.2018.
Dehydrogenation and Hydrogenation of N-Heterocycles, 01.094.
J. Am. Chem. Soc., 2014, 136, 8564–8567, DOI: 10.1021/ 56 D. Chamorro-Arenas, U. Osorio-Nieto, L. Quintero,
ja504523b. L. Hernandez-García and F. Sartillo-Piscil, Selective,
45 X. Cui, Y. Li, S. Bachmann, M. Scalone, A. Surkus, Catalytic and Dual C(sp3)-H Oxidation of Piperazines and
K. Junge, C. Topf and M. Beller, Synthesis and Morpholines Under Transition Metal-Free Conditions,
Characterization of Iron-Nitrogen-Doped Graphene/Core- J. Org. Chem., 2018, 83, 15333–15346, DOI: 10.1021/acs.
Shell Catalysts: Efficient Oxidative Dehydrogenation of joc.8b02564.
N-Heterocycles, J. Am. Chem. Soc., 2015, 137, 10652– 57 Q. Wang, H. Chai and Z. Yu, Acceptorless
10658, DOI: 10.1021/jacs.5b05674. Dehydrogenation of N–Heterocycles and Secondary
46 A. V. Iosub and S. S. Stahl, Catalytic Aerobic Alcohols by Ru(II)-NNC Complexes Bearing a Pyrazoyl-
Dehydrogenation of Nitrogen Heterocycles Using indolylpyridine Ligand, Organometallics, 2018, 37, 584–
Heterogeneous Cobalt Oxide Supported on Nitrogen- 591, DOI: 10.1021/acs.organomet.7b00902.
Doped Carbon, Org. Lett., 2015, 17, 4404–4407, DOI: 58 Y. Li, Y. Han, Z. Wang, R. Xu, W. Zhang, W. Chen,
10.1021/acs.orglett.5b01790. L. Zheng, J. Zhang, J. Luo, K. Wu, Y. Zhu, C. Chen,
47 R. Xu, S. Chakraborty, H. Yuan and W. D. Jones, Q. Peng, Q. Liu, P. Hu, D. Wang and Y. Li, Ordered Porous
Acceptorless, Reversible Dehydrogenation and N-Doping Carbon Matrix with Atomically Dispersed Co
Hydrogenation of N-heterocycles with a Cobalt Pincer Sites as an Efficient Catalyst for Dehydrogenation/
Catalyst, ACS Catal., 2015, 5, 6350–6354, DOI: 10.1021/ Transfer Hydrogenation of N-Heterocycles, Angew. Chem.,
acscatal.5b02002. Int. Ed., 2018, 57(35), 11262–11266, DOI: 10.1002/anie.
48 Y. Wang, C. Li and J. Huang, External ligand-free aerobic 201805467.
oxidation of N, C-containing cyclic systems under Pd-cata- 59 S. Saadati, N. Ghorashi, A. Rostami and F. Kobarfard,
lyzed conditions, Asian J. Org. Chem., 2016, 6(1), 44–46, Laccase-based oxidative catalytic systems for the aerobic
DOI: 10.1002/ajoc.201600465. aromatization of tetrahydroquinazolines and related
49 W. Zhou, P. Taboonpong, A. H. Aboo, L. Zhang, J. Jiang Nheterocyclic compounds under mild conditions,
and J. Xiao, A Convenient Procedure for the Oxidative Eur. J. Org. Chem., 2018, (30), 4050–4057, DOI: 10.1002/
Dehydrogenation of N-Heterocycles Catalyzed by FeCl2/ ejoc.201800466.
DMSO, Synlett, 2016, 27, A–D, DOI: 10.1055/s-0035- 60 Y. Zhang, S. Pang, Z. Wei, H. Jiao, X. Dai, H. Wang and
1561613. F. Shi, Synthesis of a molecularly defined single-active site
50 M. K. Sahoo, G. Jaiswal, J. Rana and E. Balaraman, heterogeneous catalyst for selective oxidation of
Organo-Photoredox Catalyzed Oxidative Dehydrogenation N-heterocycles, Nat. Commun., 2018, 9(1465), 1–10, DOI:
of N-Heterocycles, Chem. – Eur. J., 2017, 23, 14167–14172, 10.1038/s41467-018-03834-4.
DOI: 10.1002/chem.201703642. 61 W. Zhou, Q. Tao, F. Sun, X. Cao, J. Qian, J. Xu, M. He,
51 C. Deraedt, R. Ye, W. T. Ralston, F. D. Toste and Q. Chen and J. Xiao, Additive-free aerobic oxidative dehy-
G. A. Somorjai, Dendrimer-Stabilized Metal Nanoparticles drogenation of N-heterocycles under catalysis by NiMn
as Efficient Catalysts for Reversible Dehydrogenation/ layered hydroxide compounds, J. Catal., 2018, 361, 1–11,
Hydrogenation of N-Heterocycles, J. Am. Chem. Soc., 2017, DOI: 10.1016/j.jcat.2018.01.030.
139, 18084–18092, DOI: 10.1021/jacs.7b10768. 62 N. O. Balayeva, N. Zheng, R. Dillert and
52 X. Sun, J. Zhu, Y. Xia and L. Wu, Metal-Carbon Covalent D. W. Bahnemann, Visible-Light Mediated Photocatalytic
Bonds Stabilized Palladium Nanoparticles as Expeditious Aerobic Dehydrogenation of N-Heterocycles by Surface

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Grafted TiO2 and 4-Amino-TEMPO, ACS Catal., 2019, 9, 75 A. Vidal-Albalat, S. Rodríguez and F. V. González,
10694–10704, DOI: 10.1021/acscatal.9b03322. Nitroepoxides as Versatile Precursors to 1,4-Diamino
63 X. Bi, T. Tang, X. Meng, M. Gou, X. Liu and P. Zhao, Heterocycles, Org. Lett., 2014, 16, 1752–1755, DOI:
Aerobic oxidative dehydrogenation of N-heterocycles over 10.1021/ol500444z.
OMS-2-based nanocomposite catalysts: preparation, 76 M. M. Ibrahim, D. Grau, F. Hampel and S. B. Tsogoeva,
characterization and kinetic study, Catal. Sci. Technol., α-Nitro Epoxides in Organic Synthesis: Development of a
2020, 10, 360–371, DOI: 10.1039/c9cy01968e. One-Pot Organocatalytic Strategy for the Synthesis of
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

64 T. Liu, K. Wu, L. Wang and Z. Yu, Potassium tert- Quinoxalines, Eur. J. Org. Chem., 2014, (7), 1401–1405,
Butoxide-Promoted Acceptorless Dehydrogenation of DOI: 10.1002/ejoc.201301591.
N-Heterocycles, Adv. Synth. Catal., 2019, 361, 1–8, DOI: 77 T. I. Filyakova, A. Ya. Zapevalov, M. I. Kodess,
10.1002/adsc.201900499. P. A. Slepukhin and V. I. Saloutin, Perfluoroepoxyoxolanes
65 S. Wang, H. Huang, C. Bruneau and C. Fischmeister, in the Synthesis of Fluorine-Containing Heterocycles,
Iridium Catalyzed Hydrogenation and Dehydrogenation of Russ. J. Org. Chem., 2009, 45(6), 884–889, DOI: 10.1134/
N–Heterocycles in Water under Mild Conditions, S1070428009060141.
ChemSusChem, 2019, 12(11), 2350–2354, DOI: 10.1002/ 78 L. V. Saloutina, A. Y. Zapevalov, V. I. Saloutin,
cssc.201900626. M. I. Kodess, V. E. Kirichenko, M. G. Pervova and
66 Y. Wu, Z. Chen, W. Cheong, C. Zhang, L. Zheng, W. Yan, O. N. Chupakhin, Reactions of epoxide derived from
R. Yu, C. Chen and Y. Lia, Nitrogen-Coordinated Cobalt internal perfluoroolefins, with o-phenylene diamine and
Nanocrystals for Oxidative Dehydrogenation and 2-aminophenol, Russ. J. Org. Chem., 2006, 42(4), 558–566,
Hydrogenation of N-Heterocycles, Chem. Sci., 2019, 10, DOI: 10.1134/S1070428006040130.
5345–5352, DOI: 10.1039/C9SC00475K. 79 S. Uçar, S. Essiz and A. Dastan, Bromination of
67 K. Yu, H. Zhang, C. Su and Y. Zhu, Visible-Light-Promoted Quinoxaline and Derivatives: Effective Synthesis of Some
Efficient Aerobic Dehydrogenation of NHeterocycles by New Brominated Quinoxalines, Tetrahedron, 2017, 73,
Tiny Organic Semiconductor under Ambient Conditions, 1618–1632, DOI: 10.1016/j.tet.2017.02.014.
Eur. J. Org. Chem., 2020, (13), 1956–1960, DOI: 10.1002/ 80 Z. Dong, W. Zhu, Z. Zhang and M. Li, An efficient and
ejoc.202000170. mild ortho-zincation of aromatics and heterocycles by
68 S. Bera, A. Bera and D. Banerjee, Nickel-Catalyzed using TMP2Mg2LiCl in the presence of ZnCl2,
Dehydrogenation of N-Heterocycles Using Molecular J. Organomet. Chem., 2010, 695, 775–780, DOI: 10.1016/j.
Oxygen, Org. Lett., 2020, 22, 6458–6463, DOI: 10.1021/acs. jorganchem.2009.12.016.
orglett.0c02271. 81 V. L. Blair, D. C. Blakemore, D. Hay, E. Hevia and
69 F. Xie, G. Lu, R. Xie, Q. Chen, H. Jiang and M. Zhang, D. C. Pryde, Alkali-metal mediated zincation of
MOF-Derived Subnanometer Cobalt Catalyst for Selective N-heterocyclic substrates using the lithium zincate
C-H Oxidative Sulfonylation of Tetrahydroquinoxalines complex, (THF)Li(TMP)Zn(tBu)2 and applications in
with Sodium Sulfinates, ACS Catal., 2019, 9, 2718–2724, in situ cross coupling reactions, Tetrahedron Lett.,
DOI: 10.1021/acscatal.9b00037. 2011, 52, 4590–4594, DOI: 10.1016/j.tetlet.2011.
70 S. Murata, T. Sugimoto and S. Matsuura, A Novel Ring 06.090.
Formation Of 1,2-Dihydroquinoxalines, Heterocycles, 1987, 82 (a) G. Liu, Transition Metal-Catalyzed Fluorination of
26(4), 883–884. Multi Carbon-Carbon Bonds: New Strategies for
71 S. Antoniottia and E. Dunach, Direct and catalytic Fluorinated Heterocycles, Org. Biomol. Chem., 2012, 10,
synthesis of quinoxaline derivatives from epoxides and 6243–6248, DOI: 10.1039/C2OB25702E; (b) X.-G. Hu and
ene-1,2-diamines, Tetrahedron Lett., 2002, 43, 3971– L. Hunter, Stereoselectively fluorinated N-heterocycles: a
3973. brief survey, Beilstein J. Org. Chem., 2013, 9, 2696–2708,
72 E. C. Taylor, C. A. Maryanoff and J. S. Skotnicki, DOI: 10.3762/bjoc.9.306; (c) D. DesMarteau, Fluorinated
Heterocyclization with Cyano and Sulfonyl Epoxides. Heterocyclic Compounds: Synthesis, Chemistry, and
Preparation of Quinoxalines and Tetrahydroquinoxalines, Applications, J. Am. Chem. Soc., 2010, 132(9), 3229–3230,
J. Org. Chem., 1980, 45, 2512–2515. DOI: 10.1021/ja100496w.
73 G. Srikanth, K. V. S. Ramakrishna and G. V. M. Sharma, 83 R. D. Chambers, M. Parsons, G. Sandford, C. J. Skinner,
A Double Activation Method for the Conversion of M. J. Atherton and J. S. Moilliet, Elemental fluorine. Part
Vinyl Epoxides into vic-Amino Alcohols and Chiral 10.1 Selective fluorination of pyridine, quinoline and qui-
Benzoxazine/Quinoxaline Derivatives, Org. Lett., 2015, noxaline derivatives with fluorine–iodine mixtures,
17(18), 4576–4579, DOI: 10.1021/acs.orglett.5b02304. J. Chem. Soc., Perkin Trans. 1, 1999, 1, 803–810, DOI:
74 M. S. Tu, H.-W. Xu, W. Fan, B. Jiang and S. J. Tu, [4 + 2] 10.1039/A809838G.
Heterocyclization for Efficient Formation of Substituted 84 P. S. Fier and J. F. Hartwig, Selective C-H Fluorination of
Quinoxalines through Carbon–Oxygen Bonds Cleavage, Pyridines and Diazines Inspired by a Classic Amination
J. Heterocycl. Chem., 2015, 52(3), 719–725, DOI: 10.1002/ Reaction, Science, 2013, 342(6161), 956–960, DOI: 10.1126/
jhet.2128. science.1243759.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

85 N. Hirasawa, Y. Takahashi, E. Fukuda, O. Sugimoto and 94 J. Chen, M. Wan, J. Hua, Y. Sun, Z. Lv, W. Li and L. Liu,
K. Tanji, Indium-mediated dehalogenation of haloheter- TBHP/TFA Mediated Oxidative Cross-Dehydrogenative
oaromatics in water, Tetrahedron Lett., 2008, 49, 1492– Coupling of NHeterocycles with Aldehydes, Org. Biomol.
1494, DOI: 10.1016/j.tetlet.2007.12.116. Chem., 2015, 13, 11561–11566, DOI: 10.1039/
86 E. Fukuda, Y. Takahashi, N. Hirasawa, O. Sugimoto and C5OB01763G.
K. Tanji, Dehalogenation And Barbier-Type 95 Y. Siddaraju and K. R. Prabhu, Transition Metal-Free
Hydroxyalkylation Of Π-Deficient Haloheterocycles Using Minisci Reaction Promoted by NCS, and TBHP: Acylation
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

Indium, Heterocycles, 2009, 77(2), 1163–1170, DOI: of Heteroarenes, Tetrahedron, 2016, 72, 959–967, DOI:
10.3987/COM-08-S(F)95. 10.1016/j.tet.2015.12.065.
87 G. Chelucci and S. Figus, NaBH4-TMEDA and a palladium 96 L. Zhang, G. Zhang, Y. Li, S. Wang and A. Lei, Synergistic
catalyst as efficient regio- and chemoselective system for Effect of Self-assembling and Visible-light Induced the
the hydrodehalogenation of halogenated heterocylces, Oxidative C-H Acylation of N-heterocyclic Aromatic
J. Mol. Catal. A: Chem., 2014, 393, 191–209, DOI: 10.1016/j. Compounds with Aldehydes, Chem. Commun., 2018, 54,
molcata.2014.06.012. 5744–5747, DOI: 10.1039/c8cc02342e.
88 A. Gevorgyan, S. Mkrtchyan, T. Grigoryan and 97 R. Sharma, M. Abdullaha and S. B. Bharate, Oxidant-
V. O. Iaroshenko, Application of Silicon-Initiated Water Controlled C-sp2/sp3 – H Cross-dehydrogenative Coupling
Splitting for the Reduction of Organic Substrates, of N-Heterocycles with Benzylamines, J. Org. Chem., 2017,
ChemPlusChem, 2018, 83(5), 375–382, DOI: 10.1002/ 82, 9786–9793, DOI: 10.1021/acs.joc.7b00856.
cplu.201800131. 98 S. Sultan, M. A. Rizvi, J. Kumar and B. A. Shah, Acyl rad-
89 F. Fontana, F. Minisci, M. Claudia, N. Barbosa and icals from terminal alkynes: Photoredox-catalyzed acyla-
E. Vismara, Homolytic Acylation of Protonated Pyridines tion of heteroarenes, Chem. – Eur. J., 2018, 24, 10617–
and Pyrazines with a-Keto Acids: The Problem of 10620, DOI: 10.1002/chem.201801628.
Monoacylation, J. Org. Chem., 1991, 56(8), 2866–2869, 99 H. Ding, K. Xu and C. Zeng, Nickel-catalyzed electro-
DOI: 10.1021/jo00008a050. chemical Minisci acylation of aromatic N-heterocycles
90 N. R. Chaubey and K. N. Singh, Metal-Free with a-keto acids via ligand-to-metal electron transfer
Decarboxylative Acylation of Isoquinolines using α-Keto pathway, J. Catal., 2020, 381, 38–43, DOI: 10.1016/j.
Acids in Water, Tetrahedron Lett., 2017, 58, 2347–2350, jcat.2019.10.030.
DOI: 10.1016/j.tetlet.2017.04.098. 100 N. T. Ndlovu and W. Nxumalo, Nucleophilic Substitution
91 X. Wang and C. Zeng, Iron-catalyzed Minisci acylation of on 2-Monosubstituted Quinoxalines Giving 2,3-
N-heteroarenes with α-keto acids, Tetrahedron, 2019, 75, Disubstituted Quinoxalines: Investigating the Effect of the
1425–1430, DOI: 10.1016/j.tet.2019.01.060. 2-Substituent, Molecules, 2016, 21(10), 1304, DOI: 10.3390/
92 L. Guillemard, F. Colobert and J. Wencel-Delord, Visible- molecules21101304.
light-triggered, metal- and photocatalyst-free acylation of 101 Y. Zeng, B. Qian, Y. Li and H. Bao, A Metal-Free Approach
N-heterocycles, Adv. Synth. Catal., 2018, 360(21), 4184– for Brønsted Acid Promoted C–H Alkylation of
4190, DOI: 10.1002/adsc.201800692. Heteroarenes with Alkyl Peroxides, Synthesis, 2018, 50,
93 (a) L.-Y. Liu, J. X. Qiao, K.-S. Yeung, W. R. Ewing and J.,- 3250–3256, DOI: 10.1055/s-0037-1609965.
Q. Yu, meta C-H Arylation of Electron-Rich Arenes: 102 V. L. Revil-Baudard, J. Vors and S. Z. Zard, Xanthate-
Reversing the Conventional Site-Selectivity, J. Am. Chem. Mediated Incorporation of Quaternary Centers into
Soc., 2019, 141, 14870–14877, DOI: 10.1021/jacs.9b07887; Heteroarenes, Org. Lett., 2018, 20, 3531–3535, DOI:
(b) S. A. Girard, T. Knauber and C.-J. Li, The Cross- 10.1021/acs.orglett.8b01299.
Dehydrogenative Coupling of Csp3H Bonds: A Versatile 103 J. D. Galloway, D. N. Mai and R. D. Baxter, Silver-Catalyzed
Strategy for CC Bond Formations, Angew. Chem., Int. Ed., Minisci Reactions Using Selectfluor as a Mild Oxidant,
2013, 52, 2–29, DOI: 10.1002/anie.201304268; Org. Lett., 2017, 19, 5772–5775, DOI: 10.1021/acs.
(c) K. Matcha and A. P. Antonchick, Metal-Free Cross- orglett.7b02706.
Dehydrogenative Coupling of Heterocycles with 104 X. Xie, Y. Zhang, J. Hao and W. Wan, Ag-catalyzed Minisci
Aldehydes, Angew. Chem., Int. Ed., 2013, 125(7), 2136– C-H difluoromethylarylation of N-heteroarenes, Org.
2140, DOI: 10.1002/ange.201208851; (d) Z. Li and C.-J. Li, Biomol. Chem., 2020, 18, 400–404, DOI: 10.1039/
CuBr-Catalyzed Direct Indolation of C9OB02586C.
Tetrahydroisoquinolines via Cross-Dehydrogenative 105 M. A. Lytkina, E. V. Eliseenkov, V. P. Boyarskii and
Coupling between sp3 C-H and sp2 C-H Bonds, J. Am. A. A. Petrov, Sodium Difluoromethanesulfinate—A
Chem. Soc., 2005, 127(19), 6968–6969, DOI: 10.1021/ Difluoromethylating Agent toward Protonated
ja0516054; (e) F. A. H. Nasab, L. Z. Fekri, A. Monfared, Heterocyclic Bases, Russ. J. Org. Chem., 2017, 53(4), 539–
A. Hosseinian and E. Vessally, Recent advances in sulfur– 546, DOI: 10.1134/S1070428017040066.
nitrogen bond formation via cross-dehydrogenative coup- 106 S. Yanagisawa, K. Ueda, T. Taniguchi and K. Itami,
ling reactions, RSC Adv., 2018, 8, 18456–18469, DOI: Potassium t-Butoxide Alone Can Promote the Biaryl
10.1039/C8RA00356D. Coupling of Electron-Deficient Nitrogen Heterocycles and

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

Haloarenes, Org. Lett., 2008, 10(20), 4673–4676, DOI: Hydrogen Donors via Decatungstate Photocatalysis, Chem.
10.1021/ol8019764. Commun., 2017, 53, 2335–2338, DOI: 10.1039/
107 C. J. Rohbogner, S. H. Wunderlich, G. C. Clososki and C6CC09725A.
P. Knochel, New Mixed Li/Mg and Li/Mg/Zn Amides for 119 J. Leclerc and K. Fagnou, Palladium-Catalyzed Cross-
the Chemoselective Metallation of Arenes and Coupling Reactions of Diazine N-Oxides with Aryl
Heteroarenes, Eur. J. Org. Chem., 2009, 1781–1795, DOI: Chlorides, Bromides, and Iodides, Angew. Chem., Int. Ed.,
10.1002/ejoc.200801277. 2006, 45, 7781–7786, DOI: 10.1002/anie.200602773.
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

108 S. H. Wunderlich, C. J. Rohbogner, A. Unsinn and 120 G. Li, C. Jia and K. Sun, Copper-Catalyzed Intermolecular
P. Knochel, Scaleable Preparation of Functionalized Dehydrogenative Amidation/Amination of Quinoline N–
Organometallics via Directed Ortho Metalation Using Mg- Oxides with Lactams/Cyclamines, Org. Lett., 2013, 15(20),
and Zn-Amide Bases, Org. Process Res. Dev., 2010, 14, 339– 5198–5201, DOI: 10.1021/ol402324v.
345, DOI: 10.1021/op9002888. 121 L. Bering and A. P. Antonchick, Regioselective Metal-Free
109 I. Hyodo, M. Tobisu and N. Chatani, Catalytic Arylation of Cross-Coupling of Quinoline N–Oxides with Boronic
a CH Bond in Pyridine and Related Six-Membered Acids, Org. Lett., 2015, 17, 3134–3137, DOI: 10.1021/acs.
N-Heteroarenes Using Organozinc Reagents, Chem. – orglett.5b01456.
Asian J., 2012, 7(6), 1357–1365, DOI: 10.1002/ 122 G. Li, C. Jia, K. Sun, Y. Lv, F. Zhao, K. Zhou and H. Wu,
asia.201100971. Copper(II)-Catalyzed Electrophilic Amination of Quinoline
110 S. Sakurai, T. Yoshida and M. Tobisu, Iridium-Catalyzed N–Oxides with O–Benzoyl hydroxylamines, Org. Biomol.
Decarbonylative Coupling of Acyl Fluorides with Arenes Chem., 2015, 13, 3207–3210, DOI: 10.1039/C5OB00135H.
and Heteroarenes via C-H Activation, Chem. Lett., 2019, 123 M. Chen, X. You, L. Bai and Q. Luo, Metal-free phospho-
48(2), 94–97, DOI: 10.1246/cl.180857. nation of heteroarene N-oxides with trialkyl phosphite at
111 P. Wang, Z. Yang, Z. Wang, C. Xu, L. Huang, S. Wang, room temperature, Org. Biomol. Chem., 2017, 15, 3165–
H. Zhang and A. Lei, Electrochemical Arylation of 3169, DOI: 10.1039/c7ob00402h.
Electron-Deficient Arenes through Reductive Activation, 124 M. S. F. Franco, M. H. d. Paula, P. C. Glowacka,
Angew. Chem., Int. Ed., 2019, 131(44), 15894–15898, DOI: F. Fumagalli, G. C. Clososki and F. d. S. Emery,
10.1002/anie.201909600. Palladium-catalyzed C–H alkenylation of quinoxaline
112 A. Deb, S. Manna, A. Maji, U. Dutta and D. Maiti, Iron- N-oxide enabled by a mono-N-protected amino acid,
Catalyzed Direct C–H Arylation of Heterocycles and Tetrahedron Lett., 2018, 59, 2562–2566, DOI: 10.1016/j.
Quinones with Arylboronic Acids, Eur. J. Org. Chem., 2013, tetlet.2018.05.054.
(24), 5251–5256, DOI: 10.1002/ejoc.201300743. 125 D. Kim, P. Ghosh, N. Y. Kwon, S. H. Han, S. Han,
113 P. P. Singh, S. K. Aithagani, M. Yadav, V. P. Singh and N. K. Mishra, S. Kim and I. S. Kim, Deoxygenative
R. A. Vishwakarma, Iron catalyzed cross coupling of elec- Amination of Azine-N-oxides with Acyl Azides via [3 + 2]
tron-deficient heterocycles and quinone with organoboron Cycloaddition, J. Org. Chem., 2020, 85, 2476–2485, DOI:
species via innate C-H functionalization: Application in 10.1021/acs.joc.9b03173.
total synthesis of pyrazine alkaloid Botryllazine A, J. Org. 126 Y. А. Azev, О. S. Ermakova, V. А. Bakulev, I. S. Kovalev,
Chem., 2013, 78, 2639–2648, DOI: 10.1021/jo302797r. А. N. Tsmokalyuk, А. N. Kozitsina, М. G. Pervova and
114 G. A. Molander, V. Colombel and V. A. Braz, Direct V. I. Filyakova, Features of Quinoxaline Reactions with
Alkylation of Heteroaryls Using Potassium Alkyl- and C-Nucleophiles: Examples of Dimerization of Heterocycle
Alkoxymethyltrifluoroborates, Org. Lett., 2011, 13(7), in Course of Hydrogen Substitution, Russ. J. Gen. Chem.,
1852–1855, DOI: 10.1021/ol2003572. 2015, 85(7), 1635–1638, DOI: 10.1134/S1070363215070105.
115 J. K. Matsui and G. A. Molander, Organocatalyzed, 127 N. Leblanc, S. Sproules and A. K. Powell, An alternative
Photoredox Heteroarylation of 2– method to access diverse N,N′ -diquaternised-3,3′ -biqui-
Trifluoroboratochromanones via C–H Functionalization, noxalinium “biquinoxen” dications, New J. Chem., 2017,
Org. Lett., 2017, 19, 950–953, DOI: 10.1021/acs. 41, 2949–2954, DOI: 10.1039/c7nj00531h.
orglett.7b00196. 128 K. A. Juarez-Ornelas, J. O. C. Jimenez-Halla, T. Kato,
116 X. Ren, P. Wen, X. Shi, Y. Wang, J. Li, S. Yang, H. Yan and C. R. Solorio-Alvarado and K. Maruoka, Iodine(III)-
G. Huang, Palladium-Catalyzed C–2 Selective Arylation of Catalyzed Electrophilic Nitration of Phenols via
Quinolines, Org. Lett., 2013, 15(20), 5194–5197, DOI: NonBrønsted Acidic NO2+ Generation, Org. Lett., 2019, 21,
10.1021/ol402262c. 1315–1319, DOI: 10.1021/acs.orglett.8b04141.
117 C. Chen, X. Chen, H. Zhao, H. Jiang and M. Zhang, Direct 129 B. E. Maryanoff, D. F. McComsey and S. O. Nortey,
Access to Nitrogen Bi-heteroarenes via Iridium-Catalyzed Properties of Bis(trifluoroacetoxy)borane as a Reducing
Hydrogen-Evolution Cross-Coupling Reaction, Org. Lett., Agent of Organic Compounds, J. Org. Chem., 1981, 46,
2017, 19, 3390–3393, DOI: 10.1021/acs.orglett.7b01349. 355–360.
118 M. C. Quattrini, S. Fujii, K. Yamada, T. Fukuyama, 130 B. C. Ranu, U. Jana and A. Sarkar, Regioselective
D. Ravelli, M. Fagnoni and I. Ryu, Versatile Cross- Reduction of Quinolines and Related Systems to 1,2,3,4-
Dehydrogenative Coupling of Heteroaromatics and Tetrahydro Derivatives with Zinc Borohydride, Synth.

This journal is © the Partner Organisations 2021 Org. Chem. Front.


View Article Online

Review Organic Chemistry Frontiers

Commun., 1998, 28(3), 485–492, DOI: 10.1080/ Quinolines and Related Heterocycles Using Formic Acid
00397919808005103. under mild Conditions, Catal. Sci. Technol., 2017, 7, 1981–
131 A. M. McKinney, K. R. Jackson, R. N. Salvatore, 1985, DOI: 10.1039/C7CY00437K.
E.-M. Savrides, M. J. Edattel and T. Gavin, A Rapid and 143 F. Ding, Y. Zhang, R. Zhao, Y. Jiang, R. L. Bao, K. Lin and
Efficient Method for the Reduction of Quinoxalines, L. Shi, B(C6F5)3-promoted hydrogenations of N-heterocycles
J. Heterocycl. Chem., 2005, 42, 1031–1034. with ammonia borane, Chem. Commun., 2017, 53, 9262–
132 J. Tan, W. Tang, Y. Sun, Z. Jiang, F. Chen, L. Xu, Q. Fan 9264, DOI: 10.1039/c7cc04709f.
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

and J. Xiao, pH-Regulated transfer hydrogenation of qui- 144 Z. Liu, Z. Wen and X. Wang, B(C6F5)3-Catalyzed Cascade
noxalines with a Cp*Irediamine catalyst in aqueous Reduction of Pyridines, Angew. Chem., Int. Ed., 2017, 56,
media, Tetrahedron, 2011, 67, 6206–6213, DOI: 10.1016/j. 5817–5820, DOI: 10.1002/anie.201702304.
tet.2011.06.067. 145 N. Gandhamsetty, S. Park and S. Chang, Boron-Catalyzed
133 J. Wu, C. Wang, W. Tang, A. Pettman and J. Xiao, The Hydrogenative Reduction of Substituted Quinolines to
Remarkable Effect of a Simple Ion: Iodide-Promoted Tetrahydroquinolines with Hydrosilanes, Synlett, 2017,
Transfer Hydrogenation of Heteroaromatics, Chem. – Eur. 2396–2400, DOI: 10.1055/s-0036-1588442.
J., 2012, 18(31), 9525–9529, DOI: 10.1002/ 146 P. Buschelberger, E. Reyes-Rodriguez, C. Schottle,
chem.201201517. J. Treptow, C. Feldmann, A. J. v. Wangelin and R. Wolf,
134 L. Zhang, R. Qiu, X. Xue, Y. Pan, C. Xu, H. Li and L. Xu, Recyclable Cobalt(0) Nanoparticle Catalysts for
Versatile (Pentamethylcyclopentadienyl)rhodium-2,2′- Hydrogenations, Catal. Sci. Technol., 2018, 8, 2648–2653,
Bipyridine (Cp*Rh-bpy) Catalyst for Transfer DOI: 10.1039/C8CY00595H.
Hydrogenation of N-Heterocycles in Water, Adv. Synth. 147 I. D. Alshakova, B. Gabidullin and G. I. Nikonov, Ru-
Catal., 2015, 357, 3529–3537, DOI: 10.1002/ Catalyzed Transfer Hydrogenation of Nitriles, Aromatics,
adsc.201500491. Olefins, Alkynes, and Esters, ChemCatChem, 2018, 10(21),
135 B. Sun, D. Carnevale and G. Süss-Fink, Selective N-cycle 4860–4869, DOI: 10.1002/cctc.201801039.
hydrogenation of quinolines with sodium borohydride in 148 A. Dubey, S. M. W. Rahaman, R. R. Fayzullin and
aqueous media catalyzed by hectorite-supported ruthe- J. R. Khusnutdinova, Transfer hydrogenation of carbonyl
nium nanoparticles, J. Organomet. Chem., 2016, 821, 197– groups, imines and Nheterocycles catalyzed by simple,
205, DOI: 10.1016/j.jorganchem.2016.07.010. bipyridine-based MnI complexes, ChemCatChem, 2019,
136 Q. Xuan and Q. Song, Diboron-Assisted Palladium- 11(16), 3844–3852, DOI: 10.1002/cctc.201900358.
Catalyzed Transfer Hydrogenation of N–Heteroaromatics 149 Y. Fu and J. Sun, HMPA-Catalyzed Transfer Hydrogenation
with Water as Hydrogen Donor and Solvent, Org. Lett., of 3-Carbonyl Pyridines and Other N-Heteroarenes with
2016, 18, 4250–4253, DOI: 10.1021/acs.orglett.6b01999. Trichlorosilane, Molecules, 2019, 24, 401, DOI: 10.3390/
137 Z. Wei, Y. Chen, J. Wang, D. Su, M. Tang, S. Mao and molecules24030401.
Y. Wang, Cobalt encapsulated in N-doped graphene 150 S. Kim, F. Loose, M. J. Bezdek, X. Wang and P. J. Chirik,
layers: an efficient and stable catalyst for hydrogenation of Hydrogenation of N-Heteroarenes Using Rhodium
quinoline compounds, ACS Catal., 2016, 6, 5816–5822, Precatalysts: Reductive Elimination Leads to Formation of
DOI: 10.1021/acscatal.6b01240. Multimetallic Clusters, J. Am. Chem. Soc., 2019, 141,
138 Y. Ji, K. Wei, T. Liu, L. Wu and W. Zhang, “Naked” 17900–17908, DOI: 10.1021/jacs.9b09540.
Iridium(IV) Oxide Nanoparticles as Expedient and Robust 151 Y. Duan, X. Du, Z. Cui, Y. Zeng, Y. Liu, T. Yang, J. Wen
Catalysts for Hydrogenation of Nitrogen Heterocycles: and X. Zhang, Homogeneous hydrogenation with a
Remarkable Vicinal Substitution Effect and Recyclability, cobalt/tetraphosphine catalyst: a superior hydride donor
Adv. Synth. Catal., 2016, 359(6), 933–940, DOI: 10.1002/ for polar double bonds and N-heteroarenes, J. Am. Chem.
adsc.201601370. Soc., 2019, 141, 20424–20433, DOI: 10.1021/jacs.9b11070.
139 Y. Xia, X. Sun, L. Zhang, K. Luo and L. Wu, Chem. – Eur. 152 S. Wang, H. Huang, C. Bruneau and C. Fischmeister,
J., 2016, 22, 17151–17155, DOI: 10.1002/chem.201604503. Iridium Catalyzed Hydrogenation and Dehydrogenation of
140 Y. Zhang, J. Zhu, Y. Xia, X. Sun and L. Wu, Metal-Free N–Heterocycles in Water under Mild Conditions,
Hydrogen Atom Transfer from Water: Expeditious ChemSusChem, 2019, 12(11), 2350–2354, DOI: 10.1002/
Hydrogenation of N-Heterocycles Mediated by Diboronic cssc.201900626.
Acid, Adv. Synth. Catal., 2016, 358(19), 3039–3045, DOI: 153 L. Troian-Gautier, J. D. Winter, P. Gerbaux and
10.1002/adsc.201600505. C. Moucheron, A Direct Method for Oxidizing Quinoxaline,
141 R. Adam, J. R. Cabrero-Antonino, A. Spannenberg, Tetraazaphenanthrene, and Hexaazatriphenylene Moieties
K. Junge, R. Jackstell and M. Beller, A General and Highly Using Hypervalent λ3 –Iodinane Compounds, J. Org.
Selective Cobalt-Catalyzed Hydrogenation of N-Heteroarenes Chem., 2013, 78, 11096–11101, DOI: 10.1021/jo401872e.
under Mild Reaction Conditions, Angew. Chem., Int. Ed., 154 L. M. Kabadwal, S. Bera and D. Banerjee, Iron-catalysed
2017, 56, 1–6, DOI: 10.1002/anie.201612290. alkylation of 2-methyl and 4-methyl azaarenes with alco-
142 J. R. Cabrero-Antonino, R. Adam, K. Junge, R. Jackstell hols via C–H bond activation, Chem. Commun., 2020, 56,
and M. Beller, Cobalt-catalysed Transfer Hydrogenation of 4777–4780, DOI: 10.1039/d0cc01593h.

Org. Chem. Front. This journal is © the Partner Organisations 2021


View Article Online

Organic Chemistry Frontiers Review

155 A. Jana, A. Kumar and B. Maji, Manganese catalyzed RSC Adv., 2021, 11, 3477–3483, DOI: 10.1039/
C-alkylation of methyl N-heteroarenes with primary alco- d0ra09994e.
hols, Chem. Commun., 2021, 57, 3026–3029, DOI: 10.1039/ 157 R. Chemboli, R. Kapavarapu, K. Deepti, K. R. S. Prasad,
D1CC00181G. A. G. Reddy, A. V. D. N. Kumar, M. V. B. Rao and M. Pal,
156 M. K. Zilla, S. Mahajan, R. Khajuria, V. K. Gupta, Pyrrolo[2,3 -b, ]quinoxalines in attenuating cytokine storm
K. K. Kapoor and A. Ali, An efficient synthesis of in COVID-19: their sonochemical synthesis and in silico
4-phenoxy-quinazoline, 2-phenoxy-quinoxaline, and /in vitro assessment, J. Mol. Struct., 2021, 1230, 129868,
Published on 10 March 2021. Downloaded by University of Prince Edward Island on 5/16/2021 8:23:27 AM.

2-phenoxy-pyridine derivatives using aryne chemistry, DOI: 10.1016/j.molstruc.2020.129868.

This journal is © the Partner Organisations 2021 Org. Chem. Front.

You might also like