Accepted Article: Chemistry-Sustainability-Energy-Materials

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Chemistry–Sustainability–Energy–Materials

Accepted Article

Title: Light-Mediated Carboxylation Using Carbon Dioxide

Authors: Chanjuan Xi, Zhengning Fan, and Zeyu Zhang

This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). This work is currently citable by
using the Digital Object Identifier (DOI) given below. The VoR will be
published online in Early View as soon as possible and may be different
to this Accepted Article as a result of editing. Readers should obtain
the VoR from the journal website shown below when it is published
to ensure accuracy of information. The authors are responsible for the
content of this Accepted Article.

To be cited as: ChemSusChem 10.1002/cssc.202001974

Link to VoR: https://doi.org/10.1002/cssc.202001974

01/2020
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
Light-Mediated Carboxylation Using Carbon Dioxide
Zhengning Fan,[a] Zeyu Zhang,[a] and Chanjuan Xi*[a,b]

Dedicated to Professor Mingyuan He on the occasion of his 80th birthday

[a] Z. Fan, Z. Zhang and Prof. Dr. C. Xi


MOE Key Laboratory of Bioorganic Phosphorus Chemistry & Chemical Biology
Department of Chemistry, Tsinghua University, Beijing 100084, China
E-mail: cjxi@tsinghua.edu.cn
[b] Prof. Dr. C. Xi
State Key Laboratory of Elemento-Organic Chemistry

Accepted Manuscript
Nankai University, Tianjin 300071, China

Abstract: Carbon dioxide (CO2) is a green and sustainable one- Strong nucleophilic reagents, such as Grignard reagents and
carbon source, which could be utilized in the production of various fine organolithium reagents, are initially used for the carboxylation
chemicals. In recent studies, the light-mediated carboxylation using CO2.However, these reagents are generally air-sensitive
employing CO2 has received considerable attention. We will introduce and show poor functional group compatibility due to their high
the photocarboxylation of substrates with CO2 to build novel C-C reactivity. With the emergence of transition-metal catalysis,
bonds in this review. The article is arranged based on the light-driven fixation of CO2 into various substrates could be established in a
reactive intermediates, including CO2 radical anion, substrate radical catalytic fashion with improved functional group tolerance.[5]
anions, carbanions, and M-C species. Most of the cases are under
However, these protocols usually require stoichiometric metallic
the topic of photoredox catalysis, with single electron transfer as the
main driving force. Some non-catalytic examples are also discussed
reductants (Mn, Zn et. al), causing environmental pollution.
to provide more mechanistic insights. What’s more, heating at elevated temperature is almost
compulsory, which may result in net CO2 emission. It is contrary
to our primary goal. Thus, it is still highly requisite to discover a
new route for green and environmental-benign carboxylation with
higher atom efficiency.
1. Introduction Taking into mind that CO2 is a fundamental synthon of
natural photosynthesis, it seems feasible to achieve artificial CO2
The growing impact of greenhouse effect, mainly from over- fixation with the aid of light, which can ensure sustainable growth
emitted CO2, has been an inevitable threat to the global of the human civilization. For light-driven utilization of CO2, two
environment. As a result, exploring intrigue CO 2 utilization mode major categories could be employed: one is photoreduction of
is at urgent demand to alleviate the deteriorating environmental CO2 into one-carbon resources, including CO, methanol, formic
situation.[1] Although the thermodynamic and kinetic inertness of acid, and methane, which has been widely reported in the
CO2 has set a rigid barrier for its utilization, chemists think it’s generation of energy products,[2i-k] another is photocarboxylation,
reasonable to employ it as an ideal C1 source due to its which is a green and sustainable method for fixing CO2 into
nontoxicity, cheapness, and natural abundance, compared with organic molecules. This category is still at its infancy until recent
CO, phosgene, and other C1 compounds. Numerous successful years. This review will focus on the light-mediated carboxylation
transformation of CO2 into valuable chemicals, such as based on photocatalysis by generating various reactive
carbonates, carbamates, carboxylic acids, lactones, formic acid, intermediates for the achievement of C-C bond formation with
formamide, and methanol have been realized.[2] Among them, the CO2. To our delight, this hypothesis has been successfully
transformation of CO2 into carboxylic acids, with basic C-C bond accomplished with organic halides, alkenes, alkynes, and even
formation, exhibits privileged synthetic requirement, since the alkanes in a number of publications. Although a few reviews have
carboxylic acids as building blocks frequently occur in introduced part of them in the photocarboxylation,[6] herein, we
pharmaceutical and bioactive molecules, [3,4] some examples are would like to summary the latest advances on light-mediated
shown in Figure 1. Besides, the carboxylic acids are prone to carboxylation using CO2 in the manner of reactive intermediates.
undergo decarboxylation under specific conditions, These species include: CO2 radical anion, light-driven reactive
demonstrating their potential in late-stage functionalization as carbanions from substrates as well as transition-metal assisting
good leaving groups. in generation of the reactive M-C species. This review will not only
provide a description of the current methods for the carboxylation,
but also highlight the remaining challenges in this area, without
losing sight of mechanistic considerations. The selected
examples highlighted in this review are meant to showcase our
conviction that these untraditional technologies might impact the
practice of organic chemistry in the formation of carboxylic acids.

2. Photocatalytic carboxylation modes of CO2


and substrates

Fig. 1. Examples of carboxylic acids in pharmaceutical and CO2 and most common substrates such as alkanes, alkenes,
bioactive molecules. alkynes, and organic halides do not absorb either visible or UV
radiation in the wavelengths of 200-700 nm, so utilization of these

1
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
substrates in photocarboxylation normally requires a 3. Photocarboxylation via CO2 radical anions
photosensitizer to absorb UV-vis radiation and then transfer an with substrates
electron to CO2 or/and the substrates. Hence, the carboxylation
process begins with excitation of the photosensitizer (PS). The
Once an electron is accepted by CO2, the generated CO2
photosensitizer first absorbs photons creating an excited
radical anion could then undergo radical addition or radical
photosensitizer (PS*) (Scheme 1). The PS* is reductively
coupling reactions. Yanagida and co-workers have reported using
quenched by an electron donor (D) to afford reduced PS •–. The
p-terphenyl as photocatalyst to achieve a photoreduction of CO2
reduced PS•– is capable of reducing CO2 to its radical anion
to formate via the formation of CO2•– with UV irradiation.[9]
(CO2•–), significantly lowering down the activation energy required
Employing this photocatalyst, Jamison and co-workers disclosed
for CO2 activation, thus setting the basis for promoting the
the linear-selective hydrocarboxylation of styrene in continuous
targeted C-C bond-forming with suitable substrates (Scheme 1,
flow system.[10] In optimization attempts, 1,2,2,6,6-
path a). Giving an electron to CO2 from PS* directly is also
pentamethylpiperidine (PMP) is selected as the best donor and
possible, but there are only a few cases of photocatalytic CO2
large excess amount of water is integral to minimize the formation
reduction using this mechanism.[7] This is because the potential
of dicarboxylate byproduct. A number of functional groups are

Accepted Manuscript
for CO2 reduction is highly negative (E1/2 red = -2.21v vs. SCE in
tolerated under the condition, including unprotected OH, NH 2
DMF) and the catalysts or substrates usually require the PS •– to
groups and heteroaryls. α- and β-substituted alkenes also
have higher reduction potentials than the corresponding PS*.
manifest moderate reactivity. When replacing the water by D2O,
Alternatively, the reduced PS•– is also acting as an oxidant
high deuterated ratio at benzylic H suggests that a benzyl radical
promoting single electron transfer (SET) or twice single electron
might be involved in the reaction pathway. The mechanism is
transfer toward substrates to form radical anions or carbanions,
illustrated in Scheme 2. Initial SET between PMP and the excited
which could trap CO2 to afford the carboxylic acids (Scheme 1,
p-terphenyl (p-terphenyl*) generates its radical anion (p-
path b). It is frequently confounding that ‘photosensitizer’ (PS) and
terphenyl•–), which transforms an electron to CO2 to form CO2•–
‘photocatalyst’ (PC) are not differentiated in publications, which
that subsequently attacks to styrene to afford benzyl radical
does not cause serious semantic error in many contexts. We will
adduct As2. The second SET works consecutively to produce the
refer to the compounds, which directly react with CO2 or
benzyl anion Bs2. Protonation will afford the final linear product
substrates after light irradiation, as PS. These compounds are
Cs2 with high selectivity.
generally fixed into the products and cannot be recycled after the
reaction. On the other hand, PS simply serves to capture the
incident light energy and then transfers this energy to other moiety,
which are as PC. This kind of catalyst can keep its original
chemical structure unchanged after the reaction and the reaction
mode is typically called photoredox catalysis. [8] As the reactions
included in this review are mostly carried out with photoredox
protocol, we will use PC rather than PS.

Scheme 2. Photo-assisted linear-selective hydrocarboxylation of


styrene.

Scheme 1. Photocatcatalytic carboxylation modes of CO2 and Indeed, the hydrocarboxylation of alkenes can be classified
organic substrates. as net-reductive, where external reductant is required to achieve
the electron balance. If radical-involved difunctionalization
Various transition metal complexes, macrocycles, and protocols can be exquisitely utilized, synergistic construction of
aromatic hydrocarbons could act as either photosensitizers multiple C-C and/or C-heteratom bonds is of great synthetic
and/or photocatalysts for the carboxylation with CO2. Polypyridyl potential.[11,12] More recently, Yu and co-workers disclosed the
iridium and ruthenium complexes are the most widely applied as visible-light induced thiocarboxylation of alkenes employing
photocatalysts. In addition, some organic molecules are also simple FeCl3 as catalyst with excellent regio- and chemo-
utilized as photocatalysts (Fig. 2). selectivity (Scheme 3).[13] β-Thiophenylpropionic acids are
obtained. The protocol also works well with acrylates. Since the
carboxyl group is added in anti-Markovnikov fashion, the authors
surmise that the reaction involves the formation of an indivisible
Fe/S complex in-situ, which is supported by UV/Vis and CV
measurements. The Fe/S complex is of great reducing ability to
transfer CO2 into CO2•– and thiol into thiol radical As3, respectively.
Addition of CO2•– to olefin generates alkyl radical Bs3, which is
then trapped by As3 to afford product Cs3 after hydrolysis.
Replacing the aryl thiol by its potassium salt leads to the same
product, indicating that KOtBu mainly plays as a base in
deprotonation process and may probably facilitate the formation
of a Fe/S complex. Control experiments suggest that the C-C
bond formation is prior to the C-S bond.

Fig. 2. Typical structures of photocatalysts in carboxylation.

2
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
changed from piperdine to other simple aliphatic amines. The
mechanism is speculated with the initial excitation of p-terphenyl
under UV light and followed by a SET with a tertiary amine to
produce reductive p-terphenyl•– and amine radical cation As6.
Simultaneously, the single electron is transferred to CO2 forming
CO2•–, which reacts with α-amino carbon radical Bs6 to afford the
α-amino acid Cs6 after protonation (Scheme 6).

Scheme 3. Iron-promoted thiocarboxylation of olefins by


photoredox catalysis.
Iodide anion has strong absorption bands at UV region (λ max
= 247.2 nm), being an ideal reagent for charge transfer to CO 2.
González-Núñez and co-workers uncovered a unique

Accepted Manuscript
iodocarboxylation method of alkenes with UV light irradiation.[14]
Lithium iodide is selected as the iodine source in this case. The
reaction is applicable to a number of electron deficient alkenes,
including methyl acrylate, acrylamide, and cyclopentanone. The
mechanism is illustrated in Scheme 4. Compared with alkenes,
CO2 is more likely to undergo SET with iodine anion, resulting in
an iodine radical and CO2•–. Then CO2•– adds reversely to the Scheme 6. Photosynthesis of α-amino acids from tertiary amines
alkene to produce the distonic radical anion As4, which couples in continuous-flow.
with the generated iodine radical to give the designed
iodocarboxylate Bs4. Although it is difficult for the PS* to directly give an electron
to CO2 to afford CO2•–, Kubiak and co-workers have devised a
nickel complex [Ni3(μ3-I)2(dppm)3] (As7) as photosensitizer, which
is capable of reducing CO2 to CO2•– under UV light.[17] The
mechanism is shown in Scheme 7. Cyclohexene (Cs7) captures
the CO2•– to form monocarboxylate Ds7, which is subseuently
traped by another CO2 molecule to form cis- and/or trans- isomers
Es7. The second addition of CO2 is a reverse Kolbe reaction, which
is inclined to be reversible. As a result, thermodynamically more
Scheme 4. UV-light promoted iodocarboxylation of activated stable trans-isomer is prevailing (63% for trans, 37% for cis).
alkenes with LiI and CO2. Finally, reduction of Es7 by As7 or CO2•– occurs to furnish the
formation of dicarboxylate Fs7. This reaction represents a scarce
Ogawa and co-workers reported the photoinduced example where the PS is transformed to its radical cation Bs7
carboxylation of alkyl halides using SmI2/Sm and a tungsten lamp without external reductant albeit the reaction is not catalytic.
as light source.[15] A number of alkyl carboxylic acids and
dodecanedioic acid could be obtained from alkyl halides at
ambient pressure and room temperature. Phenylacetic acid could
also be synthesized in 35% yield using benzyl chloride. The
reaction mechanism is shown in Scheme 5, the first step is the
reduction of alkyl halides by Sm to give alkyl radical As5. At the
same time, CO2 is coordinated to SmI2 to form Bs5 and
subsequently photoinduced SET generates the Sm-coordinated
CO2 radical Cs5. Coupling of As5 and Cs5 affords the carboxylic
acid Ds5 after hydrolysis. Control experiments show that the
reaction does not proceed in the dark, precluding the formation of
R-SmI2 and following CO2 insertion into R-Sm bond as a light-free
pathway.

Scheme 7. UV-light mediated 1,2-dicarboxylation of cyclohexene


via CO2 radical anion.

4. Photocarboxylation via reactive


intermediates light-driven from substrates
4.1 Carboxylation via radical anions light-driven from
Scheme 5. Light-driven reductive carboxylation of alkyl halides substrates
with CO2. Polyaromatic hydrocarbons are well-known organic
photosensitizers, which can be converted into the corresponding
Jamison and co-workers reported a carboxylation of tertiary radical anions under the light. The generated radical anions would
amines in continuous flow setup under UV light to produce α- capture CO2.
amino acids.[16] The reaction exhibits high regioselectivity at Tazuke and co-workers disclosed a carboxylation of
benzylic position (>20:1), which is not accessible under traditional phenanthrene in the presence of various aliphatic and aromatic
batch conditions. Both electron-rich and electron-neutral arenes amines in DMSO.[18] This is the first example of light-driven fixing
are suitable substrates and the amine moiety could also be CO2 in a non-biological system. 9,10-Dihydrophenanthrene-9-

3
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
carboxylic acid was obtained in 46% yield under 300 W high- similar radical anions with UV light. Under this premise, it comes
pressure mercury lamp irradiation and using N,N-dimethylaniline the idea that upon utilizing photoredox catalyst, the substrates can
(DMA) as an electron donor. The proposed reaction mechanism be step-wisely transformed into radicals and then carbanions to
is shown in Scheme 8. The phenanthrene itself is excited in the afford alternative strong nucleophiles, which could react with CO2.
first stage, then DMA serves as reductant to produce Alkenes and imines are latent nucleophile precursors if two
phenanthrene radical anion As8, which captures CO2 to produce electrons can be sequentially donated from the photocatalyst. Yu
Bs8. Solvent is considered to be the general hydrogen donor and co-workers recently reported a carboxylation of enamides
which could convert Bs8 to the final product. It was found that and imines using 4CzIPN as photocatalyst (Scheme 10).[26] This
highly polar solvent such as DMSO and DMF is critical for the umpolung carboxylation is unable to proceed in the absence of
success of the reactions while protic solvents are devastating light, where β-substituted acrylic acid is the predominant product.
since they can react with the radical anion. Additives such as The reaction shows superb functional group tolerance and both
cumene and decaline could significantly increase the quantum enamides and imines with different alkyl or aryl substituents work
yield by suppressing electron back-transfer.[19] Mechanistic smoothly to provide α,α-disubstituted α-amino acids, which are
experiments from Neckers′ group[20] suggest that the addition of important motifs in life-relevant molecules. Shortly after, Walsh

Accepted Manuscript
CO2 to phenanthrene anion radical is reversible as higher yield is and co-workers revealed a similar strategy focusing on imine
obtained at higher gas concentration. transformation using Ir(ppy)2(dtbbpy)PF6 as photocatalyst
(Scheme 10).[27] In this protocol, amine Cy2NMe is chosen to
replace Hunig’s base (iPr2NEt) as a reductant. Upon these subtle
changes, the protecting group could be expanded to benzyl, alkyl,
and aryl group.
To understand the reaction pathway, Walsh’s work is taken
as an example in Scheme 10. The reaction is initiated by the
excited photocatalyst IrIII*, which was reduced to IrII by the amine.
A 2-center-3-electron intermediate, As10, is formed by
coordination of imine with the amine radical cation Ds10, promoting
the SET with IrII to form radical anion Bs10 or Cs10. The high
reactivity of resonance form Cs10 may activate a HAT procedure
with amine radical cation Ds10 to afford carbon anion Es10 and
imine cation Fs10. Subsequent CO2 insertion generates the
Scheme 8. Photocarboxylation of phenanthrene. carboxylate, and insoluble salt Gs10 is formed by reacting with
dicyclohexylamine, which comes from the hydrolysis of Fs10.
After this breakthrough, the strategy is proved to be versatile Computational studies also show that the carbon atom of the
for the carboxylation using other polyaromatic hydrocarbons such resonance pair exhibits more anionic character than the nitrogen
as anthracene and naphthalene.[21-23] Nevertheless, these atom, favoring a reaction pathway involving Cs10.
reactions are still far away from gratifying results because of the
lower yields and selectivity.
With the decrease of conjugated system, styrenes are not as
energetic as their fused-ring counterparts. Ito and co-workers
reported the photocarboxylation of phenylethylenes in the
presence of aromatic amines such as N,N-diethylaniline and CO2,
under a high-pressure mercury lamp through Pyrex filter. [24] Only
trace of acrylic acids were observed. In order to improve the yield,
N,N,N′,N′-tetramethylbenzidine (TMB) is employed as a suitable
photocatalyst in the carboxylation of alkenes such as
diphenylethylene (DPE).[25] Diphenyl acrylic acid (As9) could be
generated in 71% yield (Scheme 9). Substantial photophysical
measurements show that TMB is a much more effective
photosensitizer. Mechanistically, DPE is reduced by TMB to its
radical anion Bs9. Subsequent CO2 capture forms distonic radical
anion Cs9, which then undergoes back-electron transfer to TMB,
generating zwitterionic Ds9. Finally, Ds9 will be transformed into
As9 as a non-reductive product.

Scheme 10. Visible-light-assisted umpolung carboxylation of


enamides and imines.

Yu and co-workers further described the coupling between


electrophilic tetraalkyl ammonium salts and CO 2.[28] Not only the
protocol is applicable to primary and secondary ammonium salts
on a gram scale, but also it could accomplish the challenging for
tertiary ammonium salts. Apart from benzyl ammonium salts,
Scheme 9. Photocarboxylation of diphenylethylenes allylic C-N bonds could be functionalized under the standard
condition with lower regioselectivity. The reaction mechanism is
4.2 Carboxylation via reactive carbanions light-driven from shown in Scheme 11. The ammonium salt first releases the amine
substrates As11, which undergoes SET with the photocatalyst to generate the
In the former section, the addition of C-centered radical anion alkyl radical Bs11 while forming the amine cation Cs11.
to CO2 is viable as the substrates bear at least two aromatic rings, Deprotonation of Cs11 may generate α-amino radical Ds11, which
thus, the radical anion’s activation energy is hugely decreased. It is also a latent electron donor to promote the second SET process
is less probable for simpler non-α-substituted styrenes to form to form imine Es11 and carbanion Fs11, which reacts with CO2 to

4
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
afford carboxylic acid after hydrolysis. This contribution not only process with HCO2K to generate the CO2•–. The aryl halide is
proves the umplong transformation of nitrogen cation to carbanion reduced by the in-situ generated CO2•– to aryl radical Bs13, which
under reductive condition, but also envisions an environment- subsequently adds to the double bond to generate a benzyl
benign reaction design strategy. radical Cs13. SET between the IrII and Cs13 produces the benzyl
anion Ds13 and IrIII. Finally, nucleophilic addition of the carbanion
to CO2 generates methyl carboxylate Es13 after methylation. This
protocol provides a novel method for producing carbon dioxide
radical anion under less-energetic conditions.

Accepted Manuscript
Scheme 11. Light-mediated cross-electrophile-coupling of
tetraalkyl ammonium salts with CO2.

The difunctionalization of olefins with CO2 could also be


established. Martin and co-workers reported a carbocarboxylation
of aromatic olefins under photoredox condition (Scheme 12).[29]
The reaction could be conducted under atmospheric pressure of
CO2 with some simple radical precursors, such as CF3SO2Na, Scheme 13. Photo-induced arylcarboxylation of styrenes with
CHF2SO2Na, trifluoroborates, and oxalates. For the former two, CO2 and aryl halides.
Ir(ppy)2(dtbbpy)PF6 is the optimal choice while for the latter two,
the more oxidizing [Ir[dF(CF3)ppy]2(dtbbpy)PF6 shows better Recently, a dearomative carboxylation of indoles with
performance. The substrate scope is rather broad in both α- and visible-light photoredox catalysis is reported by Yu’s group. [31] The
β-substituted styrenes, which often display significant steric effect. protocol exhibits distinguished regioselectivity. This reaction
Mechanistically, the reaction is triggered by the light to produce shows low loading of photocatalyst, generally good yields, mild
oxidative IrIII*, which abstracts an electron from the radical reaction conditions, good functional group tolerance, and broad
precursor to release the radical entity As12 and reductive IrII. Then substrate scope. Mechanistic studies indicate that the benzylic
radical addition to olefin takes place, offering carbon-centered radical Cs14 and anion Ds14 might be generated as the key
radical Bs12. Subsequent SET between radical Bs12 and IrII intermediates. As is illustrated in Scheme 14, 4CzIPN is
produces the benzyl carbanion Cs12, which subsequently reacted reductively quenched by iPr2NEt to afford 4CzIPN•–, which is
with CO2 to afford desired product. The reversible reaction of CO2 confirmed by Stern-Volmer experiments. It could undergo SET
with the transient benzyl radical followed by SET cannot be with aryl bromide to produce radical anion As14. Fracture of this
excluded as an alternative pathway. unstable radical anion gives aryl radical Bs14, which adds to the
intramolecular double bond. The formed benzylic radical Cs14 is
reduced to benzylic anion Ds14 by another SET process. This step
is confirmed by deuterium labeling experiments. Nucleophilic
addition of the benzylic anion Ds14 to CO2 and following hydrolysis
will generate the arylcarboxylic acid as a difunctionalization
product.

Scheme 12. Visible-light driven dicarbofunctionalization of


alkenes with radical precursors.

More recently, Li and co-workers reported a multi-component


arylcarboxylation employing styrenes, CO2, and aryl halides.[30] In
this scenario, the combination of DABCO and HCO2K as HAT
catalyst and reductant is critical to the generation of CO 2•–. The
protocol exhibits broad functional group tolerance for both
olefines and arylhalides. Control experiments reveal that the
reaction provides dicarboxylated product without aryl halides,
suggesting a mechanistic pathway involving CO 2 radical anion.
The tentative mechanism is depicted in Scheme 13. Initially, the
excited IrIII* is reductively quenched by DABCO to give a radical
cation As13, which undergoes a hydrogen atom transfer (HAT)

5
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW

Accepted Manuscript
Scheme 15. Photoredox difunctionalization of alkenes with CO 2
and silanes or C(sp3)-H alkanes.

Xi and co-workers lately reported the construction of γ-amino


acids by CO2-involved dicarbofunctionalization with photoredox
catalysis.[33] The protocol is proved to be friendly with a number
of electron-deficient olefins and tertiary amines, where steric α-
substituted styrenes are tolerated to give moderate yields.
Mechanistically, radical cation As16 is generated from the
reductive quenching of excited photocatalyst (4CzIPN*) by N,N-
Scheme 14. Photoredox dearomatization of indoles with CO 2 to dimethylaniline, which is deprotonated to generate α-aminalkyl
produce aryl carboxylic acids. radical Bs16 in the presence of base. Then it is bonded to the
terminal position of styrene by radical addition and forms a C-
In 2018, Wu and co-workers reported a silacarboxylation of centered radical intermediate Cs16. A second SET between
olefins with silanes and CO2.[32] 4CzIPN, quinuclidine-3-yl-acetate 4CzIPN•– and Cs16 will deliever the benzylic carbanion Ds16, which
(H-OAc), and hydrosilanes are incorporated as photocatalyst, is stablized in the form of a chelating species with lithium.
HAT catalyst, and silicon source, respectively. Electron deficient- Subsequent nucleophilic addition to CO2 or LiCl-activated CO2
alkenes could proceed smoothly under the optimized condition in and protonation will afford the desired γ-amino acid Es16.
good to moderate yields and the extended silane scope suggest
that tris(trimethylsilyl) silane appears to be the best choice. With
the achieved effort, they also proposed the applicability of this
protocol to C(sp3)-H bonds. Delightfully, this hypothesis works
successfully under the same condition, affording protected amino
acids. The mechanism is proposed as shown in Scheme 15.
Excited 4CzIPN (4CzIPN*) is reductively quenched by the HAT
catalyst (H-OAc), yielding 4CzIPN•– and radical cation As15.
Subsequent HAT process between As15 and the Si-H bond affords
a silyl radical Bs15, which adds to alkene to form Cs15. The Cs15
could be reduced by 4CzIPN•– to carbanion Ds15, which attacks
CO2, leading to the carboxylate Es15. The Es15 abstracts a
hydrogen from the ammonium salt Fs15 finally closes the HAT
cycle while releasing the carboxylic acid.

Scheme 16. Light-induced dicarbofunctionalization of styrenes


with amines and CO2.

As difunctionalization of alkenes with C-C, C-Si, and C-S


bond construction are established, development of a similar
protocol involving phosphine moiety becomes a reasonable idea.
Yu’s group reported a photoredox phosphonocarboxylation of
alkenes to afford β-phosphono carboxylic acids (Scheme 17).[34]
This attempt is extremely challenging since hydrophosphination,
hydrocarboxylation, and C-H bond carboxylation are all latent
competitory side-reaction pathways. Notably, less reactive diaryl
phosphites and dialkyl phosphine oxides are also capable of
delivering desirable products, suggesting the versatility of this
methodology.

6
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
In the mechanism, 4CzIPN is aroused to the excited state involves organic dye 4CzIPN as photocatalyst and iPr3SiSH as
upon illumination, which undergoes SET with HP(O)R 2 to HAT catalyst at 0 oC. For some substrates with higher reduction
generate 4CzIPN•– and phosphonyl radical As17. Then, addition of potential, analogous photocatalysts 3DPAFIPN and 3DPA2FBN
the As17 to the double bond affords radical intermediate Bs17, with stronger reducing ability are employed to replace 4CzIPN. To
which is further reduced to carbanion Cs17 by the second SET demonstrate the synthetic applicability of the method, scaffolds of
process. Following CO2 insertion and protonation will generate several antibiotics are successfully synthesized, including
the difunctionalized carboxylic acid Ds17. Flurbiprofen, Ibprofen, and Naproxen.
While performing mechanistic investigation, the authors
found that one cyano group of 4CzIPN is replaced by the
ethylbenzene to form 4CzPEBN, which actually functions as the
photocatalyst. The excited 4CzPEBN then undergoes SET with
As19 to produce the thiol radical Bs19 and 4CzPEBN•–. This
electrophilic radical then abstracts a hydrogen from the benzylic
position of ethylbenzene to close the HAT cycle. The in-situ

Accepted Manuscript
generated benzylic radical Cs19 is further reduced to its carbanion
Ds19 and then it captures CO2. The observed kinetic isotope effect
suggest that the C-H bond cleavage might be the rate-determining
step. This reaction opens a new era in the carboxylation using
CO2.

Scheme 17. Visible-light driven difunctionalization of alkenes with


H-P(O) compounds and CO2.

The above-mentioned protocols generally requires an


activated olefin, which usually bears a aryl group or an amide at
α-position of alkenes. Very recently, Yu and co-workers reported
a reaction of unactivated alkenes with CO2 via visible-light
photoredox catalysis to afford series of carboxylic acids. [35] This
reaction combines radical attack to alkenes, hydrogen atom
transfer (HAT) and remote C–H functionalization. The radical
precursors, including CF3SO2Na, CHF2SO2Na, and HP(O)Ph2,
are used in the reaction. Mechanistically, the reaction is initiated
by the SET between IrIII* and CF3SO2Na, delivering Scheme 19. Photocarboxylation of benzylic C-H bonds.
trifluoromethyl radical As18. Addition of the radica A s18 to the
alkene produces a carbon-centered radical Bs18, which undrgoes 4.3 Carboxylation via other energetic species light-driven
a significant 1,5-HAT process to produce benzylic radical Cs18. from substrates
This step is believed to be rate-limiting according to kinetic isotope The aforementioned examples mainly include the generation
experiments. Afterwards,the radical Cs18 is further reduced to a of radical anions and/or carbanions for capturing CO 2. In some
benzylic anion Ds18 while releasing ground state IrIII. Facile CO2 preliminary non-catalytic studies, other transient intermediates
capture by Ds18 and following protonation will generate the are also generated as reactive species, including carbenes, ylides
carboxylic acid Es18. DFT calculations reveals that Cs18 is and 1,3-dienes.
stablized by both the phenyl ring and the amino group. Carbenes are latent nucleophilic, which can attack the
electrophilic carbon of CO2 with subsequent nucleophilic
cyclization to construct a labile 3-membered ring (Scheme 20,
top). The light-induced heterocycle formation with CO2 is reported
by Barlett in 1970.[37] They discovered that diphenyl
diazomethane (As20) can react with dry ice in CFCl3 to generate
an unstable α-lactone Cs20, which is converted to the
corresponding polyester Ds20 in 40% yield. Indeed, the lactone
exhibits strong 1,3-dipolar character even at -100 oC due to the
delocalization of positive charge by the two phenyl rings, thus
undergoing rapid zwitterionic polymerization. Sander and co-
workers reported the photoreaction of carbene with CO2 under
step-wise irradiation with different wavelength. [38] In the first stage,
the precursor As20 is converted to the carbene intermediate Bs20
under 543 nm light. This moiety shows no reactivity toward CO 2
under -263 oC, but it turns more active at -243 oC or when the light
source is changed to 436 nm, producing Cs20. Photolysis of this
α-lactone under UV light releases carbon monoxide (CO) and
diphenyl acetone. These findings provide a new perspective for
Scheme 18. Light-driven remote difunctionalization of the spin-forbidden reactions of triplet carbenes, which actually
unactivated alkenes with CO2 exhibits singlet activity.
CO2 can also be recognized as a dienophile that can undergo
Direct carboxylation of C(sp3)-H bonds is a more dreaming Diels-Alder (D-A) like reaction. 1,3-Dipolar species, for example,
reaction. König and co-workers reported a novel benzylic C-H ylides exhibiting zwitterionic character is considered to be an ideal
bond carboxylation (Scheme 19).[36] The optimized condition

7
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
cycloaddition companion with CO2 (Scheme 20, bottom). Schmid It is widely-accepted that CO2 could insert into M-C bonds. In
and co-workers disclosed the photoinduced carboxylation of the initial reports, some pre-formed organometallic species are
azirines in 1972.[39] The strained 3-membered ring Es20 is released not capable for CO2 capture under light-free condition. However,
upon light irradiation to produce a benzonitrile methylene ylide upon light irradiation, CO2 insertion takes place with some
Fs20. This reactive 1,3-dipolar moiety is prone to undergo additive assistance due to property changes in the metal-carbon
cycloaddition with CO2 to form oxazolinone Gs20. Complementary bonds.
to this pioneering work, Padwa further extended this protocol to Inoue reported the first photofixation of CO 2 by a metal
other substituted azirines to afford oxazolinone derivatives in complex, namely tetraphenylporphinatoaluminium ethyl (TPPAlEt,
excellent yield.[40] To achieve this, CO2 saturated benzene and As22), although the lower yield of carboxylates was observed
water-cooled mercury arc lamp are utilized as solvent and light (Scheme 22).[43] Notably, the reaction does not proceed in the
source. absence of 1-methylimidazole. The probable coordination of 1-
methylimidazole to aluminum is likely to increase the
nucleophilicity of the ethyl group.

Accepted Manuscript
Scheme 22. Reaction of CO2 with tetraphenylporphyrin aluminum
ethyl.

Later on, a novel photofixation system with ketones to furnish


β-ketocarboxylic acids was also reported. (Scheme 23).[44] The
metalloporphyrin skeleton remains the same in this report but the
ethyl group is substituted by a diethylamine (As23). The reaction is
proposed to undergo through enolates complexes Bs23.
Scheme 20. Early examples of photo-induced CO2 fixation via
energetic species.

In 2015, Murakami and co-workers reported a carboxylation


of o-alkylphenyl ketones under UV light.[41] Similar to the previous
noncatalytic carboxylation in this section, the driving force comes
from the internal bond of molecules, which could be transformed
into reactive moieties after light absorption. Ketones with γ-H
could undergo H abstraction to produce transient 1,4-biradical
As21 upon irradiation, which will then rearrange to o-
quinodimethane Bs21 and Cs21. The formation of these energetic
1,3-diene species is well-known as Norrish Type II reaction.[42] In
this case, the Z-isomer Cs21 is prone to return to the initial ketone
via 1,5-H shift. On the contrary, E-isomer Bs21 possessing a longer
lifetime goes through [4+2] cycloaddition with CO 2 to afford a six- Scheme 23. Synthesis of β-ketocarboxylic acids from CO2 via
membered ring adduct Ds21. Carboxylic acid Es21 is obtained after enolate complexes.
ring-opening process. Notably, most substrates show good
reactivity. Having succeeded in disclosing this scenario, Inoue and co-
workers found that when alkyl aluminum complex (As24) was
treated with α,β-unsaturated esters (Bs24) under light, malonic
acid derivatives were obtained after work-up (Scheme 24).[45]
Irradiation is essential in the first step according to previous
reports. Further mechanistic investigation ruled out the direct
insertion of CO2 into Al-R bond or cascade polymerization of Bs24
with Cs24. Notably, the reaction proceeded well using 1 mol% As24
with addition of stoichiometric dialkyl zinc as transmetallation
reagent. The formed Al-carboxylate Ds24 can be transformed to
the corresponding zinc salts Es24 while regenerating As24. This is
the first attempt of carrying out metal/light dual catalyzed
carboxylation with CO2.

Scheme 21. Carboxylation of o-alkylphenyl ketones under UV-


light.

4.4 Carboxylation via pre-formed organometallic reagents


under light irradiation

8
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
the RhI-H intermediate (As26), avoiding the employment of metallic
reductants. The proposed mechanism initiates with the
hydrometallation of alkenes with As26 to form a π-benzyl rhodium
complex Bs26, which reacts with CO2 to form Cs26. The next step
involving accepting two electrons and two protons generates RhIII
complex Ds26, which undergoes a reductive elimination to give the
carboxylic acid Es26 and regenerate As26. Electron-deficient
aliphatic or aromatic carboxylic acids can be obtained in moderate
yields. Later, improved conditions are provided by the same group
by using a benzimidazoline derivative as sacrificial donor.[52]

Accepted Manuscript
Scheme 24. Carboxylation of α, β-unsaturated esters catalyzd by
TPPAl.

Guilard and co-workers reported a CO2 insertion into the


methyl-indium bond of complex (P)InCH3 (As25) in the presence of
pyridine (Scheme 25, (a)).[46] In this reaction, ESR studies support
the existence of an indium radical, indicating its addition to CO 2
instead of initial CO2 coordination to the metal center. Cooper and
co-workers reported a reaction of titanocene with CO2 to afford Ti-
carboxylate Ds25 under 300-750 nm irradiation (Scheme 25,
(b)).[47] The Ti-carboxylate Ds25 displays a large downfield shift for
the methyl carbon adjacent to fixed CO2, while a common upfield
shift for methyl proton. This indicates an unusual polarization of
the methyl-titanium bond, favoring CO2 insertion.
Scheme 26. Visible-light-driven hydrocarboxylation of olefins with
Rh(I) catalyst.

Combining the synthetic demand of both linear and branched


carboxylic acids, König and co-workers reported a regioselective
hydrocarboxylation of styrenes using organic dye 4CzIPN as
photoredox catalyst, low-cost nickel catalyst and Hantzsch ester
(HEH) as a reductant and hydrogen source (Scheme 27).[53] For
Markovnikov addition, neocuproine is chosen as the optimal
ligand and the ligated complex L1·NiBr 2 (L1 = neocuproine)
exhibits superior catalytic performance. On the contrary, a
Scheme 25. Photo-induced CO2 insertion into metal complexes: phosphine ligand, namely 1,4-bis(diphenylphosphino)butane
(a) indium complex; (b) titanium complex. (dppb, L2), successfully promotes the generation of linear product
with adjustment on HEH and base. In this case, the reaction
generally gives linar products in lower yield than the branch
5. Photocarboxylation via light/transition-metal isomers.
dual catalysis The photoredox cycle of 4CzIPN produces single electron,
which could convert the nickel catalyst to Ni 0 species (As27), and
starting the catalytic process. For Markovnikov
The carboxylations in section 4.4 are mostly carried out with
hydrocarboxylation, tendency to form nickel hydride intermediacy
stoichiometric metallic reagents under light irradiation. Transition-
Bs27 pushes forward the oxidation of HEH, in which process
metal catalyzed carboxylation has been proved out to be an
proton is released. This hypothesis is supported by detection of
effective method for the production of carboxylic acids. Inspired
hydrogen gas and both nickel catalyst and 4CzIPN are
by this, several reports suggest that combination of the
compulsory for the evolution of dihydrogen. Following
photoredox cycle with transition metal catalysis is en route to
hydrometallation of alkene will provide the π-benzyl NiII complex
greener methods for synthesis of carboxylic acids using CO2. In
Cs27. The Cs27 then undergoes SET with 4CzIPN•– to produce the
the previous metal-free photoredox protocols, some electron-rich
NiI complex (Ds27). The low-valent metal center favors CO2
C-centered radicals show lower activity to form the corresponding
insertion and the nickel carboxylate Es27 is further reduced to the
carbanions due to elevated reduction potential.[48] Replacement of
branched carboxylate Fs27 by the second SET process. On the
the radicals by M-C bonds can probably facilitate the insertion of
other hand, use of L2 leads to the coordination of As27 with CO2,
CO2, providing a broader substrate scope with higher efficiency.
giving Gs27. Subsequent oxidative cyclization of styrene produces
5.1 Photocarboxylation via light/transition-metal dual
nickelalactone Hs27, which experiences ring-opening to form the
catalysis from alkenes or alkynes
linear carboxylate Js27 as final product.
Hydrometallation is a common strategy in the
functionalization of alkenes.[49,50] Iwasawa’s group first reported a
photocarboxylation using CO2 via Ru complex as photocatalyst
and Rh complex as metal catalyst (Scheme 26).[51] The reaction
could be mediated by in-situ generated Rh-H moiety. iPr2NEt is
used as sacrificing electron donor and proton donor to produce

9
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW

Accepted Manuscript
Scheme 28. Co/Ir dual catalyzed hydrocarboxylation and [2+2+2]
cycloaddition of alkynes.

5.2 Photocarboxylation via light/transition-metal dual


Scheme 27. Regioselective hydrocarboxylation of styrenes by catalysis from organic (pseudo)halides
photoredox / nickel dual catalysis. Functionalization of organic halides is one of the fundamental
areas in organic chemistry. Martin and co-workers realized
The photoredox carboxylation of alkenes with CO 2 has palladium-catalyzed direct carboxylation of aryl bromides with
received considerable attention in recent years, while the alkyne CO2 in the presence of Et2Zn.[56] Following this path, reductive
counterpart is still less investigated. Wu and co-workers reported carboxylation of (pseudo)halides has received more attention with
a hydrocarboxylation of alkynes with CO 2 by light/cobalt dual various substrates.[57]
catalysis (Scheme 28).[54] Subsequent isomerization and tandem In 2017, Iwasawa and Martin co-reported the
cyclization or group migration could generate a bank of useful photocarboxylation of aryl halides with palladium catalysts
heterocycles. For hydrocarboxylation of aklynes, CO2 inserts at (Scheme 29). This protocol works well with a variety of chlorides
the less steric site. Intriguingly, when alkyl terminal alkynes are and bromides in moderate to good yields. [58] Ir(ppy)2(dtbbpy)PF6,
employed, [2+2+2] cycloaddition takes place to produce 2- Pd(OAc)2, and Cs2CO3 are employed as photocatalyst, catalyst,
pyrones as unprecedented products. To overcome the selectivity and base, respectively. Hunig’s base, iPr2NEt, is used as
problem, an ancillary cyclization-directing group (OMOM or reductant that can avoid the addition of metallic reductants. The
NHBoc) at the ortho-position of phenylbutylacetylene was authors propose that the excited Ir III* is reductively quenched by
introduced. The cyclization proceeds smoothly under standard iPr2NEt to afford IrII, which then undergoes double SET with Pd II
condition to generate coumarin and 2-quinolone derivatives in to produce reactive LnPd0 species (As29) and regenerate IrIII.
good yields. Oxidative addition of aryl halides to As29 produces PdII complex
Control experiments suggest that isomerization step cannot Bs29. Afterwards, two mechanistic pathways are plausible: a) CO2
proceed without light or photocatalyst. In this scene, the Co II insertion first takes place to produce Cs29, followed by single
complex is reduced to reactive CoI (As28) by IrII, which is electron reduction to form PdI complex Ds29; b) CO2 is initially
generated from classical photoredox cycle with iPr2NEt as coordinated to the Pd center, and the formed complex Es29 is
electron donor. Oxidative cyclization of As28, CO2, and alkyne then reduced to PdI complex Fs29, while CO2 insertion is at the end.
affords the five-membered ring intermediate Bs28, which is Further single electron reduction of Ds29 to As29 closes the
protonated by the amine radical cation Cs28 to generate cobalt catalytic cycle and carboxylate Gs29 is obtained. Notably, the
carboxylate Ds28. The Ds28 undergoes further SET with IrII and yields for using aryl chlorides are similar to aryl bromides.
transmetallation with ZnBr2 to release carboxylate Es28 and to
regenerate CoI. In addition, the alkenes can be controlled at Z-
type (Fs28) by Ir-mediated energy transfer process,[55] which can
undergo subsequent cyclization to form coumarin Gs28 in the
presence of acid. For terminal alkyne (R2 = H), the similar
oxidative cyclization intermediate Hs28 is expanded to a seven-
membered ring complex Is28 by insertion of another molecule of
alkyne. The Hs28 then undergoes a reductive elimination to afford
2-pyrone (Js28) and releasing As28.

10
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
undergoes ligand change with Es31 to generate palladium alkoxide
Gs31. β-Carbon migration of the ester group will yield the palladium
ester Hs31 and subsequent reductive elimination gives the desired
benzoate while releasing As31. This carbon migration tactic
renders the replacement of organometallic species by alcohol
possible and exhibits uniqueness since the fixation of CO 2 is
completed in the metal-free photocatalytic cycle, which is
mechanistically different from other reports. This protocol
represents a novel type of photocarboxylation where the
photocatalyst functions not only as a single electron mediator.

Accepted Manuscript
Scheme 29. Carboxylation of aryl halides by the combination of
palladium and photoredox catalysts.

Later in the same year, König and co-workers reported a


similar carboxylation by the combination of 4CzIPN, NiBr2·glyme,
neocuproine (L1), and HEH.[59] This protocol is applicable to
aliphatic bromides and aryl triflates with broad substrate scope.
K2CO3 is used as carbon source and a CO2 balloon is equipped
for increased yield. It is interesting that electron-deficient aryl Scheme 31. Stepwise photofixation of CO2 into aryl bromides in
bromides give better yield while electron-rich aryl triflates shows the presence of ketone.
higher yield. The reaction is initiated by ligand change between
NiBr2•glyme and L1 to produce LnNiII species (As30), which is then Compared with organic halides, the photocarboxylation of
reduced to LnNi0 species (Bs30) by 4CzIPN•–. Oxidative addition of other electrophiles is less explored, although transition-metal-
halides and following one-electron reduction afford an catalyzed the carboxylation of sulfonates and esters have been
intermediate Ds30. After CO2 insertion, further one step of SET reported.[61,62] To achieve this aim, Iwasawa's[63] and Jana's[64]
between nickel carboxylate Es30 and 4CzIPN•– regenerates Bs30. group discovered the palladium-catalyzed visible-light-driven
carboxylation of aryl or alkenyl triflates independently (Scheme
32). A number of alkenyl carboxylic acids are synthesized under
light-driven conditions and some synthetic applications are also
achieved with valuable products, such as estrone derivative,
adapalene, and bexarotene. Indeed, conditions for the two
protocols are analogous and the main discrepancy seems to be
the choice of ligand and additives. Iwasawa′s group observed that
the addition of tetrabutylammonium bromide (TBAB) or other
similar salts could improve the yield of aryl triflates. For alkenyl
triflates, electron-deficient PhXphos ligand is more suitable. The
reaction′s mechanism is similar to Martin and Iwasawa’s report.[58]
In this case, the more labile triflate group may be in its ionic form,
resulting in a cationic PdII intermediate Bs32. Coordination and
following migratory insertion of CO2 to Bs32 gives Cs32 which
lowers the reduction potential. The SET between Cs32 and
reductive IrII gives PdI carboxylate Ds32, which undergoes a
second SET to release the carboxylate and catalytic As32.

Scheme 30. Visible-light-nickel dual catalyzed carboxylation of


aryl and alkyl bromides and triflates.

Ishida and Murakami and co-workers developed a three-step


photocarboxylation of aryl bromides and CO2 in the presence of
ketone (Scheme 31).[60] Substrates bearing both electron-
donating and electron-withdrawing groups can be tolerated with
good to excellent yields. The author speculated that the reaction
can be divided into three independent steps. In step 1, vicinal diol
Bs31 is formed by the photoinduced dimerization of As31. After
deprotonation, the strained central C-C bond of Cs31 is cleaved
and the released electron pair is capable of capturing CO 2 to
produce carboxylate Ds31. Methylation will give the ester Es31 as
product in step 2. In the third step, oxidative addition complex Fs31

11
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW

Accepted Manuscript
Scheme 33. Carboxylation of gem-difluoroalkenes with CO2 by
light-driven palladium catalysis.

5.3 Photocarboxylation via light/transition-metal dual


catalysis from sp3 C–H bonds
Selective sp3 C-H bond dissociation with HAT catalyst
assisstance is reported in recent years.[66] Murakami and co-
workers discovered a UV light-mediated photoredox
carboxylation of benzylic C-H bonds with ketone as bifunctional
photo-/HAT catalyst and nickel (Scheme 34).[67] A number of
phenyl acetic acids are synthesized in good to moderate yields.
Scheme 32. Palladium-catalyzed visible-light-driven However, ethylbenzene gives relatively low yield (15%), which
carboxylation of aryl or alkenyl triflates. can be attributed to radical stabilizing effect and steric hinderance
of the additional methyl group. After developing the benzylic C-H
Fluoride is an indivisible part of the halide family according to bond carboxylation, they further developed C-H bond
its unique reactivity and biocompatibility. However, the high bond- carboxylation using alkanes. For example, cyclopentane is
dissociation energy and resistance to oxidative addition of C-F treated under a modified condition and the corresponding
bonds remains a problem in its utilization. cyclopentylcarboxylic acid is observed in 80% yield. When n-
Feng and co-workers reported an innovative carboxylation of pentane is employed, the carboxylatic acids are observed in 1:8:3
gem-difluoroalkenes with CO2 by photoredox/palladium dual ratio at 1-, 2-, and 3- position, respectively.
catalysis (Scheme 33).[65] Preliminary mechanistic studies In the proposed mechanism, ketone As34 as photocatalyst is
exclude the possibility of fluorovinyl carbanion and two-electron excited by the UV light and then the carbonyl oxygen is supposed
oxidative addition pathway, indicating the probability of radical to abstract a hydrogen from the fragile benzylic C-H bonds,
intermediates. In the proposed mechanism, the reaction is generating benzyl and ketyl radical (Bs34 and Cs34). Deprotonation
initiated by reductive quenching of the excited photocatalyst by of the ketyl radical Cs34 by tBuOK gives ketyl anion Ds34, which
iPr2NEt to afford IrII. SET between IrII and gem-difluoroalkene releases an electron and returns to As34. Simultaneously, the Ni II
generates monofluoroalkenyl radical As33. This moiety is complex gets two-electron to produce Ni0 complex Es34, which
substantially captured by Pd0 species (Bs33) to produce reactive reacts with Bs34 to afford NiI-carboxylate Gs34 after insertion of
PdI intermediate Cs33. Coordination of CO2 affords the complex CO2 in Fs34. The NiI-carboxylate Gs34 receives one-electron to
Ds33 and it is converted to palladium carboxylate Es33 by CO2 regenerate the nickel complex Es34. This method provides an
insertion. Afterwards, SET between IrII and Es33 regenerates the efficient way for benzylic C-H carboxylation, but the activation of
Pd0 while giving carboxylate Fs33. A facile interconversion may stable alkyl radical intermediate still requires extensive
occur at reversible radical capture process and E-isomer of Cs33 investigation.
is more kinetically favorable.

12
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW

Accepted Manuscript
Scheme 35. Remote sp3 C-H carboxylation by photoredox/Ni
catalysis.
Scheme 34. Carboxylation of benzylic C-H bonds by
nickel/ketone/UV light catalysis. Murakami and co-workers reported a light/ketone/Cu system
for the carboxylation of allylic sp3 C-H bonds (Scheme 36).[69]
Chain walking reaction is a powerful tool in the remote Substituted xanthone and NHC-carbene ligated cuprous chloride
activation of stable sp3 C-H sites by transferring the metal center. are selected as optimal cocatalyst. Ketone is initially excited by
A cooperative report from Crespi, König, and Martin discloses a light-irradiation. Carbonyl oxygen of the excited ketone Bs36
novel remote sp3 C-H carboxylation of alkyl bromides by the abstracts a hydrogen atom from the allylic C-H bond of
combination of photoredox and nickel catalysis (Scheme 35).[68] cyclohexene to generate geminal radical pair Cs36. After radical
Preliminary studies on homobenzylic bromide manifest that the coupling, the energetic homoallyl alcohol Ds36 is obtained as key
collaboration of 4CzIPN, NiBr2, neocuproine (L1), and HEH gives intermediate, which is deprotonated by the in-situ formed copper
the best yield with excellent linear/branched selectivity (66%, complex Es36 to afford copper alkoxide Fs36. The Fs36 undergoes
linear : branched = 10 : 90). More challenging remote alkyl β-carbon elimination to afford allylcopper Gs36 and As36. The
bromides carboxylation is implemented. The intriguing results formed C-Cu bond allows CO2 insertion and subsequent ligand
suggest that modification on the ancillary ligand has critical effect exchange gives the desired β,γ-unsaturated carboxylate Is36 and
on the reaction performance, wherein linear aliphatic acids are releases Es36. When the reaction is performed with the
generated in excellent yield by using a 2,2-bipyridine ligand (L5). corresponding homoallyl alcohols as starting material without light,
In this scenario, formal [1, n]-migration of Ni atom to the terminal the same acids are obtained, indicating the correctness of this
carbon provides the carboxylation of strongest primary sp3 C-H mechanism.
bonds, in which case HAT strategy is not applicable.
For the mechanism of remote C-H carboxylation, single
electron is produced incessantly by the photoredox cycle with
oxidized HEH radical cation as byproduct. Nickel bromide
undergoes photo-induced single electron reduction twice to
generate Ni0 complex As35. Subsequent oxidative addition with
alkyl bromide affords Bs35, which undergoes β-H elimination to
produce alkenyl-NiII complex Cs35. The reinsertion of hydrogen at
the terminal position is almost barrierless, leading to the
hydrometallation complex Ds35. SET from 4CzIPN•– to Ds35
generates NiI complex Es35 and the following CO2 coordination
and migratory insertion gives NiI carboxylate Fs35. It is further
reduced to liberate the linear carboxylate Gs35 and catalytic As35.
On the other hand, the homobenzyl bromide will produce a more
stable benzylic Ni complex Hs35, which undergoes similar SET
and CO2 insertion procedure, forming the branched product Is35.

Scheme 36. UV light-driven carboxylation of allylic C-H bonds.

13
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
6. Conclusion and outlook Peden, A. R. Portis, S. W. Ragsdale, T. B. Rauchfuss, J. N. H. Reek,
L. C. Seefeldt, R. K. Thauer and G. L. Waldrop, Chem. Rev. 2013,
In summary, this review collects the recent reports in the 113, 6621-6658; b) J. Artz, T. E. Müller, K. Thenert, J. Kleinekorte,
light-mediated carboxylation reactions with CO2. The employing R. Meys, A. Sternberg, A. Bardow, W. Leitner, Chem. Rev. 2018,
light as energy source allow the avoidance of stoichiometric 118, 434-504. c) T. Sakakura, J.-C. Choi, and H. Yasuda, Chem. Rev.
metallic reductants and hash conditations, thus providing a more 2007, 107, 2365-2387; d) M. Mikkelsen, M. Jørgensen and F. C.
sustainable way for synthesis of carboxylic acids . Taking a Krebs, Energy Environ. Sci. 2010, 3, 43-81; e) Q. Liu, L.-P. Wu, R.
deeper glance at mechanistic approach, strategies for generating Jackstell and M. Beller, Nat. Commun. 2015, 6, 5933-5947; f) Z.
high-energy intermediates can be generally categorized into Wang, Z. Zhao, Y. Li, Y. Zhong, Q. Zhang, Q. Liu, G. A. Solan, Y.
following aspects: (1) Light-driven direct process to access Ma and W.-H. Sun, Chem. Sci. 2020, 11, 6766-6774; g) B. Zhang,
energetic species, including radical anions, carbenes, 1,3-dienes, G. Du, W. Hang, S. Wang and C. Xi, Eur. J. Org. Chem. 2018, 1739-
ylides, and organometallic reagents, which can react with CO2; (2) 1743; h) L. Zhang, Z. Han, X. Zhao, Z. Wang and K. Ding, Angew.
In-situ generated carbanions based on the transformation of Chem. Int. Ed. 2015, 54, 6186-6189; i) W.-H. Wang, Y. Himeda, J.
substrates by photoredox catalysis; (3) photoredox/transition- T. Muckerman, G. F. Manbeck and E. Fujita, Chem. Rev. 2015, 115,

Accepted Manuscript
metal dual catalysis to generate reactive carbon-metal bond 12936-12973; j) J. Qiao, Y. Liu, F. Hong and J. Zhang, Chem. Soc.
towards CO2 insertion. In some cases, no external catalyst is Rev. 2014,43, 631-675; k) Y. Ma, X. Wang, Y. Jia, X. Chen, H. Han
required, so it is reasonable to trace back to the original and C. Li, Chem. Rev. 2014, 114, 9987-10043; l) X.-B. Lv and D. J.
photochemistry for exploring the intrinsic light-energy-promoted Darensbourg, Chem. Soc. Rev. 2012, 41,1462-1484; m) B. Yu and
reactions, which might afford inspiration to organic catalysis L.-N. He, ChemSusChem 2015, 8, 52-62.
endeavor. [3] R. Gniadecki, C. Assaf, M. Bagot, R. Dummer, M. Duvic, R. Knobler,
Despite the promising results, the area still needs much more A. Ranki, P. Schwandt and S. Whittaker, Br. J. Dermatol. 2007, 157,
investigation due to some limitations in the current reports. For 433-440.
instance, there still lacks the reports of enantioselective [4] H. Maag, in Prodrugs: Challenges and Rewards Part 1, (Eds.: V. J.
photocarboxylation, which would be a practical way for providing Stella, R. T. Borchardt, M. J. Hageman, R. Oliyai, H. Maag and J.
pharmaceutical and bioactive scaffolds. [70] As for the direct W. Tilley), Springer, New York, NY, 2007, pp. 703-729.
activation of CO2, no catalytic example involving SET from the [5] For selected examples, see: a) K. Huang, C.-L. Sun and Z.-J. Shi,
excited state PC to CO2 is reported. Moreover, the regioselective Chem. Soc. Rev. 2011, 40, 2435-2452; b) M. Cokoja, C. Bruckmeier,
carboxylation of unsaturated compounds is still beyond study.[71] B. Rieger, W. A. Herrmann, and F. E. Kuhn, Angew. Chem. Int. Ed.
Furthermore, it comes to more interesting for the direct 2011, 50, 8510-8537; c) Y. Tsuji and T. Fujihara, Chem. Commun.
carboxylation of general sp3 C-H bond other than limit to allylic 2012, 48, 9956-9964; d) S.Wang and C. Xi, Chem. Soc. Rev. 2019,
and benzylic position. To achieve these, rational design of novel 48, 382-404; e) L. Zhang and Z.-M. Hou, Chem. Sci. 2013, 4, 3395-
photocatalysts with stronger reducing ability, development of 3403; f) B. Yu, Z-F. Diao, C.-X. Guo, L.-N. He, J. CO2 Util. 2013,
chiral ligands, employment of co-catalysts are probable solutions. 1, 60-68; g) R. Martin and A. W. Kleij, ChemSusChem 2011, 4,
In some preliminary attempts, semiconductors,[72] enzymes,[73] 1259–1263; h) A. Tortajada, F. Juliá-Hernández, M. Börjesson, T.
and electrochemical method[74] are incorporated to realize Moragas and R. Martin, Angew. Chem. Int. Ed. 2018, 57, 15948-
carboxylic acids. We also hope to witness the involvement of 15982; i) J.-T. Hong, M. Li, J.-N. Zhang, B.-Q. Sun, and F.-Y. Mo,
more earth-abundant metals or other organocatalysts in this ChemSusChem 2019, 12, 6-39; j) S.-S. Yan, Q. Fu, L.-L. Liao, G.-
territory for mechanistic innovation. Q. Sun, J.-H. Ye, L. Gong, Y.-Z. Bo-Xue, D.-G. Yu, Coord. Chem.
Rev. 2018, 374, 439-463; k) W. Hang, Y. Yi and C. Xi, Adv. Synth.
Catal. 2020, 362, 2337-2341; l) W. Xiong, F. Shi, R. Cheng, B. Zhu,
L. Wang, W. Wu, C. Qi, M. Lei and H. Jiang, ACS Catal. 2020, 10,
Acknowledgements 7968-7978; m) L. Fu, S. Li, Z. Cai, Y. Ding, X.-Q. Guo, L.-P. Zhou,
D. Yuan, Q.-F. Sun and G. Li, Nat. Catal. 2018, 1, 469-478; n) T.
This work was supported by the National Natural Science Cao, Z. Yang and S. Ma, ACS Catal. 2017, 7, 4504-4508; o) S. Li,
Foundation of China (21871163, 91645120, 21472106, and W. Yuan and S. Ma, Angew. Chem. Int. Ed. 2011, 50, 2578-2582; p)
21911530097). T. G. Ostapowicz, M. Schmitz, M. Krystof, J. Klankermayer, W.
Leitner, Angew. Chem. Int. Ed. 2013, 52, 12119.
[6] a) Y.-Y. Gui, W.-J. Zhou, J.-H. Ye and D.-G. Yu, ChemSusChem
Keywords: carboxylation • carboxylic acids • photocatalysis •
2017, 10, 1337-1340; b) C. S. Yeung, Angew. Chem. Int. Ed. 2019,
radical reaction • carbon dioxide 58, 5492-5502; c) J. Hou, J.-S. Li and J. Wu, Asian J. Org. Chem.
2018, 7, 1439-1447; d) F. Tan, and G. Yin, Chin. J. Chem. 2018, 36,
Conflict of interest 545-554; f) Y. Cao, X. He, N. Wang, H.-R. Li, and L.-N. He, Chin.
J. Chem. 2018, 36, 644-659.
[7] a) B. Kumar, M. Llorente, J. Froehlich, T. Dang, A. Sathrum, and C.
The authors declare no conflict of interest. P. Kubiak, Annu. Rev. Phys. Chem. 2012, 63 ,41-69; b) N. Sutin, C.
Creutz and E. Fujita, Comments. Inorg. Chem. 1997, 19, 67-92.
[8] a) C. K. Prier, D. A. Rankic and D. W. C. MacMillan, Chem. Rev.
2013, 113, 5322-5363; b) K. L. Skubi, T. R. Blum, and T. P. Yoon,
References Chem. Rev. 2016, 116, 10035-10074; c) N. A. Romero and D. A.
Nicewicz, Chem. Rev. 2016, 116, 10075-10166; d) M. H. Shaw, J.
[1] a) M. Aresta, A. Dibenedetto and A. Angelini, Chem. Rev. 2014, 114, Twilton, and D. W. C. MacMillan, J. Org. Chem. 2016, 81, 6898-
1709-1742; M. Aresta and A. Dibenedetto, Dalton Trans. 2007, 6926; e) D. Ravelli, S. Protti, M. Fagnoni, Chem. Rev. 2016, 116,
2975-2992; c) N. Assen, L. J. Müller, A. Steingrube, P. Voll, and A. 9850-9913; f) Y.-Y. Chen, L.-Q. Lu, D.-G. Yu, C.-J. Zhu and W.-J.
Bardow, Environ. Sci. Technol. 2016, 50, 1093-1101; d) M. Aresta, Xiao, Sci. China Chem. 2019, 62, 24-57.
Carbon Dioxide as Chemical Feedstock, Wiley-VCH, Weinheim, [9] S. Matsuoka, T. Kohzuki, C. Pac, A. Ishida, S. Takamuku, M.
2010; e) A. Sternberg, C. M. Jens and A. Bardow, Green Chem. Kusaba, N. Nakashima and S. Yanagida, J. Phys. Chem. 1992, 96,
2017, 19, 2244-2259 4437-4442.
[2] For selected examples, see: a) A. M. Appel, J. E. Bercaw, A. B. [10] H. Seo, A. Liu and T. F. Jamison, J. Am. Chem. Soc. 2017, 139,
Bocarsly, H. Dobbek, D. L. DuBois, M. Dupuis, J. G. Ferry, E. 13969-13972.
Fujita, R. Hille, P. J. A. Kenis, C. A. Kerfeld, R. H. Morris, C. H. F. [11] T. Koike and M. Akita, Org. Chem. Front. 2016, 3, 1345-1349.

14
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
[12] X.-W. Lan, N.-X. Wang and Y. Xing, Eur. J. Org. Chem. 2017, 2449.
5821-5851. [48] D. D. M. Wayner and D. Griller, J. Am. Chem. Soc. 1985, 107,
[13] J.-H. Ye, M. Miao, H. Huang, S.-S. Yan, Z.-B. Yin, W.-J. Zhou and 7764-7765.
D.-G. Yu, Angew. Chem. Int. Ed. 2017, 56, 15416-15420; Angew. [49] U. M. Dzhemilev, A. G. Ibragimov in Modern Reduction Methods,
Chem. 2017, 129, 15618-15622. (Eds.: V. P. G. Andersson and I. J. Munslow), Wiley-VCH,
[14] R. Mello, J. C. Arango-Daza, T. Varea and M. E. González-Núñez, Weinheim, 2008, pp. 447-489.
J. Org. Chem. 2018, 83, 13381-13394. [50] J. Chen, J. Guo and Z. Lu, Chin. J. Chem. 2018, 36, 1075-1109.
[15] A. Nomoto, Y. Kojo, G. Shiino, Y. Tomisaka, I. Mitani, M. Tatsumi [51] K. Murata, N. Numasawa, K. Shimomaki, J. Takaya and N.
and A. Ogawa, Tetrahedron Lett. 2010, 51, 6580-6583. Iwasawa, Chem. Commun. 2017, 53, 3098-3101.
[16] H. Seo, M. H. Katcher and T. F. Jamison, Nat. Chem. 2017, 9, 453- [52] K. Murata, N. Numasawa, K. Shimomaki, J. Takaya and N.
456. Iwasawa, Front. Chem. 2019, 7.
[17] D. A. Morgenstern, R. E. Wittrig, P. E. Fanwick and C. P. Kubiak, [53] Q.-Y. Meng, S. Wang, G. S. Huff and B. König, J. Am. Chem. Soc.
J. Am. Chem. Soc. 1993, 115, 6470-6471. 2018, 140, 3198-3201.
[18] S. Tazuke and H. Ozawa, J. Chem. Soc., Chem. Commun. 1975, [54] J. Hou, A. Ee, W. Feng, J.-H. Xu, Y. Zhao and J. Wu, J. Am. Chem.
237-238. Soc. 2018, 140, 5257-5263.

Accepted Manuscript
[19] S. Tazuke, S. Kazama and N. Kitamura, J. Org. Chem. 1986, 51, [55] K. Singh, S. J. Staig and J. D. Weaver, J. Am. Chem. Soc. 2014, 136,
4548-4553. 5275-5278.
[20] A. V. Nikolaitchik, M. A. J. Rodgers and D. C. Neckers, J. Org. [56] A. Correa and R. Martín, J. Am. Chem. Soc. 2009, 131, 15974-
Chem. 1996, 61, 1065-1072. 15975.
[21] M. Minabe, K. Isozumi, K. Kawai and M. Yoshida, Bull. Chem. Soc. [57] M. Börjesson, T. Moragas, D. Gallego and R. Martin, ACS Cat.
Jpn. 1988, 61, 2063-2066. 2016, 6, 6739-6749.
[22] J. E. Chateauneuf, J. Zhang, J. Foote, J. Brink and M. W. Perkovic, [58] K. Shimomaki, K. Murata, R. Martin and N. Iwasawa, J. Am. Chem.
Adv. Environ. Res. 2002, 6, 487-493. Soc. 2017, 139, 9467-9470.
[23] H. Tagaya, M. Onuki, Y. Tomioka, Y. Wada, M. Karasu and K. [59] Q.-Y. Meng, S. Wang and B. König, Angew. Chem. Int. Ed. 2017,
Chiba, Bull. Chem. Soc. Jpn. 1990, 63, 3233-3237. 56, 13426-13430; Angew. Chem. 2017, 129, 13611-13615.
[24] Y. Ito, Y. Uozu and T. Matsuura, J. Chem. Soc., Chem. Commun. [60] N. Ishida, Y. Masuda, W. Liao and M. Murakami, Chem. Lett. 2019,
1988, 562-564. 48, 1316-1318.
[25] Y. Ito, Tetrahedron 2007, 63, 3108-3114. [61] K. Nogi, T. Fujihara, J. Terao and Y. Tsuji, J. Org. Chem. 2015, 80,
[26] T. Ju, Q. Fu, J.-H. Ye, Z. Zhang, L.-L. Liao, S.-S. Yan, X.-Y. Tian, 11618-11623.
S.-P. Luo, J. Li and D.-G. Yu, Angew. Chem. Int. Ed. 2018, 57, [62] A. Correa, T. León and R. Martin, J. Am. Chem. Soc. 2014, 136,
13897-13901; Angew. Chem. 2018, 130, 14903-14907 1062-1069.
[27] X. Fan, X. Gong, M. Ma, R. Wang and P. J. Walsh, Nat. Commun. [63] K. Shimomaki, T. Nakajima, J. Caner, N. Toriumi and N. Iwasawa,
2018, 9, 4936-4943. Org. Lett. 2019, 21, 4486-4489.
[28] L.-L. Liao, G.-M. Cao, J.-H. Ye, G.-Q. Sun, W.-J. Zhou, Y.-Y. Gui, [64] S. K. Bhunia, P. Das, S. Nandi and R. Jana, Org. Lett. 2019, 21,
S.-S. Yan, G. Shen and D.-G. Yu, J. Am. Chem. Soc. 2018, 140, 4632-4637.
17338-17342. [65] C. Zhu, Y.-F. Zhang, Z.-Y. Liu, L. Zhou, H. Liu and C. Feng, Chem.
[29] V. R. Yatham, Y. Shen and R. Martin, Angew. Chem. Int. Ed. 2017, Sci. 2019, 10, 6721-6726.
56, 10915-10919; Angew. Chem. 2017, 129, 11055-11059. [66] a) M. H. Shaw, V. W. Shurtleff, J. A. Terrett, J. D. Cuthbertson, D.
[30] H. Wang, Y. Gao, C. Zhou and G. Li, J. Am. Chem. Soc. 2020, 142, W. C. MacMillan, Science, 2016, 352, 1304-1308; b) C. Le, Y.-F.
8122-8129. Liang, R. W. Evans, X.-M. Li, D. W. C. MacMillan, Nature 2017,
[31] W.-J. Zhou, Z.-H. Wang, L.-L. Liao,Y.-X. Jiang, K.-G. Cao, T. Ju, 547, 79-83.
Y. Li, G.-M. Cao and D.-G. Yu, Nat. Commun. 2020, 11, 3263-3271. [67] N. Ishida, Y. Masuda, Y. Imamura, K. Yamazaki and M. Murakami,
[32] J. Hou, A. Ee, H. Cao, H.-W. Ong, J.-H. Xu and J. Wu, Angew. J. Am. Chem. Soc. 2019, 141, 19611-19615.
Chem. Int. Ed. 2018, 57, 17220-17224. [68] B. Sahoo, P. Bellotti, F. Juliá-Hernández, Q.-Y. Meng, S. Crespi, B.
[33] B. Zhang, Y. Yi, Z.-Q. Wu, C. Chen and C. Xi, Green Chem. 2020, König and R. Martin, Chem. Eur. J. 2019, 25, 9001-9005.
DOI: 10.1039/D0GC02254C. [69] N. Ishida, Y. Masuda, S. Uemoto and M. Murakami, Chem. Eur. J.
[34] Q. Fu, Z.-Y. Bo, J.-H. Ye, T. Ju, H. Huang, L.-L. Liao and D.-G. Yu, 2016, 22, 6524-6527.
Nat. Commun. 2019, 10, 3592-3600. [70] M. Takimoto, Y. Nakamura, K. Kimura and M. Mori, J. Am. Chem.
[35] L. Song, D.-M. Fu, L. Chen, Y.-X. Jiang, J.-H. Ye, L. Zhu, Y. Lan, Soc. 2004, 126, 5956-5957.
Q. Fu and D.-G. Yu, Angew. Chem. Int. Ed. 2020, DOI: [71] A. Tortajada, R. Ninokata and R. Martin, J. Am. Chem. Soc. 2018,
10.1002/anie.202008630. 140, 2050-2053.
[36] Q.-Y. Meng, T. E. Schirmer, A. L. Berger, K. Donabauer and B. [72] a) B. R. Eggins, J. T. S. Irvine, E. P. Murphy and J. Grimshaw, J.
König, J. Am. Chem. Soc. 2019, 141, 11393-11397. Chem. Soc., Chem. Commun. 1988, 1123-1124; b) H. Fujiwara, M.
[37] R. Wheland and P. D. Bartlett, J. Am. Chem. Soc. 1970, 92, 6057- Kanemoto, H. Ankyu, K. Murakoshi, Y. Wada and S. Yanagida, J.
6058. Chem. Soc., Perk. Trans. 2, 1997, 317-322.
[38] W. W. Sander, J. Org. Chem. 1989, 54, 4265-4267. [73] Y. Amao, Sustainable Energy & Fuels 2018, 2, 1928-1950.
[39] H. Giezendanner, M. Märky, B. Jackson, H. J. Hansen and H. [74] a) T. Sigeru, T. Hideo, H. Takeshi, M. Kazuo, J. Anny, F. Peluger
Schmid, Helv. Chim. Acta 1972, 55, 745-748. and J.-F. Fauvarque, Chem. Lett. 1986, 15, 169-172. b) K.-J. Jiao,
[40] A. Padwa and S. I. Wetmore, J. Am. Chem. Soc. 1974, 96, 2414- Z.-M. Li, X.-T. Xu, L.-P. Zhang, Y.-Q. Li, K. Zhang and T.-S. Mei,
2421. Org. Chem. Front. 2018, 5, 2244-2248.
[41] Y. Masuda, N. Ishida and M. Murakami, J. Am. Chem. Soc. 2015,
137, 14063-14066.
[42] N. C. Yang and C. Rivas, J. Am. Chem. Soc. 1961, 83, 2213-2213.
[43] S. Inoue and N. Takeda, Bull. Chem. Soc. Jpn. 1977, 50, 984-986.
[44] Y. Hirai, T. Aida and S. Inoue, J. Am. Chem. Soc. 1989, 111, 3062-
3063.
[45] M. Komatsu, T. Aida and S. Inoue, J. Am. Chem. Soc. 1991, 113,
8492-8498.
[46] P. Cocolios, R. Guilard, D. Bayeul and C. Lecomte, Inorg. Chem.
1985, 24, 2058-2062.
[47] R. F. Johnston and J. C. Cooper, Organometallics 1987, 6, 2448-

15
This article is protected by copyright. All rights reserved.
ChemSusChem 10.1002/cssc.202001974

MINIREVIEW
Entry for the Table of Contents

Accepted Manuscript
Photocarboxylation: Fixation of CO2 into organic molecules with C-C bond formation is a challenging project. With UV/visible light
assistance, alkenes, alkynes, organometallic reagents, organic halides, and sp3 C-H bonds are successfully transformed to the
corresponding carboxylic acids. These protocols exhibit divergent chemo- and regioselectivity with tolerating broad functional group.

16
This article is protected by copyright. All rights reserved.

You might also like