You are on page 1of 84

View Article Online

Organic &
View Journal

Biomolecular
Chemistry
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: B. Hong, Org.
Biomol. Chem., 2020, DOI: 10.1039/D0OB00759E.
Volume 15
Number 47
This is an Accepted Manuscript, which has been through the
Organic &
21 December 2017
Pages 9945-10124
Royal Society of Chemistry peer review process and has been
Biomolecular accepted for publication.
Chemistry
rsc.li/obc
Accepted Manuscripts are published online shortly after acceptance,
before technical editing, formatting and proof reading. Using this free
service, authors can make their results available to the community, in
citable form, before we publish the edited article. We will replace this
Accepted Manuscript with the edited and formatted Advance Article as
soon as it is available.

You can find more information about Accepted Manuscripts in the


Information for Authors.

Please note that technical editing may introduce minor changes to the
text and/or graphics, which may alter content. The journal’s standard
ISSN 1477-0520 Terms & Conditions and the Ethical guidelines still apply. In no event
PAPER
I . J. Dmochowski et al.
Oligonucleotide modifi cations enhance probe stability for
shall the Royal Society of Chemistry be held responsible for any errors
single cell transcriptome in vivo analysis (TIVA)

or omissions in this Accepted Manuscript or any consequences arising


from the use of any information it contains.

rsc.li/obc
Page 1 of 83 Organic & Biomolecular Chemistry

View Article Online

Enantioselective Synthesis Enabled by Visible Light Photocatalysis


DOI: 10.1039/D0OB00759E

Organic & Biomolecular Chemistry Accepted Manuscript


Bor-Cherng Hong

Department of Chemistry and Biochemistry, National Chung Cheng University, Chia-Yi, 621,
Taiwan, R.O.C.

chebch@ccu.edu.tw
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Abstract:
Enantioselective photoreaction has been a synthetic challenge for decades. With the continuous
development of modern visible light photocatalysis and asymmetric catalysis, remarkable advances
have been achieved through the synergistic action of these catalytic reactions, allowing the
construction of various enantiomerically enriched molecules that were once inaccessible using
photocatalytic reactions. This review presents some of the contemporary developments in
enantioselective visible-light photocatalysis reactions, covering the period from 2008 to March 2020,
with the contents classified by catalysis type.

Contents:
1. Introduction
2. Chiral enamine/iminium catalysis with/without photocatalyst
3. Chiral Lewis acid catalysis with photocatalyst
4. Chiral Lewis base catalysis with photocatalyst
5. Phase transfer catalysis with photocatalyst
6. Chiral Brønsted acid catalysis with photocatalyst
7. Chiral photocatalysts
8. Chiral anion binding catalysis with photocatalyst
9. Metal catalysis
10. N-Heterocyclic carbene catalysis with photocatalyst
11. Enzyme catalysis with/without photocatalyst
12. Miscellaneous
13. Conclusions

1. Introduction
Visible light-catalyzed reactions have received much attention in modern chemistry since the
turn of the 21st century. Many visible light photocatalysis techniques are convenient and efficient
synthetic methods for selectively creating chemical bonds and producing multiple stereocenters.
Organic & Biomolecular Chemistry Page 2 of 83

More than a thousand publications have documented photocatalytic organic reactions driven byOnline
View Article
DOI: 10.1039/D0OB00759E
visible light, some of which have been summarized in various review papers. Among these
photocatalytic reactions, over 200 papers have described enantioselective reactions, showing an

Organic & Biomolecular Chemistry Accepted Manuscript


increasing trend in recent years. In particular, the recent development of electron donor–acceptor
complexes for enantioselective photocatalyst-free reactions has further expanded this field.1
However, fewer reviews have covered enantioselective examples of visible light photocatalysis, with
most concentrating on particular topics and niche areas,2 while several mini-reviews on
enantioselective visible light photocatalysis have focused on specific topics, including chiral
phosphoric acids,3 chiral Lewis acid catalysis,4 enantioselective [2+2] cycloadditions,5 chiral metal
photocatalysts comprising octahedral RuII and IrIII complexes,6 the activation of α,β-unsaturated
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

carbonyl compounds,7 enantioselective catalysis,8 iminium and enamine catalysis,9


biscyclometalated rhodium complexes,10 deracemization,11 and enzymes in biocatalysis.12
The vigorous development of visible light photocatalytic technology will undoubtedly continue
for decades. Therefore, to obtain a practical perspective on enantioselective visible light
photocatalysis, an update and broad overview of different aspects of this research area is needed to
establish the more effective synthetic methods. Accordingly, this review article summarizes some of
the progress on this theme between 2008 and March 2020. The content is categorized according to
the type of catalysis with or without photocatalyst, including chiral enamines/imine catalysis, chiral
Lewis acid catalysis, chiral Lewis base catalysis, phase-transfer catalysis, chiral Bronsted acid
catalysis, chiral anion binding catalysis, metal catalysis, N-heterocyclic carbene catalysis, enzyme
catalysis, and chiral photocatalysis, and a few other uncommon reaction types.

2. Chiral enamine/iminium catalysis with/without photocatalysts


In 2008, Macmillan and coworkers reported the first example of combining photoredox catalyst
Ru(bpy)3Cl2 with an imidazolidinone organocatalyst in the direct enantioselective α-alkylation of
aliphatic aldehydes with racemic α-bromocarbonyl compounds (Scheme 1).13 Subsequently, this
synthetic method was applied to the enantioselective α-trifluoromethylation of aldehydes (Scheme
1).14 More recently, this strategy was applied to the direct enantioselective α-amination of aldehydes
using a photonic excitable amine reagent (Scheme 1).15 Under light irradiation, the amine reagent
extrudes the photolabile dinitrophenylsulfonyloxy (ODNs) leaving group and provides the
electrophilic N-centered radical, which then couples with the enamine (derived from condensation of
the imidazolidinone catalyst and starting aldehyde) to obtain the α-amination product.
Page 3 of 83 Organic & Biomolecular Chemistry

2008 MacMillan O View Article Online


N R2
DOI: 10.1039/D0OB00759E
• HOTf O
t
N Bu 4
R2 H R3

Organic & Biomolecular Chemistry Accepted Manuscript


O H (20 mmol %)
R1 + R1 3
H Br R3 Ru(bpy)3Cl2 (0.5 mmol %)
2,6-lutidine, DMF, rt, 15-W CFL 63–93% yield
1 2
up to 99% ee
R1 = nhex, EtCH=CH(CH2)3, Bn, chex, 4-piperidinyl,1-adamantyl,
R2 = CO2Et, H; R3 = CO2Et, Bz, CO2CH2CF3

2009 MacMillan
O
N
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

• TFA O F F
t
N Bu 4
O H (20 mmol %) H R2
F F
R1 + R1 7
H I R2 Ir(ppy)2(dtb-bpy)•PF6 (0.5 mol %)
61–86% yield
5 6 2,6-lutidine, DMF, –20 °C
up to 99% ee
26-W CFL

R1 = nhex, (CH2)3CO2Et, (CH2)3OBn, (CH2)3NPhth, Bn, MeOC6H4, MeCHPh, 1-adamantyl


R2 = CF3, C6F5, Br, CO2Et, F

2013 MacMillan O
N R3
R3 O O • HOTf O
NO2 Et
O N
N S N (30 mmol %) H R2
R1 + R2 O 11 H
H R1
2,6-lutidine, DMSO–CH3CN, 15 °C 10
8 9 NO2
26-W CFL 67–79% yield
up to 94% ee
R1 = nhex, iPr, Bn, (CH2)3CO2Et, (CH2)3OBn, (CH2)3NPhth, 4-MeOC6H4, chexCH2, allyl
R2 = CO2Me, Cbz, Boc, Alloc, Fmoc; R3 = Me, nBu, Ph(CH2)3, MOM
Scheme 1 Enantioselective -alkylation or -amination of aldehydes.

In 2010, Macmillan et al. developed the enantioselective -benzylation of aliphatic aldehydes


12 with benzyl bromides 13 by the synergistic action of photoredox catalysis and organocatalysis.
Combining the Macmillan imidazolidinone organocatalyst and photoredox catalyst fac-Ir(ppy)3 under
household compact fluorescent light irradiation provided -alkylation products with a homobenzylic
stereocenter in good yields (68%–93%) and good enantioselectivities (82%–97% ee) (Scheme 2).16
In their study, benzylic bromide 13 reacted with fac-Ir(ppy)3* via single-electron transfer (SET) to
generate benzylic radical 15. Concomitantly, exposure of aldehyde 1 to organocatalyst 4 provided
chiral enamine 16, which then reacted with radical 15 through Si-face addition to afford radical 17.
Subsequent oxidation of 17 by fac-Ir(ppy)3+ followed by hydrolysis to give -alkylation product 14
completed the photoredox catalysis and organocatalysis cycles.
Organic & Biomolecular Chemistry Page 4 of 83

View Article Online


DOI: 10.1039/D0OB00759E
O N
O
HOTf
N

Organic & Biomolecular Chemistry Accepted Manuscript


O Bn 4 (20 mole %) H Ar
H
R + Ar Br
H R
fac-IrIII(ppy)3 (0.5 mol %), CFL 14
12 13 2,6-lutidine (2 equiv), DMSO, rt
68-93% yield
82-97% ee
R = n-hexyl, Bn, BnO(CH2)3, cyclohexyl, PhthNCH2
Ar = benzyl, 2,4-(NO2)2C6H3, 4-pyridinyl, quinolinyl, pyrazinyl, triazinyl, 1H-benzo[d]imidazolyl

Me
15 O
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ar Br reduction Ar Me N O
H R
13 + e- CH2 H
N 12

16
R
O
fac-IrIII(ppy)3* fac-IrIV(ppy)3+ N
O N
photoredox Bn N organocatalysis
catalysis
Bn N
Ar H 4
CFL e- 17
R O
oxidation N
fac-IrIII(ppy)3 Bn N O

Ar Ar H
18
R R 14

Scheme 2 Enantioselective α-benzylation of aldehydes.

Later, Zeitler and coworkers reported an enantioselective -alkylation of aliphatic aldehydes


with diethyl 2-bromomalonate, 2-bromo-1-(4-nitrophenyl)ethan-1-one, or 1-iodononafluorobutane
via enamine–iminium organocatalysis and organophotocatalysis with eosin Y and green-light
irradiation (Scheme 3).17 This process provides a new approach to the -alkylation of aliphatic
aldehydes beyond the scope of benzylic halides. The mechanism was proposed, as depicted in Scheme
3. After irradiating with green light, excited photoredox catalyst 1eosin Y* transitioned to its more
stable triplet state (3eosin Y*) through intersystem crossing (ISC) process. The authors assumed that
eosin Y served as a reducing agent, which depended on sacrificial oxidation of a catalytic amount of
the enamine as the initial electron supply, to afford alkyl radical 22 through single electron transfer
with alkyl halide 20. Subsequent addition of alkyl radical 22 to enamine 23, obtained from the
organocatalysis system, provided radical 24. Oxidation of radical 24 by reductive quenching of 3eosin
Y* gave iminium 25, which was hydrolyzed to afford product 21, and generated catalyst 4 to complete
the organocatalytic cycle. Subsequently, Zeitler et al. applied microflow technology to synergistic
organocatalysis and visible-light photoredox catalysis, which accelerated the reaction rate of
enantioselective alkylation of fatty aldehydes by more than one order of magnitude.18 For example,
one of the reactions required 18 h to reach completion under traditional batch irradiation, but only 45
Page 5 of 83 Organic & Biomolecular Chemistry

min under microreactor irradiation. Further extensive study on related methods has been reported byOnline
View Article
DOI: 10.1039/D0OB00759E
the same group.19 More recently, Melchiorre extended the enantioselective alkylation of aldehydes
to alkyl chloride using a dithiocarbamate anion catalysis strategy (Scheme 4).20 Alkyl chloride is

Organic & Biomolecular Chemistry Accepted Manuscript


activated by an SN2 pathway using the chromophoric unit of nucleophilic dithiocarbamate. The
resulting photon-absorbing intermediate provides alkyl radicals through visible-light-induced
homolytic cleavage, and undergoes a C–C bond-forming reaction using the asymmetric amino
catalyst. Furthermore, recyclable visible-light photocatalysis of the enantioselective -alkylation of
aldehydes with alkyl bromides has been reported by Li, Zhang, and coworkers by merging the
Macmillan catalyst and porous Ru(bpy)32+-linked polymers.21 Additionally, the enantioselective -
alkylation of aldehydes with alkyl bromides using covalent organic frameworks (COFs) as
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

heterogeneous photocatalysts, the Macmillan imidazolidinone organocatalyst, and 2,6-lutidine in


DMF under LED irradiation at 440 nm has been reported by Cui.22

O N
t HOTf
Bu O
N (20 mol %) R2
H 4
O H
+ R2 X Br Br lutidine (2 equiv)
H R1
HO O O DMF, rt 21
20
19 R1 green LED 56-85% yield
( = 530 nm) 86-96% ee
R1 = n-hexyl, PhCH2 Br Br
X = Br, I HO2C
eosin Y
(2.5 mol %)
examples of halides:
NO2
CO2Et

Br CO2Et Br I(CF2)3CF3
O

O R2 R2 X
N O
22 20
t
Bu reduction
N N
23 N
O H H t
24 Bu
eosin Y eosin Y
H R1 R2
R1
R1 O
19
organocatalysis oxidation photoredox Green
N catalysis light
HN 4 O
t
Bu 3 1
O N eosin Y* eosin Y*
R2 N
H
H t
Bu
R1 21 ISC
R1 R2 25

Scheme 3 Enantioselective -akylation of aldehydes.


Organic & Biomolecular Chemistry Page 6 of 83

View Article Online


29 (20 mol %)
OHC R2 DOI: 10.1039/D0OB00759E
30 (20 mol %)
OHC R1 + R2 Cl
2,6-lutidine (1.2 equiv) R1

Organic & Biomolecular Chemistry Accepted Manuscript


26 27 CH3CN, 25 °C
blue LEDs, 24 h 29
R1 = nBu, Bn 57-77% yield
R2 = CH2C(O)Ph, CH2CN, CH(CO2Et)2 74-90% ee

Ar Br
N Ar
H OTMS N
29 30 SK
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ar = 3,5-(CF3)2C6H3 S

Scheme 4 Enantioselective -alkylation of aldehydes with alkyl chlorides.

In contrast, an enantioselective domino Michael addition/oxyamination of ,-unsaturated


aldehydes through iminium catalysis and photoinduced singly occupied molecular orbital (photo-
SOMO) catalysis was reported by Jang and coworkers (Scheme 5).23 The reaction introduced an
efficient photosensitizer, di-tetrabutylammonium cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′
dicarboxylato)ruthenium(II) (N719 dye), and the Jørgensen–Hayashi catalyst to enals, affording -
alkyl--oxyamination products in good yields (53%–80%) and good enantioselectivities (90%–99%
ee).
Page 7 of 83 Organic & Biomolecular Chemistry

View Article Online


DOI: 10.1039/D0OB00759E
Ph CO2TBA
Ph EtO2C CO2Et
N
34

Organic & Biomolecular Chemistry Accepted Manuscript


H OTMS OHC HO2C
EtO2C (20 mol %) R N
OHC
R + O N NCS
EtO2C N719 (0.04 mol %), TiO2 N Ru
31 adamantanecarboxylic acid N NCS
32 (30 mol %)
(3 equiv) TEMPO (2 equiv), CH3CN, rt N
33 HO2C
8 W cool whilte fluorescent
53-80%yield
R = Ph, C6H4OMe, C6H4Me, C6H4NO2, C6H4Cl, 90-99% ee N719
C6H4Br, 2-naphthyl, 2-thiophenyl CO2TBA

TBA = tetrabutylammonium
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

CFL
EtO2C CO2Et

N719 32
Ru(II) Ru(II)*
photoredox N Ph
catalysis
Ph
e- N
OTMS
- e- EtO C
2 H 2O
TEMPO Ru(I) R
or O2 CO2Et 36 35
SOMO R
OHC
R
= CPh2OTMS R Ph 31
organocatalysis
37 Ph
N
N CO2Et H EtO2C CO2Et
OTMS
EtO2C
OHC
R
O N
N O
TEMPO 33 N

38
EtO2C H 2O
TEMPO
R
CO2Et

Scheme 5 Domino Michael addition/oxyamination

In 2013, Melchiorre et al. reported applying photoactivated chiral electron donor–acceptor


(EDA) complexes to the alkylation of aldehydes with alkyl bromides (Scheme 6).24 This chiral
catalyst does not contain any photosensitive groups, but instead directs photochemical activation of
substances by evoking the transient formation of chiral EDA complexes to establish a stereoselective
process.
Organic & Biomolecular Chemistry Page 8 of 83

View Article Online

42 (20 mol %) O H DOI: 10.1039/D0OB00759E


O O
2,6-lutidine (1 equiv) O Ar
R1 Br H
H + R2 MeOtBu, rt, 23 W CFL R1 R2 N Ar

Organic & Biomolecular Chemistry Accepted Manuscript


39 40 41 H H
OTMS
70-96% yield 42
up to 94% ee Ar = 3,5-(CF3)2C6H3
O NO2
O NO2 42 (20 mol %)
2,6-lutidine (1 equiv) H
R1 Br
H
+ t
MeO Bu, rt, 23 W CFL
R1
NO2
39 NO2 44
43
73-92% yield
90-93% ee
R1 = iPr, Me(CH2)5, cyclohexyl, TIPSO(CH2)3
R2 = Ph, 4-Ph-C6H4, 4-NO2C6H4, 2-NO2C6H4, 2-BrC6H4, 2-thiophenyl, 2-naphthyl
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

NO2

Br
= CPh2OTMS
N NO2 43
colorless
45
O
R1 R1
H
39 colorless N

H Br
46 R1

H N
H H yellow-orange EDA complex
O 42
R1
e
SET
NO2

O2N 44

N N Br

48 R1 47 R1

chiral radical ion pair

Br

Scheme 6 Enantioselective photocatalytic alkylation of aldehydes with alkyl bromides.

Palomares, Pericàs, and coworkers introduced nontoxic and inexpensive semiconductor Bi2O3
as photocatalyst and a second-generation Macmillan catalyst for the enantioselective -alkylation of
hydrocinnamaldehyde with diethyl bromomalonate (Scheme 7).25 The strategy was accomplished by
combining bismuth-based semiconductors (such as Bi2O3 and Bi2S3) as low-band-gap photocatalysts
with Macmillan’s imidazolidinone as the chiral catalyst under sunlight or white light from a regular
fluorescent bulb. The semiconductor catalysis mechanistic cycle was presented by the authors, as
shown in Scheme 7. The reaction cycle is initiated by the promotion of electrons from the valence
band (VB) to the conduction band (CB) by light irradiation, followed by single electron transfer
Page 9 of 83 Organic & Biomolecular Chemistry

reduction of halide 50 to generate alkyl radical 52, which then underwent addition to enantiopure
View Article Online
DOI: 10.1039/D0OB00759E
enamine 53 to afford radical 54. Finally, oxidation to an iminium species by transfer of an electron
to the semiconductor afforded product 3, completing the semiconductor catalysis cycle. The EDA

Organic & Biomolecular Chemistry Accepted Manuscript


complex catalysis cycle was also found to be involved, but with a lower reaction rate than the
semiconductor catalysis cycle. As bismuth semiconductors are low-cost and nontoxic, while sunlight
is a low-cost and renewable energy source, this enantioselective photocatalysis process is
environmentally friendly and sustainable.
Melchiorre et al. has also applied the EDA complex strategy to the intermolecular
enantioselective -alkylation of cyclic ketones and alkyl bromides, using catalytic quantities of a
quinidine-derived primary amine (20 mol%) and TFA (40 mol%) in toluene under CFL irradiation
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

(23 W) to afford -alkylated products with high diastereoselectivity and enantioselectivity (Scheme
8).26 Subsequently, this strategy was also successfully employed in enantioselective - and -
alkylations of enals and aldehydes, respectively, with bromomalonates, using the Jørgensen–Hayashi
aminocatalyst under CFL irradiation condition. The key element of this process is the use of derived
chiral enamines as photosensitizers to trigger the asymmetric radical chain reactions (Scheme 8).27
The same group has since reported a more detailed mechanism for this photochemical
enantioselective -alkylation of aldehydes with electron-deficient alkyl halides.28 This reaction
could proceed via the EDA complex, as well as through direct photoactivation of the chiral enamine
to attain an electronically excited state that acts as a photoinitiator in the subsequent radical reaction.
Alternatively, similar -alkylations of aldehydes with -bromoesters (or -bromoketones) using the
Macmillan catalyst and [Fe(bpy)3]Br2 as photocatalyst under visible-light irradiation have been
reported by Ceroni, Cozzi, and coworkers (Scheme 8).29 This study demonstrated that [Fe(bpy)3]Br2
can adequately substitute for [Ru(bpy)3]2+ or other photosensitizers in enantioselective visible-light
catalysis reactions.
Organic & Biomolecular Chemistry Page 10 of 83

10

View Article Online


O N DOI: 10.1039/D0OB00759E
t
Bu • HCl R3
R3 N OHC
H 4 (20 mol %)

Organic & Biomolecular Chemistry Accepted Manuscript


R1 CHO R2
+ R2 Br Bi2O3 powder or Bi2S3 nanoparticles R1
(ca. 20 mol %), 2,6-lutidine, DMF, rt 51
49 50 23 W CFL 57-91% yield
82-98% ee
R1 = (CH2)6Me, (CH2)2Ph, cyclohexyl-CH2, (CH2)4CH=CHEt,
R2 = CO2Et, CO2Me, H; R3 = CO2Et, CO2Me, C(O)Ph, C(O)C6H4OMe

t N O
Bu
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R2 X
R1

N radical ion pair


O t O
N X Bu 57
t
Bu N
N
EWG R2 X
R1 EDA O
R1  to  EDA cycle
56 complex N
N
H t
Bu

R1 R2 58
R1 CHO

49
O
t N O R2
Bu H
R2 52
N R1 51
O
53
reduction
R1 N
R2 X
N
H t 50
Bu

R1 R2
cat. 4
semiconductor 54
O cycle
oxidation e- CB
R2
H
O light
R1
51 N h+ VB
N
H t
Bu

R1 R2 55

Scheme 7 Enantioselective -alkylation of aldehydes using the Macmillan catalyst and semiconductors.
Page 11 of 83 Organic & Biomolecular Chemistry

11

View Article Online


2014 Melchiorre
DOI: 10.1039/D0OB00759E

O O NH2
O 62 (20 mol %)
TFA (40 mol %) O
N

Organic & Biomolecular Chemistry Accepted Manuscript


Br
+ R1 NaOAc, tolutene n R1 N
n 60 0 °C, 23 W CFL 62
61
59 38–94% yield OMe
up to 95% ee
O NO2 62 (20 mol %) R2
O
TFA (40 mol %)
Br
n
+ NaOAc, tolutene
R2 0 °C, 23 W CFL n
63 NO2
59 64
R1 = Ph, 4-Ph-C6H4, 4-NO2C6H4, 2-NO2C6H4, 2-BrC6H4, 2-thiophenyl, 2-naphthyl
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R2 = NO2, CN

2015 Melchiorre
R2 Ar
CHO R2 29 (20 mol %) R3
2,6-lutidine (1 equiv) CHO N Ar
R3
+ EtO2C H OTMS
R1 Br CO2Et MeOtBu, rt, 23 W CFL 29
65 66 R1
67 Ar = 3,5-(CF3)2C6H3
R1 = Me, Me(CH2)4, TIPSO(CH2)2, MeCH2CH=CH(CH2)2
R2 = Me, Bn, CO2Et, HO(CH2)5; R3 = H, CO2Et

2015 Cozzi and Ceroni O N


t HOTf
Bu
R2
O N
H 4 (20 mol %)
Br OHC R3
OHC R1
+ R3
R2 [Fe(bpy)3]Br2 (2.5 mol %) R1 O
68 2,6-lutidine (2 equiv)
69 70
DMF, rt, 23 W CFL
40–83% yield
R1 = Ph, Me(CH2)4, cyclohexyl, EtCH=CH(CH2)3, 4-piperidinyl up to 97% ee
R2 = H, Ph, CO2Et, HO(CH2)5; R3 = CO2Et, 4-NO2C6H4,

Scheme 8 Enantioselective -alkylation of aldehydes.

In 2013, MacMillan et. al. reported a β-arylation of cyclohexanone with terephthalonitrile and
the photocatalyst Ir(ppy)3 and a cinchona-derived organocatalyst, Scheme 9.30

CN O
O Ir(ppy)3 (1 mol %) H N
74 (20 mol %)
+
DABCO, HOAc, DMPU NH2
H2O, rt, CFL N
CN
71 82% yield; 50% ee 73 CN 74
72

Scheme 9 Enantioselective -arylation of cyclohexanone.

In 2014, Luo and coworkers reported an enantioselective photocatalytic -alkylation of -


ketocarbonyls with -bromocarbonyls using a primary amine catalyst (Scheme 10).31 The author
believed that the reaction proceeded through direct photoredox catalysis of the alkyl bromide, with
the EDA reaction pathway also making a minor contribution. Phenacyl radicals are generated by both
photocatalytic pathways, followed by an additional reaction. Later, the same group reported the
Organic & Biomolecular Chemistry Page 12 of 83

12

feasibility of enantioselective reactions without any external photosensitizer.32 View Article Online
DOI: 10.1039/D0OB00759E
t
H
Bu Et O
O O N 78
O

Organic & Biomolecular Chemistry Accepted Manuscript


O NH2 Et OTf (20 mol %) R1
R1 R3 + Br
R4 R2 O R4
R2 Ru(bpy)3Cl2•6H2O (1 mol %)
76 NaHCO3 (1.0 equiv), CH3CN R3
75 77
rt, 33-W CFL
cyclic or acyclic 30-96% yield
up to 99% ee
acyclic: R1 = Me; R2 = Me, nPr, Bn, CH2COPh, CH2COPh-4-OMe
cyclic: R1–R2 = (CH2)3, (CH2)4, (CH2)5
R3 = OEt, OBn, OtBu, iBu, Et,
R4 = Ph, 4-PhC6H4, 4-CF3C6H4, 4-NO2C6H4, OBn,
R6
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

O O H
t
Bu Et O
R5 O N 78 OH
N Br
H + NH2 Et OTf (20 mol %)
n N
n R5
79 80 Ru(bpy)3Cl2•6H2O (1 mol %)
O
R6 NaHCO3 (1.0 equiv), CH3CN 81
n = 1, 2, 3 rt, 33-W CFL 55-90% yield
up to >99% ee
R5 = Ph, Bn, 3,4-Cl2C6H3, 4-MeC6H4, 4-MeOC6H4
R6 = H, Me, Br, Ph
Scheme 10 Enantioselective ‑alkylation of -ketocarbonyls.

Pericàs and coworkers developed an enantioselective cross-dehydrogenative coupling of


xanthenes with aliphatic aldehydes using visible-light photocatalysis and amino organocatalysis
(Scheme 11).33 The mechanism for the synergistic catalytic enantioselective cross-coupling of two
C(sp3)–H bonds has been studied using computational calculations and experimental evidence to
explain the origin of the enantioselectivity observed.
(1) Ru(bpy)3(PF6)2 (2 mol %)
Ar
Ar 29 (10 mol %) R3 CHO
R1 R2 + R3 CHO N Ar = 3,5-(CF3)2C6H3
X H OTMS
83
82 CH2Cl2, white LED, 35 °C R1 R2
X = O, S X
(2) NaBH4, MeOH, rt
84
R1, R2 = H, Me, MeO, di-Me, di-MeO, Et2N, Br, F up to 89% yield
R3 = Me, Et, nPr, iPr, nPent, BnO, BnO(CH2)2, BnO(CH2)3, PhCH2, 4-FPhCH2 up to 99% ee

Scheme 11 Enantioselective photoredox cross-dehydrogenative coupling.

In 2017, Luo et al. reported the enantioselective -alkylation of -ketocarbonyl compounds with
propiolic acids using photocatalytic decarboxylation and chiral aminocatalysis (Scheme 12).34 The
reactions afforded the expected -alkynylation product 87 and -alkylation adducts 88 as side
products. As shown in Scheme 12, a rational mechanism was proposed for this photoreaction,
showing that the addition of -imino radicals proceeded in a characteristic amino-catalyzed manner.
Page 13 of 83 Organic & Biomolecular Chemistry

13

View Article Online


t
Bu 78 DOI: 10.1039/D0OB00759E
NEt2
CO2H
NH2 TfOH R4
O O O R2
(20 mole %) O R2
R4

Organic & Biomolecular Chemistry Accepted Manuscript


R1 R3 + R1
R2
O R1 +
O
R4 R3 O 87 R3 O
O 88
85 I 18-35% yield
17-57% yield
86 OH up to 99% ee up to 92% ee
(1.5 equiv)
Ru(bpy)3Cl2 6H2O (1 mol %)
R1 = Me; R2 = Me, CH2Cl2, rt, Ar, blue LED
R1 = R2 = -(CH2)3-, -(CH2)4-
R3 = OMe, OEt, Me, Et, iPr, nBu, iBu, tBu, Bn, allyl, cinnamyl,
R4 = Me, F, Cl, Br, OMe, n-pentyl,
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

t
Bu 85
87
NHEt2
NH2
CO2
O
t
Bu I
NHEt2
O
H
N O I
t
Bu O
NHEt2 RuIII
R1 R3 SET BI
N R2 Ph R2
asymmeric
R1 aminocatalysis
SET photocatalysis
R3 O
RuII*

t
Bu
NHEt2
H blue LEDs
N O RuII
t
Bu
NHEt2 O R2
R1 R3
Ph
N O -H+ R2 R1
O
R1 R3 R3 O
R2 88

O O

O O O O
O
I I
I O I O I O
OH OH -CO2
O 86 O O
Ph -H2O hydration Ph Ph
O HO2C

Scheme 12 Enantioselective alkylation of -ketocarbonyls with propiolic acids.

In 2017, Melchiorre et al. reported an enantioselective -alkylation of enals with alkyl silanes
enabled by visible-light excitation of chiral iminium ions (Scheme 13).35 The enantioselectivity and
yield of the reaction were directly proportional to the amine catalyst oxidation potential (92–94), with
94 having the greatest oxidation potential owing to the introduction of gem-difluoride atoms on the
pyrrolidine ring and the bulky perfluoro-isopropyl substituents on the phenyl groups.
Organic & Biomolecular Chemistry Page 14 of 83

14

View Article Online


F
O F DOI: 10.1039/D0OB00759E
O 92 (20 mol %) Ar
TFA (40 mol %) H
R TMS N Ar
H
+

Organic & Biomolecular Chemistry Accepted Manuscript


CH3CN, rt R H OTDS
Ar
single high-power LED 92
89 Ar 90 (max = 420 nm) 91
Ar = 3,5-(C(CF3)2F)2C6H3
up tp 93% yield
TDS = thexyldimethylsilyl
up to 94% ee
Ar = Ph, 1-MePh, 4-FPh, 4-ClPh, 4-BrPh, 4-MeOPh, 4-CF3Ph, 2-ClPh, 3-BrPh, 3-MeOPh, 3,5-tBu2Ph
R = Ph, 4-MePh, 2-MePh, 2,4,6-Me3Ph, 4- MeOPh, 4-FPh, 4-BrPh, 2-naphthyl, 3-benzo[b]thiophenyl
PhS, N(Boc)Ph, N-carbazolyl, N-tosyl-3-indolyl

X
X
Ar 93 X = H, Ar = 3,5-(CF3)2C6H3 Eox (93 /93 = +1.57 V, Ag/Ag+ in CH3CN)
94 X = F, Ar = 3,5-(CF3)2C6H3 Eox (94 /94 = +2.20 V, Ag/Ag+ in CH3CN)
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

N Ar 92 X = F, Ar = 3,5-(C(CF3)2F)2C6H3 Eox (92 /92 = +2.40 V, Ag/Ag+ in CH3CN)


H OTDS

O N Ph TMS 96
t
Bu
Ph N
H 95 Eox (96 /96 = +1.74 V, Ag/Ag+ in CH3CN)
Eox (95 /95 = +1.80 V, Ag/Ag+ in CH3CN)

F = C(3,5-(C(CF3)2F)2C6H3)2OTDS
F
H 2O
89a
N
TFA LED
O
H TFA
H F
Ph
Ph F *
Ph 92 colored
91a
N

H TFA
H 2O
F Ph
H F excited
Ph N as a strong oxidant

Ph F Ph TMS
F
SET 90a
N

H Ph TMS

Ph solvent

PhCH2 solvent–TMS TFA

Scheme 13 Photochemical enantioselective -alkylation of enals with alkylsilanes.

In 2018, Melchiorre et al. reported the enantioselective photocatalytic conjugated alkylation of


enals using alkyl dihydropyridine as a radical precursor (Scheme 14).36 This reaction was dependent
on violet LED and the excitation of chiral imine ions, which made them strong oxidants to produce
alkyl radicals from 4-alkyl-1,4-dihydropyridine and activate the enantioselective radical conjugate
addition pathway of the ,-unsaturated aldehyde.
Page 15 of 83 Organic & Biomolecular Chemistry

15

Ar View Article Online


CHO DOI: 10.1039/D0OB00759E
O F
97 92 (20 mol %) F
+ H Ar
TFA (80 mol %)

Organic & Biomolecular Chemistry Accepted Manuscript


R N Ar
CH3CN–perfluorohexane H
EtO2C CO2Et R Ar OTDS
(4:1), –10 °C
LED (max = 420 nm) 99 92
N Ar = 3,5-(C(CF3)2F)2C6H3
up tp 83% yield
H 98 TDS = thexyldimethylsilyl
up to 94% ee

Ar = Ph, 2-MePh, 4-MePh, 4-MeOPh, 4-FPh, 4-ClPh, 4-BrPh, 2-ClPh


R = iPr, tBu, sBu, 3-pentyl, Cyp, Cy, cyclobutyl, tetrahydro-2H-pyran-4-yl , N-Boc-piperidin-4-yl,
pent-1-en-4-yl, 2-methylhept-2-en-6-yl
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ph OHC
CHO F
F
+ 97a 94 (20 mol %)
O O Ar
TFA (80 mol %) Ph
DHP N Ar
CH2Cl2, –10 °C, 20 h O H OTDS
O O
O LED (max = 420 nm) O 94
O Ar = 3,5-(CF3)2C6H3
O
99a TDS = thexyldimethylsilyl
O
98a
Scheme 14 Enantioselective conjugate alkylation of enals with alkyl dihydropyridines.

In 2017, MacMillan et al. developed an enantioselective photocatalytic -alkylation of


aldehydes with olefins (Scheme 15).37 The intra- and intermolecular variants of these reactions were
feasible via hydrogen atom transfer (HAT) catalysis, single-electron transfer (SET), and electron
borrowing through photoredox catalysis, in addition to enamine–iminium catalysis.
Organic & Biomolecular Chemistry Page 16 of 83

16

View Article Online


DOI: 10.1039/D0OB00759E
O H 102 (20 mol %) R2 H O
R2 Me
103 (3 mol %) O N
R3 104 (20 mol %) R1 t

Organic & Biomolecular Chemistry Accepted Manuscript


R1 t Bu
R3 Bu
N 1-naph N
DMSO–NMA X n
X n LEDs N H
101 Ir F 102
100 F
dr up to >20:1 60-91% yield N
n = 0, 1, 2 i
Pr
up to 95% ee t N
Bu
i
Pr SH
Me
103 PF6
i
Ir(Fmppy)2(dtbbpy)PF6 104 Pr

R1 = R2 = Me, -(CH2)3-, H
R1 = H and R2= Me, (CH2)2TMS, prenyl, 1,3-diphenyl-1H-pyrazol-4-yl
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R3 = H, Me
X = NTs, NBoc, CH2, C(CO2tBu)2, O

OHC R 108 (20 mol %) CHO Ar


109 (1 mol %) R Me
105 110 (10 mol %) N Ar
+ DME, 10 °C
107 t
Bu
N
H OTMS
LEDs 108
Ar Ar N
Me Ar = 3,5-(C(CF3)2F)2C6H3
Ir
106 56-92% yield Me
N
up to 93% ee t
Bu
t N
Bu
t
Bu SH
Me
PF6
109 t
Bu
Ir(dmppy)2(dtbbpy)PF6 110

R = Me, nHex, iPr, (CH2)4Cl, (CH2)3CO2Et, CH2NHCbz


Ar = Ph, 4- MeOPh, 4-FPh, 4CF3Ph, 3-pyridinyl, 1-benzyl-1H-pyrrol-3-yl, 2,3-dihydro-1H-inden-2-yl

107 105
= C(3,5-(C(CF3)2F)2C6H3)2OTMS

H 2O N H 2O
H

N
N
Ar H *IrIII
H oxidant
R
R

SET Photoredox IrIII


catalysis
HAT

IrII
N
reductant e
N
Ar H
H ArS
R
R

Ar HAT
catalysis ArS
106

H
H 2O
ArSH
110

Scheme 15 Enantioselective photocatalytic -alkylation of aldehydes.


Page 17 of 83 Organic & Biomolecular Chemistry

17

View Article Online


DOI: 10.1039/D0OB00759E
In 2018, Bach et al. showed that eniminium ions prepared from the corresponding enals could
be promoted by visible-light-induced energy transfer to their respective triplet states.38 The

Organic & Biomolecular Chemistry Accepted Manuscript


intermolecular [2+2] photocycloaddition of eniminium ion 111 with 2,3-dimethylbutadiene using
Ru(bpy)3(PF6)2 (2.5 mol%) as photocatalyst at 40 °C under visible-light irradiation for 3.5 h afforded
the adduct in 78% yield with 88% ee (Scheme 16). Later, Bach and coworkers reported a visible-
light-enabled ( = 457 nm) chiral oxazaborolidine-catalyzed ortho photocycloaddition of
phenanthrene-9-carboxaldehydes and alkenes (Scheme 17).39 The authors observed that coordination
of the chiral Lewis acid with the aldehyde caused a bathochromic shift in the original n* absorption,
which enabled the reaction to be activated by visible-light irradiation.
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ar
(1) Ru(bpy)3(PF6)2 (2.5 mol %) OHC
N Ar
light, (max = 420 nm)
OTDS
H + (2) aq. NaOH Ph
BF4
Ph 111 112 113
TDS = thexyldimethylsilyl 78% yield; dr = 94:6
88% ee
Ar = 3,5-(CF3)2C6H3

Scheme 16 Enantioselective photocatalytic [2+2] cycloaddition.

OHC Ar1 Ar
117 (20 mol %) 1

–78 °C, CH2Cl2 OHC H


+ O
visible light N B
R1 R2 ( = 457 nm)
Br3Al Ar2
24 h R1 R2
114 115
116 117
R1 = H, 2-Me, 2-Cl, 3-F, 3-Me Ar1 = 3,5-dimethylphenyl
66-93% yield
R2 = H, 5-Me, 6-CF3, 6-F Ar2 = 2,6-dimethylphenyl
up tp 96% ee

Scheme 17 Photocycloaddition of phenanthrene-9-carboxaldehydes and olefins.

In 2018, Melchiorre reported a visible-light photocatalysis strategy for the intramolecular EDA
complex, applying this method to enantioselective radical conjugate additions of enones (Scheme
18)40 Key to this method was a chiral carbazole catalyst that contained no photosensitive unit, but
was able to induce the transient formation of an intramolecular EDA complex that absorbed visible
light and guided the enantioselective reaction of the photochemically formed radicals.
Organic & Biomolecular Chemistry Page 18 of 83

18

View Article Online

R4 O CF3
DOI: 10.1039/D0OB00759E
O R3 121 (20 mol %) t
high-power single LED N Bu
R1
+ ( = 420 nm) n

Organic & Biomolecular Chemistry Accepted Manuscript


R2
n R1 N salicyclic acid (40 mol %) R4
H2O (2 equiv) R2 N 121
118 119 SiMe2Ph R3
CH3CN, 35 °C, 48 h
n = 1-3
120 NH2
R1 = Me, Et, Bn, cyclopropyl 44-95 % yield
up to 95% ee t
R2 = H, Me, OMe, Ph, 4-MePh, 4-ClPh, 2-furyl, 2-thienyl Bu
R3 = H, Et, F; R4 = H, Me
CF3

R3 O
O
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R4 121 (20 mol %) R1


R2 high-power single LED N
R2
+ R1
( = 420 nm)
n
n N benzoic acid (40 mol %)
H2O (2 equiv) R3
122 123 SiMe2Ph CH3CN, 35 °C, 48 h R4
n = 1-2 124
R1 = Me, Et; R2 = H, Cl; R3 = H, Cl, F, MeO 57-82 % yield
up to 93% ee

Scheme 18 Enantioselective radical conjugate additions of enones.

In 2018, Melchiorre et al. reported a visible-light-induced enantioselective radical cyclization–


addition cascade of alkenoic acids and ,-unsaturated aldehydes by activating unfunctionalized
alkenes with a photoactivated chiral iminium ion (Scheme 19).41 This method is a further example
of visible light photocatalysis without a photocatalyst.
Page 19 of 83 Organic & Biomolecular Chemistry

19

92 (20 mol %) R2 R2 View Article Online


F DOI: 10.1039/D0OB00759E
R3 TFA (60 mol %) OHC F
HexF /CH3CN (1:1) Ar
OHC CO2H O
R1 + R3 30 °C, 16 h R1 O N Ar

Organic & Biomolecular Chemistry Accepted Manuscript


R2 R2 High power single LED H OTDS
125 126 (420 nm) R3 R3 127
92
up to 77% yield Ar = 3,5-(C(CF3)2F)2C6H3
up to 95% ee TDS = thexyldimethylsilyl
R1 = Ph, 4-FPh, 4-ClPh, 4-MePh, 3-ClPh, 3-MePh, 2-ClPh, 4-TMSPh, 4-BpinPh
R2 = Me, (CH2)5; R3 = Me, H, Ph, Et, (CH2)5

F = C(3,5-(C(CF3)2F)2C6H3)2OTDS
F
H 2O
125a LED
N
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

TFA 420 nm
OHC H TFA
O
F
Ph O Ph F *
92 colored
127a
N

H TFA
H 2O
Ph
F excited
F as a strong oxidant
Ph N
F
F 126a
SET
O N

O H
HO2C
Ph

O O

Scheme 19 Enantioselective radical addition and cyclization cascade.

In 2019, Pedro, Vila, and coworkers reported an enantioselective Mannich reaction of 3,4-
dihydroquinoxalin-2-one derivatives with ketones by merging visible-light organophotocatalysis
(eosin Y) with asymmetric aminocatalysis ((S)-proline), achieving good yields and excellent
enantioselectivities (up to 99% ee) for the chiral quinoxaline products (Scheme 20).42 Furthermore,
the reaction was feasible at a 5-mmol scale under sunlight irradiation.

O R3
H eosin Y (2 mol %)
N H
(S)-Proline (20 mol %) N
R1 R4
+ R3 R4
DMF, rt R1
O
N O blue LEDs N O
R2 129 R2
128
130
R1 = Me, MeO, CF3, F, CO2H, Cl,
R2 = H, Me, Bn, allyl, CH2CO2Me 55-94% yield
R3 = H, OH 75-99% ee
R3 = R4 = (CH2)3, (CH2)2
R4 = H, Me, iPr, noctyl, di-MeO

Scheme 20 Enantioselective Mannich reaction.


Organic & Biomolecular Chemistry Page 20 of 83

20

View Article Online


DOI: 10.1039/D0OB00759E
He, Guan, and coworkers reported a photocatalytic oxidative dearomatization of 2-arylindoles
and enantioselective Mannich reaction with ketones through amino catalysis (using L-proline and D-

Organic & Biomolecular Chemistry Accepted Manuscript


proline, respectively) to give C2-quaternary indolin-3-ones (Scheme 21).43
O R3
O Ru(bpy)3Cl2•6H2O O
(S)
R2 L-Pro, EtOH R4
R1 + R3 R1 (S) R2
N R4 2,6-lutidine, O2 N
H 9 W blue LEDs H
131 132 133
R1 = H, F, Cl, Br, Me, MeO
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

up to >20:1 dr 55-86 % yield


R2 = H, F, Cl, Br, Me, CO2Me, MeO, Me 95-99% ee
R3 = H, Me; R4 = H, Me, Bn, nC8H17
R3 = R4 = (CH2)4, (CH2)3

Scheme 21 Photocatalyzed oxidative dearomatization and enantioselective Mannich reaction.

Recently, Melchiorre et al. accomplished a photocatalytic and enantioselective reaction of in-


situ-generated chiral and colored iminium ions of ,-unsaturated aldehydes with allenic acids via
single-electron oxidation, which involved a cascade addition and cyclization process, affording
polysubstituted bicyclic lactones (Scheme 22).44
O
R1 92 (20 mol %) O F
R2 R2 H R2 F
high power LED(420 nm)
Ar R1 C R2 Ar
CH2OH CO2H Zn(OTf)2 (25 mol %) H Ar
N
134
+ H CH3CN, 35 °C, 18-72 h Ar O H OTDS 92
135 R1 R1
136 Ar = 3,5-(C(CF3)2F)2C6H3
Ar = Ph, 4-ClPh, 4-MePh, 3-MeOPh
R1 = Me, Et, (CH2)5, (CH2)11 42-56% yield TDS = thexyldimethylsilyl
R2 = Me, Et, (CH2)5, 57-74% ee

Scheme 22 Enantioselective cyclization of enals with allenic acids.

An enantioselective -alkylation of ketones using a cinchona-based primary amine catalyst and


alkyl radicals, generated in situ from alkyl chloride and dithiocarbamate anion catalyst A under blue-
light irradiation, has been reported by Melchiorre and coworkers (Scheme 23).45
Br
O 30 (20 mol %) R O
140 (20 mol %)
Cl R N
+ NaOAc (2 equiv) 30 SK
X n 138 toluene, 10 °C n S
137 blue LEDs, 24 h 139 OMe Et
n = 1, 2 up to 87% yield
up to 95% ee NH2
X = CH2, MeCH, tBuCH, PhCH, C(OCH2CH2)2, BocN N
R =CN, 2,4-(NO2)2Ph, PhC(O) •2TFA
N
140

Scheme 23 Enantioselective α-alkylation of ketones with alkyl chloride


Page 21 of 83 Organic & Biomolecular Chemistry

21

View Article Online


DOI: 10.1039/D0OB00759E

3. Chiral Lewis acid catalysis with photocatalyst

Organic & Biomolecular Chemistry Accepted Manuscript


In 2014, Yoon et al. reported an enantioselective catalytic [2+2] photocycloaddition of ,-
unsaturated ketones using a dual-catalysis approach comprising a photocatalyst and chiral Lewis acid
catalyst (Scheme 24).46 This strategy addressed the long-standing problem of how to suppress the
racemic background [2+2] reaction under regular photocatalytic reactions. However, as shown at the
bottom of Scheme 24, under visible-light conditions, the uncatalyzed racemic background
cycloaddition was unable to proceed, reducing the effect on the net enantiomeric excess of the product.
Inspired by the pioneering study of Bach,47 Yoon observed that complexing the Lewis acid with
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

enones resulted in a bathochromic shift that provided longer-wavelength absorption, allowing the
[2+2] reaction to be activated under visible-light irradiation. Furthermore, introducing an H-bonded
substrate into the reactants produced a similar effect and enantioselectivity. Consequently, the [2+2]
cycloaddition of enones was achieved using photocatalyst Ru(bpy)3Cl2, Lewis acid Eu(OTf)3, and
chiral Schiff base ligand 144 to give cyclobutane adduct 143 in 34%–71% yield and 86%–93% ee.
Furthermore, different diastereoselective cyclobutane adducts were obtained by the alteration of the
chiral Lewis acid structure resulting from the reduction of chiral Schiff base ligand 144 to 146
(Scheme 24). Subsequently, Yoon applied a similar methodology to the enantioselective [2+2]
cycloaddition of 2′-hydroxychalcone and dienes using chiral Lewis acid-catalyzed photoexcitation
energy transfer (Scheme 25).48 Chiral Lewis acid coordination of the chalcone substrate significantly
reduced its triplet energy, allowing cycloaddition reactions to be performed with enantioselectivity.
Recently, Yoon used this approach in crossed enantioselective photocatalytic cycloadditions of 2’-
hydroxychalcone and styrene (Scheme 26).49 As expected, Lewis acid coordination of 2′-
hydroxychalcone 152a (ET = 54 kcal/mol) dramatically decreased the energy of Lewis-acid-bound
compound 157 (ET = 32 kcal/mol), allowing triplet energy transfer from 3Ru*(bpy)32+ (ET = 45
kcal/mol) and affording an enantioselective cycloaddition process. More recently, merging Lewis
acid catalysis and photocatalysis for enantioselective [2+2] cycloaddition has been extended to the
reactions of regular cinnamate esters with styrenes (Scheme 27).50
Organic & Biomolecular Chemistry Page 22 of 83

22

View Article Online


DOI: 10.1039/D0OB00759E
N
N
O O O O
OH

Organic & Biomolecular Chemistry Accepted Manuscript


HN
R1 O O n R2 R3
144 (20 mol %) Bu
+ R3
141 R2 142 Ru(bpy)3Cl2 (5 mole %)
R1 143
5 equiv Eu(OTf)3 (10 mol %)
i-Pr2NEt (2 equiv), CH3CN 34-71 % yield
R1 = Me, Et, i-Pr, CH2OBn
23 W compact fluorescent light 2.5-7 : 1 dr
R2 = Ph, C6H4Cl, C6H4Br, C6H4OMe,
C6H4F, 2-naphthyl, 2-furanyl 86-93 % ee
R3 = Me, Et
N
N
H
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

O O O O
R1 O O OH HN
+ R
146 (20 mol %) n
Bu R2 R3
141 R2 142 3
5 equiv Ru(bpy)3Cl2 (5 mole %)
Eu(OTf)3 (10 mol %) R1 145
R1 = Me, i-Pr, CH2OBn
R2 = Ph, C6H4Cl, C6H4Br, C6H4OMe, i-Pr2NEt (2 equiv), CH3CN 49-80 % yield
C6H4F, 2-naphthyl, 2-furanyl 23 W compact fluorescent light 1.5-4.5 : 1 dr
R3 = Me, Et, i-Bu 84-95 % ee

CFL

MLn* Ru(bpy)32+ Ru(bpy)32+*


O
141 photoredox i
R2 Pr
catalysis
N
Lewis acid R1 + e- Et i
Pr
- e-
catalysis
MLn* i
Pr
O Ru(bpy)3+
O N
i
Et Pr
R3 + R2 147

142 R1

MLn* [2 + 2]
– e-
direct photocycloaddition
O O O O
R2 R3 [2 + 2] R2 R3
+
CFL
R1 141 R1 142
143 or 145

Scheme 24 Enantioselective photocatalyzed [2 + 2] cycloaddition.


OH O
OH O Ru(bbp)3(PF6)2 (2.5 mol %)
(S,S)-tBu-PyBox (15 mol %)
Sc(OTf)3 (10 mol %) R1
R1 R2 + i
PrOAc–CH3CN v/v = 3:1
148 23W CFL, 20h
149 150 R2
R1 = H, Me, Cl
R2 = H, Br, Cl, OMe dr 2-4:1 66–84% yield
up to 98% ee
O O
N
N N
t t
Bu 151 Bu
(S,S)-tBu-PyBox

Scheme 25. Enantioselective catalytic [2 +2] cycloadditions.


Page 23 of 83 Organic & Biomolecular Chemistry

23

View Article Online


OH O DOI: 10.1039/D0OB00759E
t
OH O R2 (S,S)- BuPyBox (15 mol %)
Sc(OTf)3 (10 mol %)
R1 Ar2
Ar1 + Ar2 Ru(bpy)3(PF6)2 (2.5 mol %)

Organic & Biomolecular Chemistry Accepted Manuscript


R1 i Ar1 R2
153 PrOH–CH3CN (3:1)
152 (10 equiv) 23W CFL, rt 154
up to 97% yield
up to >99% ee
R1 = F, Cl, Me, MeO; R2 = H, Me
Ar1 = Ph, 2-FPh, 2-ClPh, 2-CF3Ph, 2-MeOPh, 4-BrPh, 4-CF3Ph, 4-MeOPh, 3,5-(MeO)2Ph, 3-furanyl
Ar2 = Ph, 2-MePh, 2-ClPh, 4-MePh, 4-ClPh, 4-BrPh, 4-CF3Ph, 4-CO2MePh, 4-B(pin)Ph, 4-tBuOPh,
3-ClPh, 3-MeOPh, 3,5-(MeO)2Ph, SPh

OH O O
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

OMe OMe
MeO
O O
N
N N Br
O O OMe
t
Bu 151 t
Bu OMe OMe
(S,S)-tBuPyBox O OMe O OMe
155 156
80% yield, dr: 3:1; 94% ee a norlignan cyclobutane
natural product

OH O
OH O
Ph
152a
Ph (ET = 54 kcal/mol)
Ph (S,S)-tBuPyBox
154a

Sc
O O

Sc asymmetric Ph
O O catalysis 157
(ET = 32 kcal/mol)

Ph
Ph 3
Ru*(bpy)32+
159 (ET = 45 kcal/mol)
Sc
O O
photoredox
catalysis ISC
Ph
Ph Ru(bpy)32+ 1
158 Ru*(bpy)32+
153a (ET = 32 kcal/mol)
CFL

Scheme 26 Enantioselective photoredox cross-cycloaddition.


Organic & Biomolecular Chemistry Page 24 of 83

24

O View Article Online


DOI: 10.1039/D0OB00759E
O 163 (25 mol %)
RO
[Ir(Fppy)2(dtbbpy)](PF6) (1 mol %)
RO Ar1 + Ar2

Organic & Biomolecular Chemistry Accepted Manuscript


Ar2
161 CH2Cl2 Ar1
160 (5 equiv) blue LED, –25 °C, 24h 162
dr up to >20:1 67-97% yield
i
R = Me, Et, Pr, Bn up to 99% ee
Ar1 = Ph, 4-MeOPh, 3-BrPh, 2-MePh, 2-thienyl
Ar2 = Ph, 4-MePh, 3-MeOPh, 3-FPh, 4-CF3Ph, 4-ClPh, 2-ClPh, 4-PinBPh

PF6
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

t
Bu
N
H
N
F IrIII
F O
N B NTf2
N
N H F
t
Bu F

[Ir(Fppy)2(dtbpy)]PF6 F
163

Scheme 27. Enantioselective catalytic [2+2] cycloadditions of cinnamate esters and styrenes.

Yoon et al. reported the enantioselective conjugate addition of -amino radical 167, generated
in situ from silylmethyl aniline 164, to crotonyl oxazolidinone 165 using synergistic Lewis acid
catalysis and photocatalysis (Scheme 28).51 The process provided Michael adduct 166 in good
yields (60%–96 %) with high enantioselectivities (85%–96% ee). A rational mechanism was
proposed by the authors. Initial photoredox catalysis of the -silylamine 167 affording -amino
radical 165. Concomitantly, chiral Lewis acid (pybox)Sc(OTf)3 catalyzed the conjugate addition of
167 to 165 and controlled the stereochemistry by activating the N-acyl pyrazolinone amide group in
165.
Page 25 of 83 Organic & Biomolecular Chemistry

25

View Article Online


R1 O Ru(bpy)3Cl2 (2 mol %), DOI: 10.1039/D0OB00759E
Sc(OTf)3 (15 mol %), (S,S)-151 (20 mol %) R1 R3 O
N TMS R3 R4 Bu4N+Cl (30 mol %)
R2 + N
R2 R4

Organic & Biomolecular Chemistry Accepted Manuscript


164 165 (2 equiv) degassed MeCN, 23 W CFL
166
R1 = Me, Ph, Bn, Boc
60-96% yield
R2 = Ph, 4-FC6H4, 2-FC6H4, 4-ClC6H4, 4-BrC6H4, 4-MeC6H4, 4-MeOC6H4, 3-MeOC6H4,
85-96% ee
Pr, iPr, BnOCH2, tBu, 3-furanyl
R3 = Me,
R4 = O O
O O
N N
N O
N N N
Et i
Bu (S,S)-151 i
Bu
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

CFL

ScLn*
Ru(bpy)32+ Ru(bpy)32+*
+ TMS+ R1 R3 O
photoredox
N R1
R2 R4 catalysis
N TMS
170 R2
+ e- - e-
R1 R3 OTMS 164
N R1
R2 R4 Ru(bpy)3+
N TMS
171 ScLn* R2
R1 R3 O
H+ N TMS+
5 R2 R4
7 169 R1
166 164
N
R2
Lewis acid
catalysis 167

ScLn*
O O
ScLn*
168
R3 R4 chiral Lewis acid R4
activation
165 R3

Scheme 28 Enantioselective -amino radical addition.

In 2018, Zheng and coworkers reported an enantioselective reductive alkylation of nitrones to


aromatic aldehydes using the synergistic action of photocatalysis and Lewis acid catalysis to prepare
vicinal amino alcohols (Scheme 29).52 The chiral Lewis acid was an essential template for
assembling the two intermediates in this reaction, and activated the asymmetric radical coupling
reaction.
Organic & Biomolecular Chemistry Page 26 of 83

26

View Article Online


i
Pr DOI: 10.1039/D0OB00759E
Ru(bpy)3(PF6)2 (2.0 mol %) OH
O O Sc(OTf)3 (15 mol %) R3 i
Pr
N R3 175 (18 mol %) R2
+ O

Organic & Biomolecular Chemistry Accepted Manuscript


R1 H i
TEEDA (4.0 equiv) N NH Pr
R2 R1 OH
CH2Cl2, 0 °C R4
173 R4 O
65 W CFL, 48 h 174 N
172
up to 99% yield
R1 = Bn, Me, iPr; R2 = iPr, nPr, Me, cHex, tBu; R3 = H up to 99% ee
or 175
R2 = R3 =Me, (CH2)4, (CH2)5
R1 = R2 = (CH2)3, (CH2)4, CH2CH2CMe2 N O
R4 = H, F, MeS, MeO, Br, NHAc
i
NH Pr
O
i
CFL Pr
i
Pr H OH
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

i
Pr
H OH i
Pr H O
N
Ph Ru(bpy)32+ Ru(bpy)32+*
Bn 174a H
N
O [Sc] *
Ph photoredox i
Pr
Bn catalysis
N
Et i
- e- Pr
+H + e-
172a i
Pr
+ Ru(bpy)3+
173a N
Et i
Pr
Lewis acid
catalysis
i
Pr H H i
PrH O
O
H
N
O [Sc] *
H
N
O [Sc] *
Ph
Ph Bn
Bn

i +H
Pr i H
Pr O
N
i
Pr i
Pr H
N
O *
[Sc]
N Ph
Et i
Pr Bn

Scheme 29 Enantioselective -amino radical addition.

In 2019, Xu et al. reported an enantioselective alkylation of 2-acyl imidazole with


bromodifluoroacetamide under the synergistic action of photocatalysis and chiral Lewis acid catalysis,
providing chiral -keto amides bearing a geminal difluoro group (Scheme 30).53 Catalyst –179 was
used as a chiral Lewis acid here, rather than as a conventional chiral photocatalyst.
Me
Ir(ppy)2(dtbbpy)PF6 Ar
O (1 mol %) N O
O t
Bu
N Ar -179 (4 mol %) R S
Br R N N
N PF6
Me
N + F F R
DIPEA (2.0 equiv) O F F R N C
Me CH2Cl2, rt N
23W CFL, N2 178 71-96% yield Rh
176 177 up to >99% ee N
N C
Me
Ar = Ph, 4-FPh, 4-ClPh, 4-BrPh, 4-CF3Ph, 4-MePh, 4-tBuPh, 4-MeOPh,
4-EtOPh, 4-PhPh, 4-MeSPh, 3-ClPh, 3-MePh, 2-MePh, 3,4-di-MeOPh, S t
Bu
3,4,5-tri-MeOPh, 3-EtO-4-EtC(O)OPh, 3-thienyl, 2,3-dihydrobenzofuran-5-yl chloride
R = (CH2)2, (CH2)3, (CH2)4, (CH2)2O(CH2)2, (CH2)2S(CH2)2, (CH2)2NPh(CH2)2, -179
(CH2)2NBoc(CH2)2

Scheme 30 Enantioselective alkylation of 2-acyl imidazole.

In 2019, Jiang et al. reported enantioselective coupling reactions of toluene with N-tosyl isatins
Page 27 of 83 Organic & Biomolecular Chemistry

27

using a La(OTf)3/pybox catalyst (pybox comprises a chiral 3-hydroxy-3-benzyl-substituted 2-Online


View Article
DOI: 10.1039/D0OB00759E
oxindoles) under blue LED irradiation (Scheme 31).54 Furthermore, the enantioselective coupling
reaction of toluene and acenaphthoquinone was achieved using a chiral phosphoric acid (CPA)

Organic & Biomolecular Chemistry Accepted Manuscript


catalyst under visible light irradiation. Notably, PhCl was the best solvent for this reaction, with other
solvents, including CH2Cl2, CMPE, MTBE, DME, THF, EtOAc, tBuPh, and BrPh, failing to afford
the product or resulting in very low product yields.
Ar R1 NMe2
O La(OTf)3 (20 mol %) HO
R1 183 (22 mol %) R2
X O + R2 Ar
O
N 4 Å MS, argon N O O
N
PhCl, 25 °C
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ts 181 X Ts N N
3 W blue LED, 48-60 h 182
180
63-94 % yield Et 183 Et
X = H, F, Cl, Br, Me Me Me
t
up ro 98% ee
Ar = Ph, 4-MePh, 4-FPh, 4-ClPh, 4-BrPh, 4- BuPh, 4-PhPh,
4-MeOPh, 4-EtPh, 3-MePh, 2-MePh, 2-MeOPh, 3,5-Me2Ph, Ph
2,4,5-Me3Ph, 2,3,4,5,6-Me5Ph, 3-tBu,5-MePh, 5-methylfuran-2-yl,
5-methylthiophene-2-yl
R1 = H, and R2 = Me, Et, Bn, Ph; or R1 = R2 = Me
Ph

O O
O O 187 (20 mol %) O HO R
P
183 (22 mol %) OH
O
+ R–H
4 Å MS, argon
Ph
185 PhCl, –5 °C
3 W blue LED, 60 h
184 186
187
40-76 % yield Ph
R: Bn, 4-PhCH2, 4-MePhCH2, 3-MeOPhCH2, PhMe2C up ro 90% ee
9H-fluoren-9-yl

Scheme 31 Enantioselective coupling of toluene with N-tosyl isatins or acenaphthoquinone.

4. Chiral Lewis base catalysis with photocatalyst


Smith et al. reported a sequential and stepwise enantioselective alkylation of 4-nitrophenyl esters
to 2-phenyl-1,2,3,4-tetrahydroisoquinoline using Ru(bpy)3Cl2 as the photocatalyst and isothiourea
tetramisole as the Lewis base catalyst (Scheme 32).55 The ester product was converted to the
corresponding amide by treating with benzylamine. The reaction with 4-
nitrophenoxytetrabutylammonium (TBAPNP) added resulted in an increased yield, which mighty be
due to the combined effect of increased reaction mixture polarity and the promotion of catalyst
turnover by 4-nitrophenoxylate.
Organic & Biomolecular Chemistry Page 28 of 83

28

View Article Online


(1) Ru(bpy)3Cl2 (0.5 mol %) DOI: 10.1039/D0OB00759E
CBrCl3 (1.5 equiv) H
CH3CN, blue LED, 2 h NPh N+
NPh NR2R3 O

Organic & Biomolecular Chemistry Accepted Manuscript


(2) O R1
189
188 H - N+
R1 (1.5 equiv) O 191 O
OPNP
51-86 % yield
PNP = p-nitrophenyl dr ~75:25 TBAPNP O-
up to 88% ee
190•HCl (5 mol %), TBAPNP (1 equiv) N
Ph
i
Pr2NEt (1.5 equiv), THF, –10 °C, 24-36 h N S
(3) R2R3NH (5 equiv), –10 °C, 24 h 190

R1 = 4-MePh, 2-MePh, 4-CF3Ph, 4-BrPh, 4-PhPh, 4-MeOPh, 3-MeOPh, 3-thienyl, 2-naphthyl,


1-naphthyl, MeCH=CH, iPrCH=CH
R2 = Bn and R3 = H
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R2 = R3 = 1-pyrrolidinyl, 4-morpholinyl, 1-Boc-piperazin-4-yl chloride

BrCCl3
2+
Ru(bpy)3
blue LED
CCl3 + Br

photoredox
Ru(bpy)32+* catalysis Ru(bpy)3+ HCCl3

Br
N – e- N –H N
Ph amine oxidation Ph Ph
188 191 192
O
S
R
base–H N
N
Ph
base

O H
S NPh
R Ph
N
N N
OAc N
Ph R
H S
O

N BnNH2
Ph
189 N S

190 191

Scheme 32 Enantioselective addition using isothiourea catalyst.

5. Phase transfer catalysis with photocatalyst


Melchiorre and coworkers developed an enantioselective perfluoroalkylation of cyclic
‑ketoesters using a chiral phase transfer catalyst with the in-situ-generated electron donor–acceptor
complex (EDA complex) strategy (Scheme 33).56 The yellow-colored EDA complex was initially
produced after the aggregation of chiral enolate 193 with R3I (a perfluoralkyl iodide). Promoted by
white LED light irradiation, the electron-deficient R3 radical was formed via single-electron transfer
Page 29 of 83 Organic & Biomolecular Chemistry

29

and reductive cleavage of the C–I bond. Subsequently, the R3 radical reacted with the chiral PTCOnline
View Article
DOI: 10.1039/D0OB00759E
196–enolate 193 complex to give radical adduct 195a, which was trapped by R3I to afford iodide
195b and regenerate the R3 radical. Iodide 195b then released an iodide anion, affording product 195.

Organic & Biomolecular Chemistry Accepted Manuscript


O 196+Br- (20 mole %), O
Cs2CO3 (2 equiv) R3
R1 CO2R2 + R3 I
C6H5Cl–C8F18 (2:1)
R1
CO2R2
194 rt, white LEDs, 64 h
193 195
R1 = F, Cl, Br, CF3, Me, MeO 38-71% yield
R2 = Me, tBu 78-96% ee
R3 = CF3, C4F9, C6F13, C8F17, C10F21

R3 I
194
Cs2CO3 chiral PTC R3 I Aceptor
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

196 O  to 
196+Br-
CO2R2

F3C Donor
OH
EDA catalysis
N O
colored EDA complex
N F
R1 CO2R2
Br
F light
193
196+Br- (a chral PTC) F

chiral PTC
196 O R3 +I

CH2O2R2
Cs2CO3

196+193–

196+X– R3
chiral PTC
O 196 O
R3
R1 CO2R2
CO2R2
PTC asymmetric
195 catalysis 196+193–
196
I O
R3

CO2R2
195b
O
R3

R3 CO2R2
195a
R3 I
194

Scheme 33 Enantioselective perfluoroalkylation of cyclic -ketoesters.

6. Chiral Brønsted acid catalysis with photocatalyst

The first procedure combining visible-light photocatalysis with chiral Brønsted acid catalysis
was developed by Knowles et al. in 2013 for an enantioselective aza-pinacol cyclization (Scheme
34).57 This enantioselective intramolecular reductive coupling of ketones and hydrazones was
achieved via a ketyl radical intermediate, which was enabled by a concerted proton-coupled electron
transfer (PCET) process. The operation was promoted by both photocatalyst Ir(ppy)2(dtbpy)PF6 and
Organic & Biomolecular Chemistry Page 30 of 83

30

CPA catalyst (R,R)-199. The synergistic action of the two catalysts is depicted in Scheme 34. Initial
View Article Online
DOI: 10.1039/D0OB00759E
excitation of the iridium photocatalyst followed by reduction of the Hantzsch ester (HE, diethyl 1,4-
dihydro-2,6-dimethyl-3,5-pyridinedicarboxylate, 200)58 provides IrII(ppy)2(dtbpy) as a strong

Organic & Biomolecular Chemistry Accepted Manuscript


reducing agent for the subsequent PCET-mediated ketone reduction, affording ketyl intermediate 203.
The intramolecular C–C bond coupling reaction of 203 produces 204, followed by H-atom transfer
from HE to give product 198 and HE radical 201, which is then oxidized by IrIII(ppy)2(dtbpy)* to
yield acidic pyridinium ion 202. Phosphate anion 205 can be protonated by ion 202 to regenerate
CPA 199 and complete the acid catalysis cycle. This synthetic methodology provided a new pathway
for catalytic enantioselective radical reactions.
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ir(ppy)2(dtbby)PF6 (2 mol %) SiPh3


(R,R)-199 (10 mol %) NHR2
O HE (1.5 equiv), dioxane R1 O O
R2 blue LEDs, rt
R1 n N HO P
n O OH
197 n = 1, 2
198
SiPh3
R1 = Ph, C6H4OMe, C6H4CF3, C6H4Me, 45-96% yield
2-benzofuranyl, 2-naphthyl, 2-thiophenyl, 77-95% ee (R,R)-199

H
EtO2C CO2Et EtO2C CO2Et
EtO2C CO2Et
Me N Me Me N Me
O H 202 H
P
O * 201 Me N
H
Me
O O oxidation
H O O e- or 200
204 O H O
P *
O oxidation
O
R1 e-
N R1 II EtO2C CO2Et
Ir (ppy)2(dtbpy)
R2 N 203
H R2
H Me N Me
200 H
+H Acid
201 reduction photoredox
catalysis + e- IrIII(ppy)2(dtbpy)*
catalysis

OH
HO PCET
R1
P
O *
NHR2
O O Blue LED
H 198 O O
205
HO
P * IrIII(ppy)2(dtbpy)
O O
202
EtO2C CO2Et 199 R1
N
R2 197
Me N Me H
206

Scheme 34 Enantioselective aza-pinacol cyclization.

Ooi et al. reported an enantioselective -coupling of N-arylaminomethanes with N-sulfonyl


aldimines through the cooperation of a photocatalyst and ionic Brønsted acid catalyst under 15-W
white LED irradiation (Scheme 35).59 The process was achieved using P-spirocyclic
arylaminophosphonium barfate (210•BARF, BARF = [3,5-(CF3)2C6H3]4B) as the chiral ionic
Brønsted acid catalyst and an Ir(III) species as the photoredox catalyst via synergistic catalysis. Initial
reductive quenching of the excited [*IrIII]+BARF species by amine 207 affords the corresponding
amine radical cation with negatively charged counterion BArF (213•BARF) and a neutral Ir(II)
complex. The resulting Ir(II) complex reduces imine 208 via single electron transfer to give radical
anion 212, which then pairs with Ir(III)+ to form complex 212•Ir(III)+. After prompt ion exchange
with 210•BARF, complex 212•Ir(III)+ forms chiral ion pair 210•212 and regenerates the photocatalyst
Page 31 of 83 Organic & Biomolecular Chemistry

31

(211•BARF) to complete the photocatalysis cycle. Meanwhile, aminomethyl radical 214, generated View Article Online
DOI: 10.1039/D0OB00759E
by the deprotonation of radical cation 213•BARF with base, reacts with imine radical anion complex
210•212 under asymmetric induction by the chiral P-spirocyclic Brønsted acid to give product 209

Organic & Biomolecular Chemistry Accepted Manuscript


and regenerate chiral acid 210 for continuation of the ionic Brønsted acid catalysis cycle.
Ms 210•BARF (4 mol %) R1 Ms
R1 HN
N 211•BARF (1 mol %)
N Me + N
H R3 visible light, toluene, rt R2 R3
R2
209
207 208
63-90% yield; 85-97% ee

R1, R2 = Ph, Ph; 2-naphthyl, Ph; 3-MeC6H4, Ph; 4-BrC6H4, 4-BrC6H4; iPr, Ph; nHex, Ph
R3 = Ph, 4-MeC6H4, 4-FC6H4, 4-ClC6H4, 4-MeSC6H4, 3-MeC6H4, 3-MeOC6H4, 2-MeC6H4,
2-napthyl, 3-thiophenyl,
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ph CF3
Me F3C CF3
N F3C CF3
H H N
N N
P Ir B
N N N
H H
N F3C CF3
Me
F3C CF3
Ph CF3 211
Ir(ppy)2(Me2phen) BARF
210

Ms R1
N [IrII]
211 N Me
H R3 R2
207
208

Ms photoredox
N catalysis [*IrIII]+BARF-
[IrIII]+
H R3 R1
white
IrIII•212 LED N Me
III + - R2 213•BARF
[Ir ] BARF
(211•BARF)
B

B•HBARF

NH
*
Ms
HN P NH N R1
210•BARF NH H R3 N CH2
* 210•212 R2
214
IBA
catalysis

R1 HN Ms
B: base, e.g. 207, 212, BARF-, etc 209
N
IBA: ionic Brønsted acid R2 R3

Scheme 35.Enantioselective -coupling of N-arylaminomethanes with imines.

An intramolecular hydroetherification of alkyl pent-4-enol with poor to moderate


enantioselectivity enabled by blue LED irradiation and a chiral ion pair acridinium photocatalyst was
reported by Luo et al., which afforded tetrahydrofuran (Scheme 36).60 In this study, the ion pair
Organic & Biomolecular Chemistry Page 32 of 83

32

catalyst demonstrated higher activity compared with its parent catalyst, 2 (Fukuzumi catalyst).
View Article Online
DOI: 10.1039/D0OB00759E

Cl

Organic & Biomolecular Chemistry Accepted Manuscript


MesAcr+217– (5 mol %) Mes
R1 R1 R2 R2
PhCH(CN)2 (100 mol %) R1 O
O O
HO R2 CH2Cl2, –25 °C R1 R2 P
blue LEDs N O O
215 216
Me
50-88% yield
2-64% ee MesAcr+ 217–
R1 = Ph, 4-ClPh, 2-MePh, 3-MePh, 4-MePh Cl
R2 = Me, -(CH2)4-, -(CH2)5-, -(CH2)6-, 4-ClPh, 4-FPh, 4-MeOPh

Scheme 36 Enantioselective intramolecular hydroetherification.


Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Jiang et al. reported an enantioselective substitution reaction of alkyl halides and -amino
radicals derived from N-aryl amino acids via photoredox catalysis (Scheme 37).61 Photosensitizer
dicyanopyrazine-derived chromophore (DPZ) 221 was used as the photoredox catalysis to generate
achiral alkyl radicals from racemic -bromoketone and N-phenyl glycines, and CPA catalyst SPINOL
222 was applied to control the stereoselectivity of the free radical coupling reaction between both
alkyl radicals.

DPZ (0.5 mol %)


O R3 222 (10 mol %) R3 NC N OMe
O S
Br Ar NaHCO3 (3.0 equiv)
R1 + HO2C N R1 NHAr S
H 3Å MS, DME NC N OMe
R2 R2
0 °C, 60 h
218 219 3 W blue LEDs 220
DPZ (221)
46-86 %yield
up to 96% ee t
Bu
R1 = Ph, 4-ClPh, 4-MePh, 4-MePh, 4-MeOPh, 4-OHPh, 4-BrPh, 3-ClPh, 3-MePh,
3-MeOPh, Me, tBu, 2-thienyl t
Bu
R2 = Me, Et, Bn, 2-MePhCH2, 3-MeOPhCH2, 4-ClPhCH2, 4-MeOPHCH2,
R3 = H, Me ; Ar = Ph, 4-ClPh, 4-MePh, 4-MeOPh, 3-ClPh, 3-MePh
O O
P OH
DPZ*
t
O Bu
LED
219

t
222 Bu
DPZ photocatalysis
R3
Ar
N
H
DPZ

218
O O R3
222
R1 R1 NHAr
R2 enantioselective acid catalysis
R2 220

Scheme 37 Enantioselective C(sp3)–C(sp3) coupling reaction.

Jiang et al. reported an enantioselective decarboxylative -aminoalkylation of -amino acid-


Page 33 of 83 Organic & Biomolecular Chemistry

33

derived redox-active esters (RAEs) to give isoquinolines using a SPINOL-CPA 226 andOnline
View Article
DOI: 10.1039/D0OB00759E
photosensitizer DPZ 221 under blue LED irradiation. This reaction was used to prepare a series of -
isoquinoline-substituted amines in good yields with enantioselectivity up to 93% ee (Scheme 38).62

Organic & Biomolecular Chemistry Accepted Manuscript


The same group also reported an enantioselective semipinacol rearrangement and aerobic oxidation
of 2-aryl-3-alkyl indoles, affording chiral 2,2-disubstituted indolin-3-ones, through synergistic
operation of SIINOL-CPA catalysis and visible-light photocatalysis (Scheme 39).63

O O R2 R1
DPZ (1 mol %) N
226 (15 mol %)
R1
N + N O HN CO2Et
R2
4 Å MS, DME, –10 °C CO2Et
N
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

223 O 3W blue LEDs, 60 h H O


224 225 O
P
R1 = H, Cl, Br, OMe 77-95% yield O OH
R2 = Ph, 4-FPh, 4-MePh, 4-iPrPh, 4-MeOPh, 4-AcOPh, 3-FPh, up to 93% ee 226
3-ClPh, 3-BrPh, 2-FPh, 2-MePh, 2-BrPh, iPr, iBu, Me, Bn,3-indolyl methyl

NC N OMe
S
S
NC N OMe

DPZ (221)

Scheme 38 Enantioselective decarboxylative -aminoalkylation.

NC N OMe
S
R3
O S
NC N OMe
DPZ (0.5 mol %)
229 (10 mol %) R1 R3
R1 R2 DPZ (221)
4 Å MS, tBuPh, 8 °C, air N R2
N H
H 3 W blue LED, 40 h
228 Ar
227
64-90% yields
up to 94% ee O O
R1 = H, Me, F P
R2 = Ph, 4-FPh, 4-ClPh, 4-iPrPh, 4-tBuPh, 3-MePh OH
O
R3 = Ph, 4-FPh, 3-FPh, 4-ClPh, 3-ClPh, 2-ClPh, 3-BrPh, 4-MePh, 3-MePh, 229
4-iPrPh, 4-tBuPh, 4-MeOPh, 3-MeOPh, 2-naphthyl, 2-thienyl, Cy,
Ar Ar = 4-pyrenyl

Scheme 39 Enantioselective semipinacol rearrangement and aerobic oxidation.

In 2018, Xia et al. reported the enantioselective cyclization of tryptamines with TEMPO under
visible-light irradiation (Scheme 40).64 Nitroxide TEMPO converted the indole to its free radical
through visible-light enabled hydrogen atom transfer (HAT) and NH activation, followed by imine
radical cyclization catalyzed by a CPA. Subsequent reaction with another TEMPO molecule afforded
octahydropyrrolo[2,3-b]pyrroles in high yields with high enantioselectivities. The total synthesis of
natural product ()-verrupyrroloindoline was achieved in five steps from Cbz-tryptamine using this
synthetic method.
Organic & Biomolecular Chemistry Page 34 of 83

34

View Article Online

NHR2 233 (10 mol %), DMAP (1.5 equiv)


i
Pr DOI: 10.1039/D0OB00759E
i
Pr
CyNCO (1.0 equiv), toluene, –15 °C N
N blue LEDs (max = 455 nm) O
R1 R3 +

Organic & Biomolecular Chemistry Accepted Manuscript


i
N O N Pr
H R1 O O
R2 P
231 (TEMPO) N R3
230 H OH
O i
(2.5 equiv) 232 Pr

R1 = Me, F, Cl, Br, OMe; R2 = Ac, CO2Me, Fmoc, Cbz, Alloc, Boc 63-94 % yield
R3 = H, Me up ro 98% ee
i i
233 Pr Pr

NHR2 O
N C
233 N
OH
R1 R3 cyclohexyl isocyanate
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

N 234 (CyNCO)
N
blue LEDs R2
O
N
231 NHR2 H

R1 R3 O
R1 R3 N
N P O
H H O *
O
230

233

N N
O O
N
R1 R2
N R3 232
H

Scheme 40 Enantioselective photocyclization.

In 2018, Jiang et al. reported an enantioselective decarboxylation and radical coupling reaction
of N-aryl glycines to give 1,2-diketones using CPA SPINOL and a photosensitizer DPZ under blue
LED irradiation (Scheme 41).65 N-arylglycine produced an -amino intermediate through a single-
electron oxidative decarboxylation reaction under photoredox catalysis, then underwent a coupling
reaction with an electrophilic -keto radical that was stereocontrolled by H-bonding with the CPA.
Similarly, the CPA SPINOL and DPZ method has been applied to the conjugate addition and
enantioselective protonation of N-aryl glycines to SPINOL -branched 2-vinylazaarenes for the
synthesis of 3-(2-pyridine/quinoline)-3-substituted amines (Scheme 42).66 In addition, cooperative
Brønsted acid catalysis and visible-light photocatalysis using a combination of photosensitizer DPZ
and CPA SPINOL has been applied in the enantioselective -deuteration of prochiral azaarene-
substituted ketones or racemic -chloroazaarenes with D2O as the deuterium source (Scheme 43)67
Similarly, reactions of azaarene-substituted ketones in the absence of D2O afforded the corresponding
chiral alcohols with good enantioselectivities.68 Furthermore, the strategy combining photosensitizer
DPZ with a chiral H-bonding catalyst (L-tert-leucine-based squaramide amine 255) has been used for
the enantioselective -dehalogenative protonation of chiral cyclic and acyclic ketones under visible-
light photocatalysis, with 2,3-dimethyl-1,2,3,4-tetrahydroquinoxaline (256) as the terminal reductant
(Scheme 44).69
Page 35 of 83 Organic & Biomolecular Chemistry

35

238 (10 mol %) O


View Article Online
R3 DPZ (1.5 mol %) R1 DOI: 10.1039/D0OB00759E
H OMe
N CO2H O (nBu)4PBr (30 mol %) NC N
N R2 S
R1 H
+ R2 Na2S2O4 (0.5 equiv) HO
NC N
S

Organic & Biomolecular Chemistry Accepted Manuscript


cyclopentyl methyl ether OMe
O 237
235 3 W blue LED R3
236 5Å MS, 10 °C, 36 h yield up to 89% DPZ (221)
up to 97% ee
Ar
R5
O O
H O 239 (20 mol %) HO P
N CO2H N OH
DPZ (1.0 mol %) H O
R4 R5 O
+ N THF, 10 °C
R4 O
SPINOL-CPA
N Ar
235 3 W blue LED, 36 h
Boc Boc 241 238: Ar = 2,6-Me2-4-tBuC6H2
240
yield up to 91% 239: Ar = 3,5-tBu2-4-MeOC6H2
up to 94% ee
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R1 = PH, Cl, Br, Me, MeO


R2 =Ph, iPr, 4-FPh, 4-ClPh, 4-BrPh, 4-tBuPh, 4-MePh, 3-MePh, 4-MeOPh, Naph
R3 = H, F, Cl, Br, tBu, Me, MeO
R4 = 2-Me-4-MeOPh; R5 = H, F, Cl, Br, Me, MeO

DPZ*
235
LED
H
N CO2H
Ph
DPZ photocatalysis
–H+
–CO2

H
N
Ph CH2
DPZ

236
O O
238 Ph
Ph N
R2 R2
enantioselective H-bonding H
O catalysis HO Ph
237

Scheme 41 Enantioselective radical coupling reaction.


R2 R2
H 245 (15 mol %) R3
N CO2H DPZ (0.2 mol %) NC N OMe
N S
R1 R3
+ N LiPF6 (40 mol %) R1 H
N S
THF, –35 °C 244 NC N OMe
242
243 3 W blue LED, 48-72 h yield up to 94%
up to 95% ee DPZ (221)

Ar
R2 R2
H 246 (15 mol %)
N CO2H N DPZ (0.2 mol %) N O O
R1 N P
+ LiPF6 (40 mol %) R1 H
O OH
242 THF, –35 °C 248
247 3 W blue LED, 48-72 h yield up to 85% Ar
up to >99% ee
SPINOL-CPA
R1 = H, F, Cl, Br, Me, CF3, MeO; R3 = H, Br, Me, MeO 245: Ar = 2-CF3Ph
R2 = Ph, 4-CNPh, 4-CF3Ph, 4-FPh, 4-ClPh, 4-BrPh, 4-MePh, 4-PhPh, 4-MeOPh, 3-FPh, 3-MePh, 246: Ar = 2-naphthyl
2-MePh, 2-napthyl, 1-naphthyl, 5-benzofuranyl, Me

Scheme 42 Conjugate addition and enantioselective protonation.


Organic & Biomolecular Chemistry Page 36 of 83

36

View Article Online


222 (20 mol %) DOI: 10.1039/D0OB00759E
DPZ (1 mol %) NC N OMe
R1 D S
R1 Cl 200 (1.5 equiv) R2
R2 N
N S

Organic & Biomolecular Chemistry Accepted Manuscript


D2O (150 equiv) NC N OMe
R3 NaHCO3 (1.0 equiv) R3
249 250
mesiylene, 25 °C 58-89% yield
(racemate) DPZ (221)
3 W blue LED up to 95% ee
Ar
245 (10 mol %)
O O EtO2C CO2Et
DPZ (0.5 mol %) D
R4 200 (1.5 equiv) R4 R5 P
R5 N O OH
N Me N Me
D2O (200 equiv) H
252 OH
251 O NaCl (0.2 equiv), –5 °C
Ar 200
(racemate) cyclopentyl methyl ether 65-99% yield
SPINOL-CPA
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

3 W blue LED, 14 h up to 99% ee


222: Ar = 2-tBuPh
245: Ar = 2-CF3Ph
R1 = H, Me, Br, MeO
R2 = Ph, 4-CF3Ph, 4-FPh, 4-MePh, 3-MePh, 3-ClPh, 3-MeOPh, 2-MePh, 4-MePh, 4-tBuPh, 4-PhPh, 4-vinylPh
1-naphthyl, 9-phenanthrenyl
R3 = iPr, Me, Et, Bn, allyl, (CH2)3C=CH2, 3-pentanyl, cyclopentyl, Cy, tetrahydro-2H-pyran-4-yl chloride, cycloheptanyl
R4 = H, F, Me; R5 = Ph, 4-CF3Ph, 4-FPh, 4-MePh, 4-BrPh, 3-MePh, 2-MePh, 2-thenyl, 2-furanyl, 2-naphthyl, (CH2)2CH=CH2, Me

Scheme 43 Enantioselective reductive deuteration and dehalogenative deuteration.


O O
DPZ (0.5 mol %) t
O Br Bu
O 255 (5 mol %) F
F N N
256 (0.6 equiv)
Cl R H H H
R Br N
n Na2CO3 (2.5 equiv) X n
X H2O (1.0 equiv) 255
DCE, 5 °C, 3 W blue LEDs 254
253 H
55-96% yield N
X = C, O, S up to >99% ee
n = 0, 1, 2 256
R = H, F, Cl, Br, Me, OMe, OTs, OAllyl, OH, OMs, OTBS N
H

DPZ* 256 H
LED N
+ H+
SET
N
photocatalysis H
DPZ

N
O O
SET N
t DPZ
Br Bu
N N
H H
Br O O
O +e O
F
F Br
t
Bu + H+
enantioselective H
Cl N N
protonation
H H
Br
O 254

acid catalysis F

Scheme 44 Enantioselective dehalogenative protonation of ketones.

In 2018, Phipps et al. report an enantioselective Minisci-type addition of N-acyl -amino alkyl
radicals, generated in situ from the corresponding amino acid derivatives, to pyridines and quinolines
by merging photocatalysis with Brønsted acid catalysis (Scheme 45).70 This method has
Page 37 of 83 Organic & Biomolecular Chemistry

37

demonstrated the highly regio- and stereoselective late-stage functionalization of twoOnline


View Article
DOI: 10.1039/D0OB00759E
pharmaceuticals, namely, metyrapone, for the diagnosis of adrenal insufficiency, and etofibrate, a
fibrate.

Organic & Biomolecular Chemistry Accepted Manuscript


O O R2 NHAc
(Ir(dF(CF3)ppy)2(dtbpy)]PF6 (1 mol %)
N
(R)-TRIP or (R)-TCYP (5 mol %) N
+ N O NHAc
1,4-dioxane
R2

R1 blue LEDs, 14 h
O 258 R1
257
259
R1 = Me, MeO, Ph, F, Cl, CO2Me, EtC=O, CN, CF3, 4-BrPhO
R2 = Bn, Me, iPr, sBu, (CH2)2Ph, (CH2)4NHBoc, (CH2)2CO2tBu, (CH2)2SMe, 4-I-PhCH2, up to 98% yield
4-AcOPhCH2, (1H-indol-3-yl)methanyl up to 97% ee
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R R
CF3 PF6
F
t
Bu
N
N R O
O
F Ir III P
F O
N R OH
N
t
Bu
F
CF3 R R
i
260, (R)-TRIP: R = Pr
[Ir(dF(CF3)ppy)2(dtbpy)]PF6
261, (R)-TCYP: R = Cy

Scheme 45 Enantioselective Minisci-type addition.

In 2019, Jiang et al. reported an enantioselective addition of alkyl or ketyl radicals to


vinylpyridines by merging asymmetric Brønsted acid catalysis with photoredox catalysis under
visible light irradiation to afford chiral -phenyl and aminopropylpyridines (Scheme 46).71 The chiral
Brønsted acid catalyst (SPINOL-CPA) activated pyridine and C=X, and controlled the
enantioselectivity. A series of -aminoalkyl and ketyl radicals generated in situ from the
corresponding ketones, aldehydes, and imines underwent conjugation addition to vinylpyridine with
good enantioselectivities.
Organic & Biomolecular Chemistry Page 38 of 83

38

DPZ 221 (0.5 mol %) R1 View Article Online


R1 DOI: 10.1039/D0OB00759E
SPINOL-CPA 233 (8 mol %)
X
+ Hantzch ester 265 (1.5-2.0 equiv)
N
Ar
N Ar R2 3 Å MS, CH3Cl, or CH3Cl–C6F5H (4:1),

Organic & Biomolecular Chemistry Accepted Manuscript


262 –5 °C,3W blue LEDs 264 R2 X
263
or or
DPZ 221 (0.5 mol %)
X N
SPINOL-CPA 268 (20 mol %)
N + Ar
Ar R2 Hantzch ester 200 (2.5 equiv)
263 toluene, –5 °C or –50 °C 267 R2 X
266
3W blue LEDs
up to 99% yield
X = O, N(4-EtOPh) up to >99% ee
R1 = H, Me, MeO; R2 = H, Me, Et, FCH2
Ar = Ph, 4-FPh, 4-ClPh, 4-BrPh, 4-CF4Ph, 4-MePh, 4-MeOPh, 4-CO2EtPh,
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

3-CF3Ph, 4-FPh, 3-FPh, 3-BrPh, 3-BrPh, 3-MePh, 3-MeOPh, 3-CNPh,


2-BrPh, 2-thienyl, 2-furanyl, cinnamyl

Ar1

O RO2C CO2R
OMe O
NC N
S P OH
S O Me N Me
NC N H
OMe
Ar1 265 R = tBu
200 R = Et
DPZ (221) SPINOL-CPA
233 Ar1 = 2,4,6-iPr3C6H2
268 Ar1 = 9-anthryl

Scheme 46 Enantioselective addition of radicals to vinylpyridine.

Wang, Yang, and coworkers reported the enantioselective reaction of 2-aminochalcone with
catalytic CPA under blue LED irradiation to afford tetrahydroquinolines (Scheme 47).72 In the
absence of photocatalyst, the process underwent a relay visible-light-enabled cyclization and chiral
Brønsted acid-catalyzed hydrogenation reaction.

O 271 (2 mol %) Ar
200 (3.0 equiv) R1
R2
N R2 O O
R1 EtOAc, rt P
H O OH
NH2 5W blue LEDs
269 270
R1 = H, Cl, Br, Ph 51-96% yield Ar 271
R2 = Ph, 4-MePh, 3-MePh, 2-MePh, 4-MeOPh, up to >99% ee
3-MeOPh, 4-PhPh, 4-FPh, 4-ClPh, 4-BrPh, Ar = 9-(10-phenyl)-anthryl
4-CF3Ph, 1-napthyl, 2-napthyl, 2-thienyl,
2-furyl, Me, Ph(CH2)2 EtO2C CO2Et

blue
Me N Me
LEDs H
200
cat. 271
R1
can be done
N R2 in the absence of light

Scheme 47. Enantioselective coupling of tetrahydroisoquinolines and alkynes.

Knowles et al. reported an intramolecular enantioselective hydroamination of alkenes with


Page 39 of 83 Organic & Biomolecular Chemistry

39

sulfonamides using a CPA and photocatalyst [Ir(dF(CF3)ppy)2(5,5’-dCF3bpy)]PF6 via N-centered View Article Online
DOI: 10.1039/D0OB00759E
radicals, formed in-situ from the proton-coupled electron transfer (PCET) activation of sulfonamide
N–H bonds (Scheme 48).73 Subsequent induction of the sulfonamidyl radicals by the CPA resulted

Organic & Biomolecular Chemistry Accepted Manuscript


in enantioselective C–N bond construction.
[Ir(dF(CF3)ppy)2(5,5'-CF3bpy)]PF6 R2 R2
R2 O O (2 mol %), 274 (2.5 mol %)
R1 O
S thiol 275 (30 mol %) O
R3 N S N
H CF3Ph, –20 °C, 72 h
272 blue LEDs R1
273
R1 = H, MeO, Me, Cl, Br, CN, CF3, etc.
up to 98%yield
R2 = R3 = Me, (CH2)5, (CH2)3, (CH2)2NBoc(CH2),
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

up to 96% ee
R2 = H, R3 = Me, tBu

Ph
N N
n
Oct N i
Pr
SH
O O Ph
P NBu4 i i
Pr Pr
O O Ph

n
275
Oct N
N N
274
Ph

Scheme 48. Enantioselective intramolecular hydroamination of alkenes.

7. Chiral photocatalysts

Chiral-only-at-metal asymmetric catalysts, including ruthenium or iridium complexes, can be


activated by visible light, allowing visible light to affect asymmetric catalysis.74 In 2014, Meggers et
al. reported a new strategy using chiral iridium photocatalyst –279 for enantioselective photoredox
catalysis (Scheme 49).75 The chiral metal center in the photocatalyst served as the sole source of
chirality, as well as providing catalytic Lewis acid activity. A conceivable mechanism for photoredox
catalysis combined with asymmetric catalysis was proposed, as shown in Scheme 49. Initial bidentate
coordination of 2-acyl imidazole 276 with catalyst –279 and subsequent deprotonation formed
iridium (III) enolate complex 281. Next, the enantioselective addition of alkyl radical 282, generated
from photoredox catalysis of alkylbromide 277, generated radical 283. Single-electron oxidation of
ketyl radical 283 to give the iridium-coordinated keto intermediate 284 by photocatalyst iridium (IV)
regenerated the photocatalyst PS [IrIII]. After exchange with starting substrate 276, product 278 was
released from complex 284, and a new asymmetric catalytic cycle proceeded. The structure of IrIII
enolate complex 281, which is the key intermediate between the asymmetric catalysis and
photocatalysis cycles, was confirmed by X-ray crystal analysis. The synthetic strategy of integrating
the asymmetric and photoredox catalytic cycles in a single photocatalyst provided a new approach to
enantioselective visible-light-induced photocatalysis. The same group later applied these catalysts to
Organic & Biomolecular Chemistry Page 40 of 83

40

the reactions of N,N-diaryl-N-(trimethylsilyl) methylamines and 2-acyl-1-phenylimidazoles under airOnline


View Article
DOI: 10.1039/D0OB00759E
and household-lamp irradiation to provide oxidative coupling products in good yields (61%–93%)
with high enantioselectivities (90%–98% ee) (Scheme 50).76 This catalyst had three unique functions,

Organic & Biomolecular Chemistry Accepted Manuscript


namely, as an active Lewis acid, a photoredox sensitizer, and as the sole source of chirality.
This chiral photocatalyst and strategy has since been successfully applied to the enantioselective
trichloromethylation of 2-acyl imidazoles and 2-acylpyridines with excellent enantioselectivities of
up to >99% ee (Scheme 50).77
O O
N R2 N
Br R3 –279 (2 mol %) R3
N + Na2HPO4 (1.1 equiv), 40 °C N R2
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R1 276 277 R1
14W white light energy-saving lamp 278

R1 = Me, iPr t
R2 = Me, Et, Ph, C6H4Me, C6H4OMe, C6H4Cl, 2-naphthyl, 3-thiophenyl Bu
S
R3 =
X N Me PF6
F O C
N
Ir –279
Y N
NO2 NO2 N C
Me
X = CN, NO2, CO2Et Y = NO2, Cl, Br S t
Bu

[IrIII]
[IrIII] CH2R3
N O
N O 282
N Br R3
R2
N R2
R1 PS [Ir ]IV +e 277
R1 281 R3
H+ 283

[IrIII]
N O 280 asymmetric
analysis
e
photocatalysis PS [IrIII]*
N R2
R1

[IrIII]
O
–279 N O PS [IrIII]
N R2
N R2 [IrIII]
N III
R1 276 R1 PS [Ir ] = –279 N O
O R3 or
284 N R2
N
R3 R1
N R2
R1 279

Scheme 49 Enantioselective -alkylation of 2-acyl imidazoles.


Page 41 of 83 Organic & Biomolecular Chemistry

41

View Article Online


O DOI: 10.1039/D0OB00759E
O t
R4 Bu PF6
N O
N R2 R4 -288 (2 mol %) N
Me3Si N Me
+

Organic & Biomolecular Chemistry Accepted Manuscript


H2Cl2, rt, air N R2 R3 N C
N R3 12-W CFL R1 N
R1 287 Ir
285 286
61-93 % yield N
N C
up to 98% ee Me
i
R1 = Ph, Me, Pr,
R2 = Ph, 2-MeC6H4, 3-MeC6H4, 4-MeOC6H4, 4-ClC6H4,3-thiophenyl, Me, Et O t
Bu
R3 = Ph, 4-MeOC6H4, 4-MeC6H4, 2-naphthyl; R4 = Ph, 4-MeC6H4
O -288
O
-279 (2-4 mol %) N R
N R
+ BrCCl3
NaHCO3, MeOH–THF N CCl3 S
t
Bu PF6
N
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ph 290 rt, 20-W CFL R1 291


289 N Me
62-96 % yield C
N
up to >99.9% ee Ir
N
R = Ph, 2-MeC6H4, 3-MeC6H4,4-MeOC6H4,4-ClC6H4, 2-ClC6H4, 2-naphthyl, C
N Me
3-thiophenyl, Me, Et, nPent
O S t
Bu
O
N R2
N R2 -279 (2-4 mol %)
+ BrCCl3
2,6-lutidine CCl3
-279
293 MeOH–CH3CN R1
R1 292 294
40 °C, 20-W CFL
65-91 % yield
R1 = H, Me, Cl, Br, Ph, nPr; R2 = Me, Ph up to 99.6% ee

Scheme 50 Enantioselective coupling reactions.

Meggers, Gong, and coworkers later applied a similar chiral rhodium (III) photoredox catalyst
in the aerobic enantioselective visible-light-driven dehydrogenative cross-coupling of 2-acyl
imidazole 295 with arylamine 296, providing dual functions as a Lewis acid and visible-light
photocatalyst (Scheme 51).78 In most cases, the reaction yields were >70% with >90 % ee, with a
few examples of 99% ee. A proposed mechanism was provided by the authors in which the enolate
rhodium complex approached the iminium species, generated in situ from the photosensitized
oxidation or autoxidation of amine 296, from the Re face. Furthermore, the chiral photocatalyst has
been applied to the visible-light-activated enantioselective addition of alkyl radicals to ,-
unsaturated acyl imidazole (Scheme 52).79
Organic & Biomolecular Chemistry Page 42 of 83

42

View Article Online


O R3 O R5 DOI: 10.1039/D0OB00759E
-288 (2-5 mol %)
R2 N R3
R1 R1 N
+ R5 R4 air, 24 W blue LED

Organic & Biomolecular Chemistry Accepted Manuscript


295 R2 R4
296 297
35-85% yield; 79-99% ee
R 1: N N For R5 = CO2tBu, dr: 5:1 - 12:1

N most cases: >70% yield; >90% ee


N
i
R R = Ph, H, i-Pr Pr

R2 = Ph, 4-MeOC6H4, 4-BrC6H4, 3-thiophenyl, 2-naphthyl


R3 = Me, Ph, p-Tol, o-Tol, 3,5-Me2C6H3, 4-MeOC6H4, 4-Cl, C6H4
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R4 = H, Me, Et, iPr


R5 = CO2tBu, H
Best condition for R5 = CO2tBu: -288 (2.0 mole %), TFA (2.0 mole %), O2,
NMP, 40 °C

O
Si R5
N
N R R3
N N
Ir
O R2 R4
N
O

Scheme 51 Enantioselective domino photo-oxidation/Mannich reaction.


R3
O
O
-179 (4 mol %) N
N - +
R2
R2 + R3 BF3 K seseitizer 301 (2 mol %), rt N
N acetone-H2O, blue LEDs, 4-24 h R1 300
R1 298 299
40-97 % yield
R1 = Ph, Me, iPr, up to 99% ee
R2 = Me, Et, nPr, iPr, CH2OBn, Ph, 2-BrC6H4, 4-FC6H4, 4-MeOC6H4
R3 = 4-MeOC6H4CH2, 4-MeC6H4CH2, 4-ClC6H4CH2, tBuOCH2, MeOCH2,
4-MeOC6H4CH2OCH2, TMSCH2OCH2, iPr, tBu, cyclpentyl, cyclohexyl,

t
Bu PF6
S

N Me
C
N
Rh
N
N C
Me
N
S t
Bu ClO4
301
-179
Scheme 52 Enantioselective addition of alkyl or aminyl radicals to alkenes.
Page 43 of 83 Organic & Biomolecular Chemistry

43

Meggers et al. reported an enantioselective synthesis of 1,2-amino alcohols from tertiaryView


amines
Article Online
DOI: 10.1039/D0OB00759E
and trifluoromethyl ketones using a chiral iridium complex as a dual Lewis acid/photoredox catalyst
(Scheme 53a).80 This visible-light-driven process incorporated single-electron transfer between a

Organic & Biomolecular Chemistry Accepted Manuscript


donor substrate and catalyst-activated acceptor substance with a stereocontrolled radical–radical
cross-coupling, affording excellent enantioselectivity of up to 99% ee. This protocol was also
implemented in the coupling reaction of -silylamines with 2-acyl imidazoles, followed by
desilylation, to prepare 1,2-amino-alcohols in 69%–88% yields with high enantioselectivity (up to
99% ee) (Scheme 53b).81 Furthermore, these catalysts have also been applied to the enantioselective
radical amination of 2-acyl imidazoles using 2,4-dinitrophenylsulfonyloxy carbamates82 (Scheme
53c) and the radical perfluoroalkylation of 2-acyl imidazoles using perfluorobenzyl iodide and
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

perfluoroalkyl iodides (Scheme 53d).83 Additionally, the enantioselective -fluorination of 2-acyl


imidazoles using this iridium catalyst was reported by Xu et al. (Scheme 53e).84 However, these
reactions did not indicate whether the compounds were obtained by visible-light irradiation, or
whether they could still be obtained in the dark. However, according to the literature, these reactions
are likely promoted by indoor compact fluorescent or LED lighting. Furthermore, Kang et al. reported
an enantioselective conjugate addition of ,-unsaturated 2-acyl imidazole with
tetrahydroisoquinolines, activated by visible-light irradiation and chiral rhodium catalyst -320
(Scheme 53f).85
Organic & Biomolecular Chemistry Page 44 of 83

44

View Article Online


DOI: 10.1039/D0OB00759E
R1
(a) t
R1 Bu PF6
S

Organic & Biomolecular Chemistry Accepted Manuscript


-279 Me
O N C
(3 mol %) HO CF3 N
N Ir
CF3 + CHCl3 N N
N
N N C
Me CFL or N Me
Ph N
blue LEDs Ph R2
R2 304 S t
Bu
302 303
60-82% yield -279
R1 = H, Me, tBu, Cl; R2 = H, Me, OMe, tBu, Cl up to 99% ee

(b)
(1) -179 (4 mol %) R1 t
Bu PF6
O HO R2 S
R3 Ru(bpy)3(PF6)2 (1 mol %)
N N N Me
R2 CH3CN–DMAC, CFL R2 N C
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

+ TMS N
N R4 N N
(2) TBAF 307 Rh
R1 R1
306 N
305 69-88% yield N C
Me
up to 99% ee
R1 = Me, Ph, o-Tol, (2-Ph)Ph S t
Bu
R2 = Me, Et, Ph, Tol, 2-MePh, 4-ClPh, 4-AcOPh, 4-CO2MePh, 4-MeOPh, AcOPh, 4-PhPh -179
R3 = Me, Bn, tol, 4-tBuPh; R4 = tol, 4-tBuPh, 3-ClPh
(c)
O R3
O O Me -311 O PF6 t
Bu
N R2 S N (2 mol %) N O
N
+ O CO2Me CO2Me Me N
N 2,6-lutidine, rt N C
R2 N
R1 O2N NO2 309 CH3CN–DMSO, R1 310 Rh
308 blue LEDs N
52-99% yield C N
Me
up to 98% ee
R1 = Ph, o-Tol, iPr t O
R2 = Ph, 4-FPh, 4-ClPh, 3-ClPh, 2-ClPh, 4-BrPh, 4-MePh, 4-MeOPh, 2-naphthyl Bu
-311
R3 = Me, nBu, iBu

(d)
O t PF6
O Bu
R3 S
-279 N
N R2 (2-4 mol %) Me
N
+ R3 I N R2 314 N
C
N NaHCO3, MeOH–THF R1 Ir
R1 312 313 rt, 21-W CFL N
24-93% yield N C
absolute config. vary Me
up to >99.5% ee
R1 = Ph, o-Tol S t
R2 = Me, Et, nBu, Ph, 4-MeOPh, 4-MeOPhO, 4-ClPh, 2-ClPh Bu
-279
R3 = CF3, C3F7, C4F9, C6F13, C8F17, C10F21, C6F5CF2

(e)
O
O PF6 t
Bu
N R2 S
N R2 -279 (4 mol)
N Me N
F C
N selectfluor (1.2 equiv) R1 316 N
R1 315 MeOH, rt Ir
60-97% yield N
up to 99% ee C N
Me
R1 = Me, iPr, Ph
R2 = Ph, 4-PhPh, 4-ClPh, 4-FPh, 4-MeOPh, 2-ClPh, 2-MePh, 3-MePh, 3,4-diMeOPh, t
Bu S
2-naphthyl, Me, Et, nBu, nPent -279

(f) O R1 R3
O
-320 N N t PF6
N Ar Bu
R1 (2 mol %) O
R2
N N
N + R3 CH2Cl2, rt R1 N C
Me
i
Pr blue LEDs R2 N
318 319 Rh
317 N
54-96% yield N C
Me
90-99% ee
Ar O t
Bu
R1 = Ph, 4-MePh, 4-ClPh, 4-BrPh, 4-NO2Ph, 4-CF3Ph, 3-MeOPh, 2-MePh, 2-furyl,
2-thienyl, 1-naphthyl, 2-naphthyl, Me, CO2Et
Ar = 3,5-(CF3)2C6H3 -320
R2 = 6-Cl, 6,7-(MeO)2; R3 = Ph, 4-MePh, 4-BrPh, 3-BrPh, 4-CO2Me, 2-MeOPh

Scheme 53 Coupling reactions of 2-acyl imidazoles with different substrates.


Page 45 of 83 Organic & Biomolecular Chemistry

45

Meggers et al. subsequently reported a proton-coupled electron transfer (PCET)86 strategy forOnline
View Article
DOI: 10.1039/D0OB00759E
the photocatalytic enantioselective -amination of ,-unsaturated 2-acyl imidazole with N-phenyl
carbamic acid methyl ester using a combination of a weak phosphate base (tetrabutylammonium

Organic & Biomolecular Chemistry Accepted Manuscript


diphenyl phosphate), iridium photocatalyst, and chiral rhodium Lewis acid -311 (Scheme 54).87
This process relied on a visible-light-induced and phosphate-facilitated single-electron transfer from
the NH group to a nitrogen-centered free radical, and the reaction of a nitrogen nucleophile with a
catalyst-bound substituted olefin, accompanied by a highly enantioselective radical–radical cross-
coupling, driven by a chiral rhodium–enolate radical intermediate. Meggers and coworkers reported
a visible-light-activated enantioselective reaction of ,-unsaturated N-acylpyrazole with allyl
sulfone using chiral rhodium catalyst -311, which provided the radical allylation product and
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

sulfonyl radical addition product in one process (Scheme 55).88 This method used the Hantzsch ester
328 and bidentate coordination of N-acylpyrazole with rhodium catalyst -311 under visible-light
irradiation to generate a radical species that reacted with allyl sulfone 325 and delivered the sulfonyl
radical and an enolate intermediate. The enolate afforded product 326 after protonation, with
regeneration of catalyst -311. Concurrently, in a sequence involving a stereodirected radical addition,
hydrogen atom transfer, and ligand exchange, the sulfonyl radical reacted with the rhodium-
coordinated N-acylpyrazole to form sulfonation product 327. In contrast, in the absence of LED
irradiation, Kang and coworkers have reported that the conjugate addition of N-protected
hydroxylamines to ,-unsaturated 2-acylimidazoles can be catalyzed by these types of chiral Rh
(III) catalysts to afford the adducts in good yields with good enantioselectivities.89 Furthermore,
Kang et al. also reported that, without intentional light irradiation, these types of chiral Rh(III)
complex can catalyze the decarboxylative conjugate addition of -keto acids to ,-unsaturated 2-
acyl pyridines or imidazoles with catalytic loadings as low as 0.05 mol%, providing the products in
excellent yields (94%–98%) with high enantioselectivities (88%–96%).90 More recently, Meggers et
al. reported an enantioselective -alkylation of ,-unsaturated 2-acyl imidazoles using a series of
Hantzsch esters as alkyl radical sources under the typical conditions (Scheme 56).91

O
PF6 t
Bu
R3 O
O -311 (5 mol %) O N OR2
O [Ir(dF(CF3)ppy)2(5,5'-dCF3bpy)]PF6 Me
N N C N
R1 R3 (2 mol %), Ph2PO4-Bu4N+ (20 mol %) R1
N OR2 N
N + H
CH2Cl2, blue LEDs, rt
N
N
Rh
Ph 321 322 Ph 323 C
Me N
up to 97% yield
R1 = Ph, 4-CF3Ph, 4-ClPh, 4-BrPh, 4-FPh, 4-MeOPh, 4-MePh, 3-MePh up to >99% ee t
Bu O
R2 = Me, Et, iPr, tBu, iBu, Bn, (CH2)2Ph, Ph -311
R3 = Ph, 4-ClPh, 4-Br, 4-MePh, 3-Me, 2-Me

Scheme 54 Enantioselective catalysis of 2-acyl imidazoles with carbamates.


Organic & Biomolecular Chemistry Page 46 of 83

46

EtO2C CO2Et View Article Online


328 O SODOI:
2Ar 10.1039/D0OB00759E
O O R2
R2 Et N Et (1.5 equiv) N
N H N R1
R1 R1
N SO2Ar -311 (8 mol %) N
+ N + 327

Organic & Biomolecular Chemistry Accepted Manuscript


326
1,4-dioxane, N2, white LEDs, rt
324 325 up to 92% yield
(2 equiv) (1 equiv) up to 85 % yield up to 89% ee
up to 97% ee
R1 = Me, Et, nPr, (CH2)2CH=CH2; R2 = CN, CO2Et, etc.
Ar = Ph, 4-MePh, 4-BrPh, 4-CF3Ph, 4-MeOPh, 2-MePh, 2,4,6-Me3Ph, 2-naphthyl, 1-naphthyl

N
DMP =
N

SO2Ar
ArSO2
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

[Rh] O R2
*
DMP R1 [Rh] O R2
R2
*
SO2Ar DMP R1
EtO2C CO2Et
325
Et N Et
[Rh] O H 328

DMP R1
EtO2C CO2Et
O R2
+ *
Et N Et DMP R1
SET
328 326
-311
photocatalysis

+ O
[Rh]
1
328 DMP R1

ArSO2

O SO2Ar

DMP * R1
327 + SO2Ar
[Rh] O

DMP * R1

EtO2C CO2Et
O + SO2Ar
[Rh] O Et N Et
DMP R1
324
DMP * R1
EtO2C CO2Et

Et N Et

Scheme 55 Visible-light-activated enantioselective radical allylation and sulfonylation.

t PF6
R2 Bu
O R2 S
t
O
N BuO2C CO2tBu N Me
R1 -179 (4 mol %) N C
R1 N
N + N CH2Cl2, 21W CFL N Rh
Ph N
H 13-18 h Ph 331 C
329 N Me
330
up to 96 %yield S t
R1 = Me, Ph, 4-MeOPh, 4-FPh, 2-BrPh, 4-Br-2-AcOPh, nPr, iPr, up to 98% ee Bu
R2 = Ph, 4-MeOPh, 4-FPh, 2-BrPh, 2-thienyl, BnO, BocNH, iPr -179

Scheme 56 Enantioselective -alkylation of 2‑acyl imidazoles using Hantzsch esters.


Page 47 of 83 Organic & Biomolecular Chemistry

47

Meggers, Wiest, and coworkers employed chiral photocatalyst -RhS2 in DOI: a visible-light-
View Article Online
10.1039/D0OB00759E
induced enantioselective -alkylation of 2-acyl imidazoles and 2-acylpyridines with -ketoesters
(Scheme 57).92 Based on the results of experimental studies and theoretical calculations, the authors

Organic & Biomolecular Chemistry Accepted Manuscript


presented a mechanism in which a photoactivated Rh-enolate, after light excitation, transfers a single
electron to the -ketoester, followed by -proton transfer, and a subsequent stereodirected carbon
radical–radical coupling reaction.

t
O R1 Bu Ph
O -336 (4 mol %) O O S
PF6
N R2 DABCO (20 mol %) Me
R1 OR3 X OR3 C N
N
+ acetone, blue LEDs, rt N
O R2 OH Rh
332
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ph 334 335 N
C N
or X = 2-imidazole or 2-pyridine Me
O up to 97% yield t S Ph
up to >99% ee Bu
N
R1 -336
333
R1 = Ph, 4-MePh, 4-tBuPh, 4-MeOPh, 4-TBSOPh, 4-HOPh, 4-MeSPh, 4-CF3Ph, 3-BrPh, 3-indole-Boc
R2 = Ph, 4-MePh, 3-MePh, 4-FPh, 3-MeOPh; R3 = Cy, damantyl

O
N [Rh]+ O
R1 cat. [Rh]+
N
332 R1 –H+

[Rh]+ O
O R1 O N R1

X OR3
R2 OH
*
335 [Rh] O H
N
 R1
[Rh] O R1 O SET
N
OR3
R2 OH O
OH R2
R 3O
R 3O
R2 334 O
O
+H+
–H+
transfer

[Rh] O R1 O [Rh] O
N N
OR3 R1
R2 OH

Scheme 57 Enantioselective -alkylation of 2-acylimidazoles and 2-acylpyridines with -ketoesters.

Furthermore, in 2015, Meggers et al. reported the enantioselective conjugate addition of aryl
groups to 2-acyl imidazole derivatives using iridium catalysts bearing octahedral chiral ligands for
chiral Lewis acids catalysis in the absence of light irradiation.93 Later, the authors also applied this
catalyst to enantioselective hydrogenation and sequential photocatalytic radical
trifluoromethylation/cyclization (or atom transfer radical addition) reactions using the iridium
Organic & Biomolecular Chemistry Page 48 of 83

48

catalyst (Scheme 58).94 Recently, this process has been improved to a one-pot operation View
in other
Article Online
DOI: 10.1039/D0OB00759E
related reactions (Scheme 59).95

Organic & Biomolecular Chemistry Accepted Manuscript


(1) -279 (2 mol %) OH
O dmp (40 mol %) t
Bu PF6
H2 (50 bar), THF, rt S
R
N Me
R (2) 339 (1.5 equiv) C
NaHCO3 (2.0 equiv)
N
337 338 CF3 Ir
MeOH, 7 °C N
R = 4-Ph, 4-OMe, 4-CF3 24 W blue LEDs 54-76% yield N C
Me
dr 1:1; 94-96% ee
S t
Bu
(1) -279 (2 mol %) HO
O
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R3 dmp (40 mol %) R3 -279


H2 (50 bar), THF, rt
R1 CF3 O
R1 (2) 2 (1.5 equiv)
N N
2,6-lutidine (2.0 equiv)
R2 O
R2 CH3CN, rt 341 I N NH
340 24 W blue LEDs
52-72% yield CF3 dmp
R1 = H, Br, R2 = Boc, CO2Me; R3 = Me, nPr 93-99% ee
339

(1) -279 (2 mol %) OH CCl3


O dmp (40 mol %)
H2 (50 bar), THF, rt R
Br
R (2) BrCCl3 (2.0 equiv) 343
342 NaHCO3 (2.0 equiv) 70-90% yield
MeOH, 7 °C dr 1:1; 97-99 % ee
R = 4-Ph, 4-OMe, 4-CF3 24 W blue LEDs

Scheme 58 Sequential hydrogenation–trifluoromethylation.


Page 49 of 83 Organic & Biomolecular Chemistry

49

View Article Online


(1) -279 (2 mol %) DOI: 10.1039/D0OB00759E
dmp (40 mol %)
O OMe CH3CN, rt OH

Organic & Biomolecular Chemistry Accepted Manuscript


21 W blue LEDs
R
Br + OTBS R CO2Me
(2) HCO2NH4
344 345 THF–H2O (1:1), 40 °C 346
71-85% yield
R = 4-BrPh, 4-ClPh, 3-NO2Ph
93-95 % ee
(1) -279 (2 mol %)
dmp (40 mol %)
O
2,6-lutidine, CH3CN, rt OH
R1
21 W blue LEDs
+ O R1
CF3
OTBS I (2) HCO2NH4
349
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

THF–H2O (1:1), 60 °C
347 348 CF3
76-81% yield
R1 = 4-PhPh, 4-MeOPh, 4-BrPh, 3-benzofuranyl 91-94 % ee

OH
(1) -279 (2 mol %)
dmp (40 mol %) R
O TFA (1.0 equiv)
N
CH3CN, rt
R2 + 6 W blue LEDs
Me
N 352
350 Me 351 (2) HCO2NH4
THF–H2O (1:1), 40 °C 71-79 % yield
R2 = Ph, 4-ClPh, 4-ClPh 98-99 % ee

(1) -279 (4 mol %) Br


O O O
R3 Na2HPO4 (1.1 equiv)
N THF–MeOH (1:1), 40 °C N
N + Br
21 W blue LEDs, 24 h
N R3 OH
Me Br Me
(2) HCO2NH4 (9 equiv) 355
353 354 dmp (80 mol %) 80%yield; dr 4.5:1
R3 = 2-MePh THF–H2O (1:1), 60 °C, 5 h 96% ee

Scheme 59 Sequential photoredox and hydrogenation transfer reactions.

Megger et al. reported an enantioselective intermolecular [2+2] photocycloaddition of 2-acyl


imidazoles with alkenes by direct visible-light excitation of the Lewis acid catalyst/substrate complex
(Scheme 60).96 This process provided cyclobutanes with high yields, diastereoselectivity, and
enantioselectivity.
Organic & Biomolecular Chemistry Page 50 of 83

50

View Article Online


DOI: 10.1039/D0OB00759E

O R2 PF6 t
Bu
O R3 R5 -179 S
R1

Organic & Biomolecular Chemistry Accepted Manuscript


(2-4 mol %) R3 Me
R1 R2
+ R4 C N
acetone, rt R5 N
356 357 blue LEDs Rh
358 R4 N
C N
up to 99% yield Me
N N N dr up to >20:1
R1 = Me t
Bu S
up to 99% ee
N N
Ph Me -179

R2 = Ph, 4-FPh, 4-CF3Ph, 4-ClPh, 4-BrPh, 4-MePh, 4-MeOPh, 3-BrPh, 2-BrPh, 3,5-(CF3)2Ph,
2-furanyl, 3-furanyl, cinnamyl,
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R3 = H, Me; R4 = n-BuO, Ph, 2-thienyl, ; R5 = H, Me

[Rh]+ O R3
358
N
R2 blue LEDs

356
[Rh]+ O R3 *
N
R2
[Rh]+ O
R2
N excited siglet
R3
R5
ISC
R4

cyclization
[Rh]+ O R3 *
N
R2
ISC
[Rh]+ O R3 excited triplet

N
R2
R5 357

R4

Scheme 60 Enantioselective [2+2] cycloaddition.

In 2014, Bach and coworkers introduced chiral thioxanthone 361 as an organic photocatalyst
under visible-light irradiation for the enantioselective intramolecular [2+2] photocycloaddition of 359,
affording corresponding cyclobutane adduct 360 (Scheme 61). At a higher concentration of 2.5 nM,
the photoreaction of 359a gave 360a in 86% yield with 92% ee, while at a dilute concentration of
360a (1.0 mM), a 99% yield and slightly lower enantioselectivity (89% ee) were observed.97 A series
of quinolones, 359, were screened with catalyst 361 under the optimal reaction conditions, affording
promising results of 79%–95% yields with 87%–94% ee (Scheme 61).
Page 51 of 83 Organic & Biomolecular Chemistry

51

View Article Online


O DOI: 10.1039/D0OB00759E
N

Organic & Biomolecular Chemistry Accepted Manuscript


O S H
NH
361
O (10 mol %)
H
N O PhCF3, – 25 °C, visible light N O
H H
2.5 nM (359a), 30 min, 86% yield, 92% ee
359a 1.0 nM (359a), 30 min, 99% yield, 89% ee 360a

R1
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

X R2
X = O, NTs, C 79-95 % yield
R2
R1 = H, Me 87-94% ee
H R2 = H, Me
N O
H 360

N
O
N
O *
O S O
visible light S
N H O R N H O R
O H N O H N

361*•359a
361•359a R = (CH2)3CH=CH2

360a triplet-energy
transfer
359a
O
O N
N
O S
O S
N H O
N H O
*
O H N
O H N
361•359a*
361•360a

Scheme 61 Enantioselective intramolecular [2+2] photocycloaddition.

Later, Bach et al. applied the yellow chiral thioxanthone catalyst to intermolecular [2+2]
photocycloadditions of 2(1H)-quinolones to electron-deficient olefins, such as acrylates, to afford
products with good regio-, diastereo-, and enantioselectivities (Scheme 62).98 The products were
found to be head-to-tail (HT) adducts with the electron-withdrawing substituent at the exo-position.
The reactions proceeded under a Rayonet RPR-4190A light source (maximum wavelength, 419 nm)
in Rayonet-type photochemical reactors. As shown in Scheme 62, two-point hydrogen bonding
between the catalyst and quinolone was crucial for reaction success, providing adequate triplet energy
transfer and high enantioface differentiation. Furthermore, the cycloaddition process was also feasible
Organic & Biomolecular Chemistry Page 52 of 83

52

under solar radiation. Alternatively, this strategy was applied to the enantioselective photocatalysis
View Article Online
DOI: 10.1039/D0OB00759E
of the intermolecular [2+2] cycloaddition between acetylenedicarboxylates and 2-pyridones under
near-UV irradiation with fluorescence lamps (max = 366 nm) (Scheme 63).99

Organic & Biomolecular Chemistry Accepted Manuscript


O
N R3
Y R2
R1
R3 O R1
S R2
R2 NH
X + Y 361 X H
N O R2 (10 mol %) N O
H O H
362 363 364
PhCF3, – 25 °C
R1 = H, C5H11 violet light ( = 419 nm) 47-94% yield
R2 = H, Me X = H, Me, Br 80-95% ee
R3 = H, Me Y = CO2Me, CO2Et, CO2Bn, COMe, COEt, CHO
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

N
O *
O S
violet light
N H O R1
O H N 362•361*
O
N
O S
N H O 362*•361
R1 O
O H N N

362•361 O S
photocatalysis
N H O R1
R2
R2
R3 O H N *
H
O Y
O
R1
HN N R3
R2
O S
362 Y
R2 H R R2
364 N H O 1
Y
363
O H N
R3
R2
365•361

Scheme 62 Enantioselective intermolecular [2 + 2] photocycloaddition.


O
N

O H
O R 3O 2C R2
R2 CO2R3
NH
369
+ (10 mol %) R 3O 2C N O
R1 N O O R1 H
H CO2R3
366 Hexafluoro-m-xylene/trifluorotoluene = 2:1 368
367 – 65 °C, near UV light ( = 366 nm), 4 h 40-88% yield
59-92% ee
R1 = H, Me; R2 = H, Me; R3 = Me, Et, iPr, tBu
Scheme 63 Enantioselective [2+2] cycloaddition of acetylenedicarboxylates and 2-pyridones.

Sibi, Sivaguru, and coworkers reported an asymmetric intramolecular [2+2] photocycloaddition


of coumarin with atropisomeric thiourea catalysts (Scheme 64).100 The reactions provided the
corresponding cycloadduct at a low catalyst loading with high conversion and good enantioselectivity
Page 53 of 83 Organic & Biomolecular Chemistry

53

(77%–96% ee). The binaphthyl-based thiourea exhibited lower excited-state energies than theView starting
Article Online
DOI: 10.1039/D0OB00759E
coumarin substrate. In this reaction, the observation and manipulation of excited-state reactivity has
been accomplished by creating dynamic and static complexes (exciplex formation).

Organic & Biomolecular Chemistry Accepted Manuscript


H

R1 (S)-372 (1-10 mole %) R1


H
toluene/m-xylene (1:1), –78 °C R2 O O
R2 O O
370 371
coversion up to 100%,
R1 = H, Me, F, OMe; R2 = H, Me, F
isolted yield 77%
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

77-96% ee
R3
S
Ar
N N
H H
OR4

R3 (S)-372

R3 = H, CF3; R4 = H, Me; Ar = 3 5-(CF3)2C6H3, C6F5

Scheme 64. Enantioselective intramolecular [2+2] photocycloaddition of coumarin.

Yoon and coworkers reported an enantioselective intramolecular [2+2] cycloaddition using an


effective chiral photocatalyst ()-[Ir(tF(CF3)ppy)2(pypz)][BArF] under visible light irradiation
(Scheme 65). This process directed the quinolones using a series of hydrogen-bonding and –
interactions.101
R4 R3
R2 R4
R3 H
O R2 375•BARF (1 mol %)
R1 O
R1 CH2Cl2/pentane (1:1) N O
N O –78 °C, blue LED, 24 h H
H
373 374
up to >98% yield
R1 = F, Cl, Br, I, CF3, Me, MeO
up to 91% ee
R2 = H, Me, R3 = H, Me, Et; R4 = H, Me

CF3

F F3C CF3
N
F3C CF3
F N
F
Ir B
F
F N
HN F3C CF3
N
F F3C CF3

375 CF3 BARF

Scheme 65 Enantioselective [2 + 2] photocycloaddition.


Organic & Biomolecular Chemistry Page 54 of 83

54

Nicewicz et al. reported a moderate enantioselective radical cation Diels–Alder reaction Viewunder
Article Online
DOI: 10.1039/D0OB00759E
visible-light photocatalysis using photosensitizer 1c, comprising an oxidizing pyrilium salt, that
contains a chiral N-triflyl phosphoramide anion (Scheme 66).102 Unfortunately, some of the

Organic & Biomolecular Chemistry Accepted Manuscript


examples afforded low yields and enantioselectivities.
H
R R OMe
X SiPh3
378 (3.5 mol %) X
LEDs (470 nm) O 378
H O
toluene, –20 °C, 72 h P
O NTf
OMe
SiPh3 O
377 OMe
376
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

up to 72% yield MeO OMe


X = O, NTs; R = iPr, Cy, Bn, Me up to 50% ee

R 378 (3.5 mol %) H


H
+ LEDs (470 nm)
MeO 381
toluene, 60 °C
379 380 38-85% yield; dr 8-10:1; 32-36% ee
R = Me, Et OMe

Scheme 66 Enantioselective cation radical Diels–Alder reaction.

Meng et al. reported the enantioselective aerobic oxidation of -ketoesters, -ketoamides, and
1,3-diketones using a phase transfer catalyst containing a photosensitizer grafted to a cinchona
derivative (Scheme 67).103 Under white LED irradiation and air, the photocatalytic reactions resulted
in the corresponding -hydroxylation of -ketoesters in yields of up to 97% and enantioselectivities
of up to 86% ee. Subsequently, the enantioselective aerobic -hydroxylation of -dicarbonyl
compounds has been improved using cinchona-based phase-transfer catalysts and can be performed
in semi-flow photoredox processes.104
Ph

Ph NH
N
O O N
O OH O
384 (5 mol %) HN
R1 R1
Cs2CO3 (2 equiv)
n R2 toluene, rt, air n R2 Ph
382 25W white LEDs
383
n = 1, 2 N
Br
R1 = H, Cl, F, Br, Me, MeO,
R2 = adamantan-1-ol, Et3CO, tBuO, MeO, 4-methylpiperidin-1-ol, 384
NMePh, NHPh Br Br

CF3

Scheme 67 Enantioselective aerobic oxidation of -keto carbonyl compounds.

In 2018, Meggers et al. reported visible-light induced photocatalytic and enantioselective radical
conjugate additions of -aminoalkyl radicals, generated in situ from the precursor of glycine
derivatives, to ,-unsaturated N-acylpyrazoles for the synthesis of -substituted -aminobutyric acid
derivatives (Scheme 68).105 Some of the reaction products were converted to related bioactive
Page 55 of 83 Organic & Biomolecular Chemistry

55

compounds or pharmaceuticals, such as (S)-nebracetam, (R)-baclofen, (S)-pregabalin, and (R)-Online


View Article
DOI: 10.1039/D0OB00759E
rolipram.

Organic & Biomolecular Chemistry Accepted Manuscript


O
EtO2C CO2Et
R1 R3
t
Bu
385 N S
R2 R3 PF6
+ H O
Me
(2.0 equiv) N N C
O R1 R4 N
O
R3 -179 (8 mol %) Rh
acetone, rt 387 N
N N N C
O R4 23 W CFL 42-89 % yield Me
O up to 97% ee S
386 t
Bu
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

N N
N N -179
R1 =

t
R2 = Ph, Bn, Me, iBu, nPr, npentel, OEt, 2-thienyl, CH2iPr Bu Ph
PF6 S
R3 = H, Me, Ph, Bn; R4 = Boc, Fmoc, Cbz
Me N
C
O F N
EtO2C CO2Et Rh
N N
N R5 C
O R5 F Me Me N
388 N N
N S
+ H N Boc t
Bu Ph
(2.0 equiv)
O 390 -336
O Me -179 (8 mol %)
N N acetone, rt 53-80 % yield
O Boc 23 W CFL up to 96% ee
O 389
R5 = Ph, 4-MePh, 3-MePh, 2-MePh, 4-MeOPh, 4-MeSPh, 4-tBuPh, 4-ClPh

Scheme 68 Enantioselective aerobic oxidation of -keto carbonyl compounds.

Some chiral copper (II) bisoxazoline complexes (CuII-BOX) have been reported by Gong et al.
to possess absorptions in the region of 400–550 nm, owing to metal–ligand charge transfer, which is
an essential characteristic required for visible-light photoredox catalysis. The catalysts were applied
by the group in the enantioselective alkylation of imines with a series of benzyl trifluoroborates under
visible-light irradiation (Scheme 69).106
Organic & Biomolecular Chemistry Page 56 of 83

56

R2 View Article Online


R2
BF3K Cu(BF4)2•H2O (10 mol %) DOI: 10.1039/D0OB00759E
R3
R1 N 394 (11 mol %)
S
+ R3 CHCl3, argon, –20 °C
R1 NH
S
O O 24 W blue LEDs

Organic & Biomolecular Chemistry Accepted Manuscript


391 O O 393
392
R1 = H, MeO, CF3O, CF3, Cl, Me ; R2 = CO2Me, CO2iPr, CO2Et, CO2Bn, 69-88 % yield
R3 = H, F, Cl, Br, CF3, MeO, Ph, CO2Me up to 98% ee

NBoc BF3K BocHN


Cu(BF4)2•H2O (10 mol %) CO2Me
R4 398 (11 mol %) R4 O
O
N + CHCl3, argon, –20 °C N
CO2Me 24 W blue LEDs R5
395 R5 396 397
R4 = H, F, Cl, Br, MeO 84-99 % yield
R5 = Bn, Me up to 94% ee
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Bn Bn
O O
N N
O O
N N
398

394

Scheme 69 Enantioselective alkylation of imines by benzyl trifluoroborates.

In 2018, Meggers et al. reported an enantioselective [3+2] cycloaddition of cyclopropanes to


alkynes or alkenes via visible-light photoredox catalysis using a rhodium-based chiral photocatalyst
(Scheme 70).107 The reactions gave high product yields with excellent enantioselectivities. The
cycloaddition reaction mechanism has been proposed to proceed through direct photoexcitation of
the single-electron transfer reduction of the catalyst–substrate complex, followed by radical ring-
opening of cyclopropane and subsequent cycloaddition of an alkyne or olefin.
O R3
O R1 R4
R2 R3 -179 (2-4 mmol %) N
N DIPEA (50-200 mol %)
N
+ R4 acetone, N2 N PF6 t
Bu
Ph R2 S Ph
Ph blue LEDs, rt R1
400 401
399 Me N
up to 99% yield C
up to 99% ee N
Rh
O R6 N
C N
O R5
R5 -179 (2-4 mmol %) N
Me
N R6 DIPEA (50-200 mol %) S
t
Bu Ph
+ acetone, N2 N
N
Mes blue LEDs, rt
Mes
R5
R5 -336
403
402 404
up to 99% yield
up to 99% ee
R1 = R2 = Me, (CH2)4, (CH2)3, or R1 = Ph, R2 = Me, etc
R3 = CO2Me, COPh, P(O)(OEt)2, SO2Ph, Ph, 4-MeOPh, 4-FPh, 4-BrPh, 4-CNPh, pyridin-2-yl, 4-CO2MePh, CN
R4 = H, Me; R5 = Me, (CH2)3; R6 = 4-MeOPh, 3,5-di-CF3Ph, 4-BrPh, 4-MePh, 3-MePh, 2-MePh, 3-thienyl

Scheme 70 Enantioselective [3+2] cycloadditions of cyclopropanes.

8. Chiral anion binding catalysis with photocatalyst

In 2014, Stephenson et al. introduced a strategy of merging asymmetric anion-binding catalysis


Page 57 of 83 Organic & Biomolecular Chemistry

57

with photoredox catalysis for an enantioselective alkylation of N-aryltetrahydroisoquinolines View


405Article
andOnline
DOI: 10.1039/D0OB00759E
silyl ketene acetal 406 using photoredox catalyst Ru(bpy)3Cl2 and asymmetric catalysis 408, affording
-amino esters 407 in good yields and enantioselectivities (Scheme 71).108 The process was achieved

Organic & Biomolecular Chemistry Accepted Manuscript


as follows. A solution of tetrahydroisoquinoline 405a, photocatalyst, and CCl4 in CH3CN was
irradiated with blue light for 16 h, followed by concentration to give crude iminium product 410.
Crude compound 410 was then reacted with silyl ketene acetal 406 and asymmetric catalyst 408 in
methyl tert-butyl ether (MTBE) at 60 °C for 16 h. As shown in Scheme 71, the process achieved
photoredox catalysis of amine oxidation and subsequent chiral anion-binding catalysis109 of the
imine to give the desired product. Applying the dual function of oxidative photoredox activation and
asymmetric anion binding catalysis in this asymmetric synthesis process provided a distinct strategy
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

for enantioselective photocatalysis. Furthermore, N-dearylation of product 408 was achieved by


reacting with [Fe(bpy)3](PF6)2 and CAN to give secondary amine 409, which was smoothly converted
to 410 using acetyl chloride and pyridine. This conversion expanded the examples of aryl-substituted
derivatives to other functionalities and demonstrates the synthetic utility of chiral
tetrahydroisoquinolines.
(1) Ru(bpy)3Cl2 (1 mol %)
OTBS CCl4 (4 equiv), CH3CN, blue LED, 16 h R
R
N
+ N
Ar OMe (2) CF3 Ar
405 406 t
Bu S 407 CO2Me
(2.0 equiv)
N
R = H, Cl, OMe N N CF3
S H H
Ar = Ph, MeOC6H4, BrC6H4, O
ClC6H4, (MeO)2C6H3
408 (20 mol %), MTBE, 16 h, -60 oC

OMe CAN, CH3CN, H2O OMe


N [Fe(bpy)3](PF6)2 N AcCl, pyridine N
87% O
MeO2C MeO2C MeO2C
408 409 410

CCl4 OTBS
Ru(bpy)32+ N OMe
Ar
blue LED 406
CCl3 + Cl 407a
CO2Me
photoredox
Ru(bpy)3 2+
* catalysis Ru(bpy)3+ HCCl3 cat. 408
anion-binding
asymmetric
catalysis

N – e- N –H N
Ph amine oxidation Ph Ph
409 N
405a 410 Ph
lower pKa, lower BDE Cl
– e- 413
–H H H
F 3C N N
CCl4 R*
N N S
Ph Ph
411 412 Cl CF3

Scheme 71 Merging photocatalysis and anion-binding asymmetric catalysis.

9. Metal catalysis
Organic & Biomolecular Chemistry Page 58 of 83

58

In 2014, Molander and coworkers developed an enantioselective photoredox cross-coupling View Article Online
DOI: 10.1039/D0OB00759E
reaction of racemic trifluoroborate with methyl 3-bromobenzoate to afford a 1,1-diarylethane product
in 52% yield with 50% ee (Scheme 72).110 Furthermore, a single-electron mechanism for

Organic & Biomolecular Chemistry Accepted Manuscript


transmetalation in the photoredox/nickel cross-coupling with organoboron was proposed by the
authors, as shown in Scheme 72. A detailed mechanistic investigation of the reaction by density
functional theory was reported later, with the dynamic kinetic resolution (DKR) of Ni-catalyzed111
cross-couplings presented.112 More recently, the Macmillan group reported an enantioselective Csp3–
Csp2 cross-coupling of aryl halides and -amino acids through the synergistic action of photoredox
and nickel catalysts, affording to the chiral benzylic amines with high enantioselectivities (Scheme
73).113
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ir[dFCF3ppy]2(bpy)PF6 (2 mol %)
417 (3 mol %) Me
Me Br CO2Me Ni(COD)2 (3 mol %) CO2Me
+ Ph
Ph BF3K 2,6-lutidine
THF–MeOH, blue LED, 24 h
414 415 416
racemic Bn N N Bn
52% yield; 50% ee

O O 417

Scheme 72 Enantioselective cross-coupling reaction.


[Ir[dF(CF3)ppy]2(dtbbpy)]PF6 (2 mol %) CN
X NiCl2•glyme (2 mol %) X
CO2H Y 421 (2.2 mol %), TBAI Y O O
R1 R2
+
NHBoc Br 4-BrPhCN R1 NH N
Cs2CO3, DME–toluene R2
418 419 rt, blue LED NHBoc
420
R1 = iPr, nPr, OBn, CO2Me, CH2CO2Me, (CH2)3NHCbz, (CH2)2Cl, yield up to 84% 421
Ph, 4-BocOPh, 3-indolyl, 2-thienyl up to 93% ee
X = CH, N; Y = CCN, CAc, CH, N
R2 = H, F, CO2Me, CF3, F

Scheme 73 Enantioselective carboxylative alkylation.

Li and coworkers reported an enantioselective cross-dehydrogenative-coupling of


tetrahydroisoquinolines and terminal alkynes under visible-light irradiation. The process was
accomplished by combining chiral copper catalysis and photocatalysis (Scheme 74).114 The reactions
were conducted under 26-W household compact florescence light with CuBr (10 mmol%), ligand 4
(15 mmol%), Ir(ppy)2(dtbbpy)PF6 (1 mmol %), and benzoyl peroxide (or under O2), in CH3CN–HF
(1:1) at 20 °C for 2 days. This methodology demonstrated the synergistic cooperation of photoredox
catalysis. Unfortunately, a rational mechanism accounting for enantioselective induction in the
copper catalysis cycle has yet to be proposed, and the absolute configurations of the adducts have not
been determined.
Page 59 of 83 Organic & Biomolecular Chemistry

59

Ir(ppy)2(dtbbpy)PF6 View Article Online


DOI: 10.1039/D0OB00759E
(1 mmol %) R1
R1
N
+ H R2
* N
Ar CuBr (10 mmol %) Ar
423 425 (15 mmol %)

Organic & Biomolecular Chemistry Accepted Manuscript


422 (BzO)2 (1.2 equiv) or O2
CH3CN–THF (1:1)
CFL, –20 °C 424 R2
N
30-90 % yield
PPh2 60-97% ee

425

Scheme 74. Enantioselective coupling of tetrahydroisoquinolines and alkynes.


Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Xiao and coworkers reported an enantioselective photocatalytic oxidation of dihydro-1H-


indene-type -ketoesters with oxygen through complexation of a linker of the chiral bisoxazoline
(boxes) ligand with photosensitizer thioxanthen-9-one and catalytic Ni(acac)2, producing the
corresponding -hydroxy products in yields of up to 97% with up to 95% ee (Scheme 75).115
Furthermore, the reactions were conducted under irradiation by the 3W violet LEDs with a
wavenumber of around 390–410 nm. Furthermore, the reaction was not only limited to benzene
examples but also applied to some examples of pyridine and thiophene systems.

O 428 (10 mol %) O


Ni(acac)2 (10 mol %) OH O
R1 CO2R2 S
R1 CO2R2
toluene, –15 °C, O2 O
n 3W violet LEDs n
O O
426 427
Ph Ph
n = 1, 2, 3 N N 428
R1 = Me, MeO, F, Cl, Br, TBSO, MeCCH2O, allyl Ph Ph
R2 = adamantanyl, tBu,

Scheme 75 Enantioselective photocatalytic aerobic oxidation.

Xiao et al. later reported the enantioselective perfluoroalkylation and difluoroalkylation


reactions of -ketoesters by combining photoredox, chiral Lewis acid, and nickel catalysis cascades
(Scheme 76).116
Ir[dF(CF3)ppy]2(dtbbpy)PF6 O
O (3 mol %), NiBr2•glyme O O
432 (20 mol %) CO2Ad
F F
R1
R1 CO2Ad + I R2 NaHCO3 (2.0 equiv) CF2R2
Bn
N N
Bn
n
n DME, degas, rt 432
430 7 W blue LEDs, 30 h 431
429
n = 1, 2 40-67% yield
R1 = H, Me, MeO, F, Cl, Br, OTf up to 90% ee
R2 = CO2Et, C2F5, C3F7; Ad = 1-adamantanyl

Scheme 76 Enantioselective photocatalytic aerobic oxidation.

Liu, Lin, and coworkers reported an enantioselective decarboxylative radical cyanation reaction
of N-hydroxy-phthalimide (NHP) esters to afford the corresponding chiral alkyl nitriles through
synergistic action of photocatalysis and asymmetric copper catalysis (Scheme 77).117
Organic & Biomolecular Chemistry Page 60 of 83

60

View Article Online


DOI: 10.1039/D0OB00759E

R2
R2 Ir(ppy)3 (0.5 mol %)
O O O

Organic & Biomolecular Chemistry Accepted Manuscript


O CuBr (1 mol %) CN
N 435 (1.2 mol %) R1 N N
R1
O 434
O TMSCN (1.1 equiv)
DMF/p-xylene (4:6) 59-98 % yield 435
433 (racemic) Ar, blue LEDs up to 99% ee
R1 = Me, Cl, Br, tBu, MeO, and other aromatic system, e.g. 1-naphthyl, 3-pyridinyl, benzo[b]thiophenyl, etc
R2 = Me, Et, iPr, Cy, allyl, (CH2)3OMe, CH2CO2Et

434 blue LEDs


435-CuICN
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

IrIII

asymmeric IrIII *
SET photocatalysis
coppercatalysis

435-CuII(CN)2
435-CuIICN
IrII SET

R2 R2 O 433
O + N
-CO2
O
O
Scheme 77 Enantioselective decarboxylative cyanation reaction.

In 2017, Lu, Xiao, and coworkers reported a sequential enantioselective Pd-catalyzed


photoactivated [4+2] cycloaddition of vinyl benzoxazinanones and -diazoketones (Scheme 78).118
Visible-light activation of -diazoketones facilitates Wolff rearrangement, providing an intermediate
ketene for the subsequent Pd-catalyzed enantioselective [4+2] annulation.
Page 61 of 83 Organic & Biomolecular Chemistry

61

View Article Online

R1 Br DOI: 10.1039/D0OB00759E
N2 Pd2(Dba)3-CHCl3 (5 mol %) R1 R2
O R2 439 (11 mol %), CH2Cl2, rt
+ R3 R3

Organic & Biomolecular Chemistry Accepted Manuscript


6W blue LEDs
N O O N O
Ts Ts Ph
437 438 S
436
439
(racemic) dr up to >95:5 75-99% yield
up to 96% ee Ph NBn
R1 = Me, Br, Cl, F, CF3, MeO
(PhO)2P
R2 = Ph, iPr, tBu, cyclopropyl; R3 = Me, iPr, Bn, (CH2)2CH=CH2, (CH2)3OBn, H

blue LEDs O
[Pd]*
437 C
Wolff O asymmetric
rearrangement allylic alkylation
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R3 R2 R2
N
Ts
R3 [Pd]*

R2
[Pd]*
R3
N O
NTs Ts

ligand
[Pd]* exchange

436
O

N O
Ts 438

Scheme 78 Enantioselective Pd-catalyzed photoactivated [4+2] cycloaddition.

Xiao et al. reported dual nickel catalysis and visible-light photoredox catalysis in desymmetric
C–O coupling reactions of 2-(2-halophenoxy)-propane-1,3-diols for the enantioselective synthesis of
1,4-benzodioxanes (Scheme 79).119 This process was realized by deploying an axially chiral 2,2’-
bipyridine ligand L3 and photocatalyst Ir(dFCF3ppy)2(dtbbpy)PF6.

CF3
F
t
Bu
Ir(dFCF3ppy)2(dtbbpy)PF6 (3 mol %) N
R2 R2
O NiCl2•glyme (5 mol %) N
OH O N F
442 (5 mol %), K2CO3 OH O Ir
R1 R1 F PF6
I OH quinuclidine (20 mol %) O
O N N
THF, 30W blue LEDs N
440 441 t
Bu
t 442 F
R1 = H, Et, Bn, Bu, Cy, F, iPr, Cl, F, Me, PhMe2C, 65-84% yield CF3
R2 = H, Me up to 76% ee
Ir(dFCF3ppy)2(dtbbpy)PF6

Scheme 79 Enantioselective synthesis of 1,4-benzodioxanes.

Mei, Han, and coworkers reported an enantioselective radical cyanoalkylation of styrenes by


incorporating copper catalysis and photoredox catalysis (Scheme 80).120 Alkyl N-
hydroxyphthalimide esters are introduced as alkylation reagents, and the chiral Box/Cu(II) catalyst is
Organic & Biomolecular Chemistry Page 62 of 83

62

responsible for enantioselectivity. View Article Online


DOI: 10.1039/D0OB00759E

O
O Ir(ppy)3 (0.5 mol %)

Organic & Biomolecular Chemistry Accepted Manuscript


CN
CuBr (1.0 mol %) O O
N R2
R1 R2 O 446 (1.2 mol %)
+ 444 R1 N N
O
+ bNMP–ClPh, rt 445
443 blue LEDs, 24 h 446
TMSCN up to 85% yield
up to 94 % ee
R1 = H, F, Cl, Br, CF3, Me, tBu
R2 = Me, Et, nPr, iPr, tBu, cyclobutyl, cyclopentyl, Ph(CH2)2, 4-MePh(CH2)2,

Scheme 80 Photocatalytic cyanoalkylation of styrenes.


Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

In 2018, Feng et al. applied a visible-light excitable alkenoylpyridine-bound N,N’-dioxide–TbIII


complex to the enantioselective [2+2] photocycloaddition reactions of enones with dienes (Scheme
81).121 Furthermore, the combination of photocatalyst [Ru(bpy)3]Cl2 with a chiral N,N’-dioxide–ScIII
complex for the enantioselective photocycloaddition reactions of 2’-hydroxychalcones using alkenes
has been established. Both methods provide the corresponding cyclobutanes with good yields and
enantioselectivities.
Sc(OTf)3/ 450 OH O
OH O (1:1, 10 mol %)
R2
[Ru(bpy)3]Cl2 R1
(5 mol %), iPrCN O N N
R1 + i
Pr
O
i
Pr
N H O O
447
5 Å MS, rt H N
448 23W CFL R2
449 i i
R1 = H, Me, F, Cl, Br Pr 450 Pr
R2 = H, Me, F, Cl, Br, CF3 dr up to 6:1 up to 99% yield
up to 92% ee
O
O Tb(OTf)3/453 N
R2
N (1:1, 10 mol %) R1 O N 2 N O
i i
+ Pr Pr
R1 N H O O
H N
i
PrCN, 5 Å MS, rt
451 23W CFL
448 i i
452 R2 Pr 453 Pr
R1 = H, Me, F
R2 = H, Me, F, Cl, Br, CF3 dr up to >19:1 up to 75% yield
up to 96% ee

Scheme 81 Enantioselective [2+2] photocycloaddition.

In 2018, Gong et al. reported an enantioselective conjugate addition of ,-unsaturated amide


derivatives and -silylamines using a NiII–DBFOX photoredox catalyst under blue LED irradiation
(Scheme 82).122 The chiral catalyst participated in the photoredox processes to initiate single-electron
transfer, while providing a coherent chiral setting for the subsequent free radical reaction.
R1 R2
O R2 Ni(ClO4)2•6H2O O
TMS (10 mol %) N
N 457 (12 mol %) N
R1
R3 N O
N + THF, Ar, 25 °C
456 O O
454 24W blue LEDs R3 N N
8-16 h yield up to 85%
455 up to 96% ee 457
Ph Ph
n c n
R1 = Me, Et, Pr, Hex, Bu, OEt, CF3, CO2Me, Ph, 4-MePh, 4-MeOPh, 4-FPh, 3-furanyl
R2 = Me, Et, iPr, Ph; R3 = H, Me, Br, F, OMe,
Page 63 of 83 Organic & Biomolecular Chemistry

63

Scheme 82 Enantioselective radical addition. View Article Online


DOI: 10.1039/D0OB00759E

In 2018, Yu et al. reported an enantioselective allylic alkylation of allyl acetates with 4‑alkyl-

Organic & Biomolecular Chemistry Accepted Manuscript


1,4-dihydropyridines by combining photoredox catalysis and palladium catalysis (Scheme 83).123
The alkyl radicals were generated from 4-alkyl-1,4-dihydropyridine under visible light redox
conditions and underwent a free radical coupling reaction with a -allyl palladium complex to
perform an asymmetric Pd-catalyzed allyl alkylation.
R3
R4
R2 R4 Ir(ppy)2(dtbbpy)PF6 (2 mmol) R3
OAc Pd2(dba)3 (2.5 mol %)
R1 EtO2C CO2Et 461 (6 mmol) R2
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

+ K3CO3, CH3CN, N2
blue LEDs, 12 h R1
N 460
H
458
459
44-84 % yield
R1 = H, OMe, Me, F, Cl, CF3, CO2Me, NHAc, 2-naphthyl, 2-thieny, 2-furanyl, up to 96% ee
R2 = Ph, 4-MeOPh, 4-MePh, 4-ClPh, 3-MeOPh, 3-MePh, 3-ClPh, 2-MeOPh,
2-MePh, 2-FPh, 2-ClPh, 2-thienyl, BnO, BocNH
R3 = H, Me, Ph; R4 = H, Me
OMe

t PF6
Bu
N
MeO PAr2
N
MeO PAr2
Ir
N
t N
Bu OMe 461
Ar = 4-OMe-3,5-tBu2-C6H2
Ir(ppy)2(dtbbpy)PF6 (R)-GARPHOS

Scheme 83 Enantioselective allylic alkylation.

In 2019, Cabrera, Alemán, and coworkers reported an enantioselective [3+2] cycloaddition of


,-unsaturated N-acyl oxazolidinones to silyl-indole derivatives using a dual catalytic system of
Lewis acid catalysis and visible-light photoredox catalysis, affording pyrrolo[1,2-a]indoles (Scheme
84).124 Chiral ligand (R,R)-iPr-PyBOX, which gave a moderate ee, was found to be the best candidate
for this enantioselective reaction.
Organic & Biomolecular Chemistry Page 64 of 83

64

Yb(OTf)3 (15 mol %) R1 View Article Online


R1 O O O DOI: 10.1039/D0OB00759E
466 (20 mol %)
R3 N 465 (2 mol %) R2 X
R2 + N

Organic & Biomolecular Chemistry Accepted Manuscript


N N CNCH2CO2Et (2 equiv)
Bn CH3CN, rt, N2 R3
TMS 464
462 463 white LED, 18 h
up to 80 % yield
R1 = Me, CO2Me; R2 = H, Br; R3 = Me, iPr up to >95:5 dr
40-70% ee
F3C
F i
t
Pr
Bu O
N PF6 N
N
F
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ir
F 466 N
N
N
t
Bu N
O
F i
Pr
F3C
465

Scheme 84 Enantioselective [3+2] annulation.

Wang, Xu, and coworkers reported a chiral Cu-catalyst-induced enantioselective


cyanofluoroalkylation of styrenes under violet LED irradiation (Scheme 85).125 For the
enantioselective formation of C–CN bonds, the chiral Cu catalyst served as both the visible-light
photoredox catalyst and electron transfer catalyst.
R1
R1 CuI (10 mol %), 435 (12 mol %)
467 O O
N,N-dimethylbenzylamine (1.5-3.0 equiv)
H2O (2.0 equiv), CH3CN N N
+ 24 W violet LED (390–410 nm)
R2
CN
R 2I + TMSCN 470
435
468 469 up to 93% yield
up to 96% ee
R1 = H, Me, MeO, F, tBu, AcO, MeS, CH2OH, CH2OCH2OMe, Ph, etc
R2 = C4F9, CF3, CF2CO2Et

Scheme 85 Enantioselective cyanofluoroalkylation of styrenes.

In 2019, Mitsunuma, Kanai, and coworkers reported an enantioselective allylation of aldehydes


with cooperative chiral chromium complex catalysis and organophotoredox catalysis using an
acridinium catalyst (Scheme 86).126 The chiral chromium hybrid complex not only served as chiral
allylchromium nucleophiles for enantioselective allylation of aldehydes, but the chromium(III)
complex intermediate also served as a reductant for the electron-transfer reduction of photocatalyst
radicals to regenerate the acridinium catalyst and complete the catalytic cycle. Furthermore, the
addition of magnesium salt, such as Mg(ClO4)2, significantly increased both the enantioselectivity
and reactivity. Additionally, the enantioselective allylation of 4-fluorobenzaldehyde with 4-allyl-1,2-
dimethoxybenzene with photocatalyst [Ir(dFCF3ppy)(dtbbpy)PF6] (2 mol%), Cr2Cl (10 mol%), and
pyridine (20 mol%) in DMF for blue light irradiation has been reported by Glorius et al.127
Page 65 of 83 Organic & Biomolecular Chemistry

65

View Article Online


DOI: 10.1039/D0OB00759E
R4 CrCl2/435 (2.5-20 mol %) R4 OH O O
O 474 or 475 (0.5-5 mol) R2
+ R2 R1 N N
R1 Mg(ClO4)2 or MgPhPO3

Organic & Biomolecular Chemistry Accepted Manuscript


H
R3 R5 472 (20-100 mol %) R5 R3
471 473
CH2Cl2, rt, 12-48 h 435
38 W LED (430 nm) up to 97% yield
85-99 % ee R

R1 = Ph, 4-ClPh, 4-BrPh, 4-IPh, 3-AcPh, 4-CO2MePh, 4-CF3Ph, 3-MePh,


2-MePh, 3-MeOPh, 2-furanyl, Cy, Ph(CH2)2, nPent, Me2CHCH2
Me Me
List of compound 472:

N
Me
ArAcr ClO4
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

Ar = Acr = 10-methylacridinium 474 R = H; 475 R = Me


or

N
Me

Ar Acr •X
472a
LED

ArAcr
ArAcr •X
X
* OH

N N R1
– H+ CrIII
473a
* X X X

N N
CrII
+ H+
X X
*
* N N
CrIII
N N
O X X
ArAcr •X CrIII

X X R1

R1 H 471

Scheme 86 Enantioselective allylation of aldehydes.

Xiao et al. reported enantioselective radical conjugated addition (photo-Giese reactions) of


enones using alkyl Hantzsch esters with photocatalyst Mes-AcrClO4 (9-mesityl-10-methylacridinium
perchlorate) and a chiral cobalt (II) catalyst, formed in situ from the addition of CoCl2•6H2O and
chiral N4-ligand 479 under blue LED irradiation (Scheme 87).128 Notably, for the reaction with acyl
Hantzsch esters, the reactions proceeded in the absence of a photoredox catalyst.
Organic & Biomolecular Chemistry Page 66 of 83

66

CoCl2•6H2O (8 mol %) R1
View Article Online
O R2 479 (9.6 mol %) O DOI: 10.1039/D0OB00759E
N EtO2C CO2Et AgOTf (16 mol %) N
R1 R2
N + Mes-AcrClO4 (2 mol %) N

Organic & Biomolecular Chemistry Accepted Manuscript


Ph N CH3CN, rt, 20-48 h Ph
H 478
3W blue LEDs
476 477 49-90% yield
up to 91% ee

R1 = Me, Et, nPr, iPr, Ph, 4-BrPh, 4-MePh, 2-BrPh, 4-FPh, 2-thienyl, 3-thienyl, nPent, Ph(CH2)2,
R2 = Bn, 4-BrPhCH2, 4-ClPhCH2, 4-MePhCH2, 4-MeOPhCH2, 2-MeOPhCH2, 3-MeOPhCH2,
3,4-diMeO2PhCH2, 2-nathylCH2, 2-thienylCH2, cyclopentyl, cyclohexyl,
tetrahydro-2H-pyranyl, N-Boc-4-piperidinyl, NBocCH2,

O R4 R3
O CoCl2•6H2O (8 mol %) O
N EtO2C CO2Et 479 (9.6 mol %) N O
R3 AgOTf (16 mol %)
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

N + N
N R4
Ph CH3CN, rt, 2-12 h Ph 482
H
480 3W blue LEDs
481 47 to >99% yield
up to 94% ee
R1 = Me, Ph, 4-BrPh, 4-FPh, 4-MePh, 2-BrPh, 2-thienyl, nPr, nhexyl, chexyl, iPr
R4 = Ph, 4-FPh, 4-MePh, 4-MeOPh, 3-MePh, 2-MePh, 2-naphthyl, 2-thienyl, cyclohexyl, 1-adamantyl

Ph Ph

N N Et Cl
N
Et Et Ar
N N N
N N Me
Cl N Ph
Co
H 2O N Ph
Me
N
i i N
i
Pr Pr iPr Pr Ar
N ClO4
Et
i i -Co•479 Mes-AcrClO4
Pr Pr
479 Ar = 2,4,6-iPr3C6H2

Scheme 87 Enantioselective photo-Giese reactions.

In 2019, Gong et al. reported an enantioselective -aminoalkylation of oxazolidin-2-one-derived


imine with -silylarylamine using asymmetric copper catalysis under blue LED irradiation (Scheme
88).129 In the presence of photocatalyst Ir[dF(CF3)ppy]2(dtbbpy)PF6 (2 mol%), the reaction yield was
higher, but the reaction was feasible without a photocatalyst under visible-light irradiation. Individual
chiral ligand L7 and oxazolidin-2-one-derived imine were found to have no significant absorption in
the visible region, but their reduced state [L7-CuI-imine] had an obviously enhanced absorption.
Therefore, the copper (I) substance generated in situ from the reaction process can be concluded to
be used as a photosensitizer in the reaction.
O O
O R2 O
R1 Cu(OTf)2 (10 mol %)
N TMS N 486 (11 mol %) N
N NH R2
+ –40 °C, THF, argon N
R1
Ar H 24W blue LEDs, 40 h Ar O O
484
483 485 N N
with or without Ir[dF(CF3)ppy]2(dtbbpy)PF6 (2 mol %)
55-82% Ph Ph
486
up to 98% ee
Ar = Ph, 2-MePh, 3-MePh, 4-MePh, 2-MeOPh, 2-PhPh, 2-FPh, 2-ClPh, 2-BrPh, 2-naphthyl
R1 = H, Br, Me, Cl, F; R2 = Me, iPr, H, Ph, Et

Scheme 88 Enantioselective -aminoalkylation of imines.


Page 67 of 83 Organic & Biomolecular Chemistry

67

View Article Online


DOI: 10.1039/D0OB00759E
Glorius et al. reported a three-component enantioselective dialkylation of butadiene with
aldehydes and Hantzsch esters by chromium catalysis (with a chromium–bisoxazoline complex) and

Organic & Biomolecular Chemistry Accepted Manuscript


visible-light photocatalysis (organic dye 4CzIPN) (Scheme 89).130
491 (2 mol %) OH
R2 R3
O 435 (10 mol %) R3
EtO2C CO2Et CrCl2 (10 mol %) R1
R1 H + +
THF, rt, 24 h
R2 N
487 488 34W blue LEDs
H 490
R1 = Ph(CH2)2, Et2CH 489 57-80% yield
R2 = H, Me 84-96% ee
R3 = iPr, iBu, Et2CHCH2, CyCH2
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

N O O
NC CN
N N

N N
435
N
491
4CzIPN

Scheme 89 Enantioselective dialkylation of butadiene.

Zhang et al. reported a photo-driven enantioselective dual arylation/alkylation and alkynylation


of styrenes, catalyzed by chiral copper (I) complex tBu-BOPA (Scheme 90).131 Presumably, the
BOPA-Cu (I) complex aggregated with alkynes, serving as a photo-reducing agent and enantiomeric
induction source. Expanding the scope of the phenyl group in styrene to other heteroaryl groups, such
as thiophene, is feasible in this three-component coupling reaction.
R1
CuI (10 mol %)
R1 R2
496 (20 mol %)
K3PO4 (3 equiv)
+ R2 H + R3 I
CH3CN, 0 °C or rt
(R)

492 493 494 40 W blue LEDs R3


495
42-87% yield
R1 = H, F, Cl, Br, CH3, tBu, MeO, CF3, Ph up to 98% ee
R2 = SiiPr3, SiMe3, SiEt3, (CH2)3CH3, (CH2)3Ph,
t
Bu, iPr, sBu, cyclopropyl, 3-pyridinyl, 2-thienyl, N
H
3-methylphenyl, 2-chlorophenyl, 4-methoxyphenyl O O
N N
R3 = CH2CHF2, CH2CF2CF2H, CH2CF2CF3, iPr, Ph,
4-tBuPh, 4-MeOPh, 4-BrPh, 4-PhPh, 2-thienyl, t t
Bu Bu
CH2F, nBu, 1-naphthyl, 3-pyridinyl
tBu-BOPA
(496)

Scheme 90 nantioselective dual alkylation and alkynylation.

Wang et al. reported an enantioselective acyl-carbamoylation of alkenes bound to aryl carbamic


chlorides with aldehydes by combining nickel catalysis (using a chiral Ni-PHOX complex) with
photocatalysis (using an advantageous hydrogen-atom transfer photocatalyst, tetrabutylammonium
Organic & Biomolecular Chemistry Page 68 of 83

68

decatungstate) to give a series of oxindoles bearing a quaternary stereogenic center (Scheme View
91). 132Online
Article
DOI: 10.1039/D0OB00759E
R2 TBADT (5 mol %) R2 O
O
t
O Ni(OTf)2 (10 mol %) Bu

Organic & Biomolecular Chemistry Accepted Manuscript


500 (12 mol %) R4 N
R1 R1 O
R3 + H
K3PO4 (1.1 equiv) N
N PPh2
R4 CH3CN, 9h R3 500
497 O Cl 498 blue LED (390 nm) 499
up to 96% yield Tetrabutylammonium
R1 = H, , MeO, BnO, Me, Cl, CF3; R2 = Me, Et, nPr, iPr, cHex, up to 96% ee decatungstate (TBADT)
R3 = Me, Bn, nPent, PMB
R4 = nBu, iPr, tBu, Bn, BzO(CH2)2, TBDPS(CH2)2, BocNHCH2,
PhC(O), 4-MePhC(O), 4-tBuPhC(O), 4-FPhC(O), 3-MePhC(O),
2-MePhC(O), 3,4-(MeO)2PhC(O),
Scheme 91 Enantioselective dual alkylation and alkynylation.
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

10. N-Heterocyclic carbene catalysis with photocatalyst


Rovis et al. developed an -acylation of tertiary amines with aldehydes by merging photoredox
catalysis with chiral N-heterocyclic carbene catalysis (Scheme 92).133 The iminium ion was formed
by visible-light photoredox catalysis of tertiary amine 1. Accordingly, the acyl anion equivalent was
derived from aldehyde 2 using an N-heterocyclic carbene (NHC) catalyst, followed by nucleophilic
addition to the subsequently formed iminium ion, affording asymmetric -acylated amine 3.
Page 69 of 83 Organic & Biomolecular Chemistry

69

View Article Online


O DOI: 10.1039/D0OB00759E
Ru(bpy)3Cl2 (1 mol %) R1
O N
m-dinitrobenzene Ar N Br
R1 504 (10 mol %), CH2Cl2 N
N + R2 H R2 O
N
Ar blue light (~450 nm)

Organic & Biomolecular Chemistry Accepted Manuscript


501 502 503 Br
504 Br
51–94% yield
up to 92 ee
R1 = H, 6,7-MeO; Ar = Ph, 4-MeC6H4, 4-Br, C6H4, 4-MeOC6H4, 4-CF3C6H4
R2 = Me, Et, nPr, iPr, cyclopropyl, MeS(CH2)2, Ph(CH2)2 CH2=CH(CH2)2,
PhthN(CH2)3, AcO(CH2)3

NO2
Ox e.g.
Ru(bpy)32+
blue light
NO2
Ox
photoredox
Ru(bpy)32+* catalysis Ru(bpy)3+
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

PC

N N N
Ph Ph – e- Ph
501
iminium ion
–H
–H

– e-
N N
Ph Ph
503
O
N Br
N
N

HO Br Br
R2 O
N Br
O NHC O N N
asymmetric Br
R2 R2 H HO Br
NHC catalysis
acyl anion 502 R2
O N
Ar
N Br
N
N

Br
504 Br
N
Ar

R2 O 503

Scheme 92 -Acylation of N-phenyltetrahydroisoquinoline.

An unprecedented N-heterocyclic carbene-catalyzed - and -alkylation using alkyl radicals


under visible-light irradiation, but affording low enantioselective, was reported by Gao, Ye, and
coworkers (Scheme 93).134
Organic & Biomolecular Chemistry Page 70 of 83

70

View Article Online


DOI: 10.1039/D0OB00759E
N N
N

Organic & Biomolecular Chemistry Accepted Manuscript


OCO2Me CN
BF4 Mes
O 508 (20 mol %)
Ph + I CN Ph
O
H Ru(bpy)3(PF6)2 (2 mol %)
*
505 506 H
CaOAc (2.0 equiv) 507
MeOH (5.0 equiv) 57% yield
DCE, 25 °C, 32 W CFL 29% ee

N N
N CO2Et
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

OCO2Me CO2Et BF4 Mes Bn CO2Et

O Br CO2Et 508 (20 mol %)


+ O
H Bn 510 Ru(bpy)3(PF6)2 (2 mol %) *
509 CaOAc (2.0 equiv) H
511
MeOH (5.0 equiv)
51% yield
DCE, 25 °C, 32 W CFL
12% ee

Scheme 93 Enantioselective - and -alkylation.

11. Enzyme catalysis with photocatalyst

Visible-light photobiocatalysis is a new method that attempts to combine the characteristics of


enzymatic catalysis and photocatalysis. The photobiocatalytic method has two approaches, namely,
the use of an external photocatalyst, or that in which no additional photocatalyst is necessary. A brief
review of recent advances in the synthesis of volatile sulfur compounds via photobiocatalytic
cascades has been reported.135
In 2016, Hyster et al. demonstrated the first enantioselective example of the synergistic action
of photocatalysis and biocatalytic techniques (Scheme 94).136 This strategy modified the enzyme
catalytic activity by applying the photo-excited state of its nicotinamide cofactors. For example,
NADPH (the reduced form of nicotinamide adenine dinucleotide phosphate) is both a source of
hydrogen atoms and a photoreductive agent, while under visible-light irradiation, the ketoreductase
is converted from carbonyl reductase into a radical species initiator and a chiral source of hydrogen
atoms. Later, the same group also extended the strategy of applying the photoexcitation of an EDA
complex of an enzyme and substrate to directly promote electron transfer in several enantioselective
reactions. Examples include the enantioselective deacetoxylation of -acetoxytetralone,
photoenzymatic catalytic ketone reduction, enantioselective radical cyclization of -chloroacetamide
to afford lactam by the photoexcitation of flavoenzymes ('ene'-reductase), and enantioselective
photoexcited radical cyclization of -haloamides to provide oxindoles using flavin-dependent ‘ene’-
reductases (Scheme 95.).137
Page 71 of 83 Organic & Biomolecular Chemistry

71

O RasADH (1 mol %) O LKADH (0.25 mol %) O View Article Online


NADP+ (1 mole %) Br DOI: 10.1039/D0OB00759E
R R NADP+ (0.45 mole %) R
O O O
GDH-105, glucose (200 mM) kPi (pH = 6.5, 1 mM MgCl)
n TRIS (pH = 7.0, 10 mM CaCl) n n
iPrOH–DMSO (460 nm, rt, 12 h)

Organic & Biomolecular Chemistry Accepted Manuscript


(R)-513 glycerol–DMSO (460 nm, rt, 12 h) 512 (S)-513

n = 1, 2; R = Ph, 4-FC6H3, 4-ClC6H3, 4-BrC6H3, 4-CF3C6H3, PhCH2 up to 92%yield; 96% ee

O H H O

O R NH2
photo
N irradiation
(S)-513
R NADPH
cofactor oz
E 1/2 = 0.57 V vs SCE
regeneration H H O *
NH2
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

O O
H N
R NH2 R
O
N NADPH*
oz
R E 1/2* = –2.60 V vs SCE
NADP+ O
Br
O R
asymmetric
H atom 512
transfer

H H O O H H O
O
Br
R NH2 R NH2
O O
N N
R R
mesolytic
cleavage

Br

Scheme 94. Enantioselective photoexcited dehalogenation.


Organic & Biomolecular Chemistry Page 72 of 83

72

2018 Hyster View Article Online


DOI: 10.1039/D0OB00759E
O NtDBR (1.0 mol %) O
R Rb (0.5 mol %), GDH-105
OAc R

Organic & Biomolecular Chemistry Accepted Manuscript


X NADP+ (1 mol %), glucose X
514 KPi (100 mM, pH 8.0)
10% glycerol, 530 nm LEDs 515
racemate
38–85%yield
R = Me, Pr, allyl, iPr, Bn; X = H, OMe, Cl, Br up to 94% ee

2019 Hyster
O MorB (1.0 mol %) OH
Ru(bpy)3Cl2 (1.0 mol %)
R R
X X
NAD(P)+ (2.0 mol %)
516
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

GDH-105, glucose 517


achiral 460 nm LEDs, 24 h
15 to >99%yield
up to 76% ee
R = Me, Pr, allyl, nC5H11, 2-thiophenyl; X = H, OMe, Cl

2019 Hyster
O O
Cl 'ene'-reductase (0.5mol %)
N NADP+ (1.0 mol %) N

518 n n
GDH-105, glucose R1
KPi (100 mM, pH 8.0) R2
R1 R2 497 nm cyan LEDs, 36 h 519
achiral 44-99%yield
R1 = Me, Et, Ph, benzyl, cyclohexyl, vinyl up to >99% ee
R2 = H, Me, Et

2020 Hyster

N O OPR1 (0.5 mol %) N


tricine buffer (pH 8.0, 100 nM) X O
X
i
R1 Cl PrOH (10% v/v), 4 °C
R2 497 nm cyan LED, 24 h. R1 R2
520 521
racemate 39–96%
up to 96% ee
R1 = CO2Me, CO2Et, CO2nPr; R2 = Me, Et, nPr, nBu, iPr, allyl
X = Me, Et, Br. Cl, OMe

Scheme 95. Examples of the combination of enzymes with photocatalysis.

The photosensitized activation of redox enzymes is of interest in promoting electron transfer,


particularly in the context of sustainable and green chemistry using clean light energy. A solar-light-
driven enantioselective ketone reduction through domino photocatalytic–biocatalytic reactions was
reported by Baeg et al. (Scheme 96a).138 The enantioselective reduction of some ketones was
achieved under visible light using an orchestration system comprising a chemically converted
graphene-based photocatalyst (CCG-BODIPY) 524, an electron donor (triethanolamine), an electron
mediator precursor ([Cp*Rh(bpy)Cl]Cl), an electron donor (TEOA), nicotinamide nucleotide
coenzyme (NADP+), and an alcohol dehydrogenase (LKADH, known as a NADPH dependent). This
process imitates the mechanism in which chlorophyll absorbs and converts solar energy in plants.
In contrast, Hollmann et al. reported an enantioselective photobiocatalytic alcohol oxidation and
lactonization of meso-3-methyl-1,5-pentanediol to give (S)-4-methyltetrahydro-2H-pyran-2-one
using NAD+ (>98% ee), which was aerobically regenerated by a flavin photocatalyst (FMN) and blue-
Page 73 of 83 Organic & Biomolecular Chemistry

73

light irradiation from horse liver alcohol dehydrogenase (HLADH) (Scheme 96b).139 Hollmann, Park,
View Article Online
DOI: 10.1039/D0OB00759E
and coworkers also reported that 2-methylcyclohexenone was hydrogenated to 2-
methylcyclohexanone with high enantioselectivity (>99% ee) under irradiation using photosensitizer

Organic & Biomolecular Chemistry Accepted Manuscript


Rose bengal with white LED lamps, and that Flavin-containing enzymes were obtained from the old
yellow enzyme family (OYEs) (Scheme 96c).140
(a) graphene-based photocatalyst
(CCG-BODIPY) 524 OH
O [Cp*Rh(bpy)Cl]Cl
R2 R1 R2 R1
NADP+, LKADH, visible light, rt
522 523
GC-yield: 58-74%
R1 = CH3, CF3; R2 = Ph,
ee: 95 to >99.9%
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

O S

N O

(b) O

HLADH O
HO OH

NAD+ NADH
> 98% ee
H2O2 h
FMN
O2
(c) O
white
LED RB* FMNox

RB
TsOYE
O
donorox
RB FMNred
donor
>99% ee
donor: triethanolamine (TEOA)
TsOYE: flavin-containing OYE homologue from Thermus scotoductus

Scheme 96 Photobiocatalytic oxidation and reduction reactions.

In 2018, Zhou et al. reported an application of redox catalysis and enzyme catalysis to the
dynamic kinetic resolution (DKR) of amines (Scheme 97).141 The racemization of various primary
amines has been achieved through photoredox-induced hydrogen atom transfer (HAT) and enzyme
catalysis.
Ir(ppy)2(dtb-bpy)PF6 (2 mol %) O
n
OctSH (50 mol %), Novozym 435
R NH2 MeO(CH2)2CO2Me (2-7 equiv) R N OMe
H 526
525 CH3CN, 38 °C
racemate 4 Å MS, 32 W white LED, 2d up to 95% yield
99% ee
R = Ph(CH2)2, 4-MePh(CH2)2, 2-MePh(CH2)2, 3-MePh(CH2)2,
4-EtPh(CH2)2, 4-MeOPh(CH2)2, 2,4-diMePh(CH2)2,
3,5-diMePh(CH2)2 PhCH2, 4-MePhCH2, 4-FMPhCH2, 4-ClPhCH2,
i
Pr, nPent, Me2CH(CH2)3, nHex, Cy

Scheme 97. Dynamic kinetic resolution of amines by the combination of enzyme and photoredox catalysis.
Organic & Biomolecular Chemistry Page 74 of 83

74

View Article Online


DOI: 10.1039/D0OB00759E
In 2018, Ward et al. reported an enantioselective synthesis of cyclic amines by incorporating
photoredox catalyst Na3Ir(sppy)3 under visible light irradiation to generate an -amino alkyl radical

Organic & Biomolecular Chemistry Accepted Manuscript


under hydrogen atom transfer (HAT) from ascorbic acid, as well as enzymatic catalytic oxidation of
one cyclic amine enantiomer back to the starting imine by enzyme monoamine oxidase (MAO-N-9)
(Scheme 98).142
MAO-N-9
H
Na3Ir(sppy)3 (1 mol %)
ascorbic Acid (0.2 M) N R NaO3S SO3Na
R H N
N phosphate buffer solution
(pH 8.0), air, 30 h 528
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

527 yield >95% Ir


405 or 450 nm LED (3 W)
up to >99% ee N N
R = nBu, Bn, cHex,

529
Na3Ir(sppy)3 Na3Ir(sppy)3 SO3Na
N R
photoredox H
catalysis

Ascorbic SET
Ascorbic acid acid
N R
H
527
N R
MAO-N-9 H
enzymatic
catalysis O2

H2O2

Scheme 98. Reduction of imines by the combination of enzyme and photoredox catalysis.

Park et al. reported an enantioselective hydrogenation of 2-methylcyclohex-2-en-1-one to give


(R)-2-methylcyclohexan-1-one by employing photosensitizer rose bengal and an ene-reductase from
Thermus scotoductus SA-01 (TsOYE), as well as triethanolamine (TEOA) as an electron donor under
irradiation with white LED (Scheme 99).143
O old yellow enzyme O
>99% ee
Rose bengal max. conversion: 70.4%
triethanolamine turnover frequency: 1.54/min
white LED turnover number: 300.2
530 531
Scheme 99. Enantioselective hydrogenation.

Guan, He, and coworkers reported an enantioselective concomitant oxidation and alkylation of
2-arylindoles by combining photocatalysis (with Ru(bpy)3Cl2•H2O) and enzyme catalysis (with
wheat germ lipase, WGL) under white compact fluorescent lamp irradiation, providing chiral 2,2-
disubstituted indol-3-ones (Scheme 100).144
Page 75 of 83 Organic & Biomolecular Chemistry

75

View Article Online


O R
DOI: 10.1039/D0OB00759E
R2 O Ru(bpy)3Cl2•6 H2O 2

R1 (2 mmol)
N
+ R1
R4

Organic & Biomolecular Chemistry Accepted Manuscript


H WGL (0.16 mol %) N
R3 R4 H
532 DMF, O2, rt, 32 W CFL
533 534 O
R3
R1 = H, F, Br, MeO, Me, O
R2 27-70% yield
R2 = H, CF3, F, Cl, Br, Me, MeO 32-86% ee
R3 = H and R4 = H, Me, Ac R1
R3 = R4 = (CH2)3 N
535

Scheme 100. Enantioselective concomitant oxidation and alkylation.


Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

In 2018, Höhne, Schmidt, and coworkers reported an enantioselective C–H bond


functionalization by merging the organo photoredox catalysis and enzyme catalysis (Scheme 101).145
In their study, alkanes were oxidized to aldehydes and ketones by photocatalyst anthraquinone sulfate
(SAS) under an oxygen balloon atmosphere and irradiated by a household 200-W white light bulb at
30 °C. After complete photooxidation, enzymes and appropriate reagents were added to the reaction
mixtures for one-pot operation. A series of chiral amines, hydroxynitriles, -chiral ketones, and
acyloins with up to 99% ee were prepared from the corresponding alkanes using this protocol.
OH

CN
*
X 538
X = H, Cl, MeO
or
O Y
SAS (1.0 mM)
R1 R2 enzymes
CH3CN–H2O R1 R2 * 539
536 various reagents
two-phase system
537 e.g. NaCN, IPA, O2,
200W white light bulb Y = NH2, OH
NADH, etc.
O or
SO3Na O

541 X = H, Cl, MeO


OH
O Y = NH2, OH
sodium anthraquinone sulfonate 540
(SAS) up to 99% conversion
up to 99% ee

Scheme 101 Enantioselective C–H bond functionalization.

Recently, Hyster et al. reported an enantioselective reduction of vinyl pyridines to alkanyl


pyridines by the synergistic action of flavin-dependent ‘ene’-reductases from N. punctiforme
(NostocER) and photocatalyst Ru(bpy)3Cl2 under irradiation by blue LEDs (Scheme 102).146
Organic & Biomolecular Chemistry Page 76 of 83

76

R1 NostocER (1.0 mol %) R1


View Article Online
DOI: 10.1039/D0OB00759E
Ru(bpy)3Cl2 (0.5 mol %)
NaDP+ (1.0 mol %)

Organic & Biomolecular Chemistry Accepted Manuscript


N GDH-105, glucose (2 equiv) N
R2 R2
tricine (100 mM, PH = 9)
i 543
542 PrOH (5% v/v), 4 °C, 48 h
2,3,4-substituted blue LEDs up to 87% yield
up to 98% ee
R1 = H, OMe
R2 = Ph, 4-MePh, 4-FPh, 4-ClPh, 4-CF3Ph, 4-MeOPh, 3-MePh,
3-MeOPh, 3-CF3Ph, 2-MePh, 3-furanyl, 3-thiophenyl,
benzo[d][1,3]dioxole-5-yl, Ph(CH2)2, Bn, CH2OH
Scheme 102 Hydrogenation by the combination of enzyme and photoredox catalysis
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

12. Miscellaneous
Photocatalysis with Brønsted base
Jiang and coworkers reported the first enantioselective visible-light photocatalytic reduction of
-ketoketimines and 1,2-diketones to the corresponding chiral -amino ketones and -hydroxy
ketones using photoredox catalyst DPZ (221), a chiral organic catalyst, and reductant 2-naphthyl
tetrahydroisoquinolines (THIQ-2) (Scheme 103).147 Notably, the same 1,2-diketone photoreaction
was also feasible in the absence of photosensitizer DPZ. A visible-light-promoted SET reduction of
1,2-diketones with THIQ-2 was initiated by the photochemically active EDA complex of THIQ-2
and 1,2-diketone. Enantioselective protonation was realized using chiral Brønsted base or Brønsted
acid. The authors proposed the detailed mechanisms with and without sensitizer DPZ, and the lack-
of-sensitizer mechanism was simplified, as outlined in Scheme 103.
Page 77 of 83 Organic & Biomolecular Chemistry

77

OMe View Article Online


DOI: 10.1039/D0OB00759E
Br OMe NC N OMe
DPZ (1 mol %) S
546 (10 mol %) Br
N THIQ-2 (2.0 equiv) S

Organic & Biomolecular Chemistry Accepted Manuscript


NC N OMe
R2 HN H
3 Å MS, 2-bromoanisole
R1 R2 DPZ (221)
0 °C, 3W blue LEDs, 72 h
O R1
544 545 O
68-95% yield
N
85-90% ee
R1 = R2 = Ph, 4-FPh, 4-MePh, 4-nBuPh, 4-iPrPh, 4-MeOPh, 3-MePh, THIQ-2
3,5-Me2Ph, 2-naphthyl

DPZ (1 mol %), 549 (10 mol %)


O THIQ-2 (2.0 equiv) HO H N O CF3
R2 NaBARF (10 mol %) R2 HN
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

R1 R1
PhCl, 0 °C HN
O O
3W blue LEDs, 24 h 548 546
547 60-99% yield CF3
80-92% ee
Ph Ph
R1 = R2 = Ph, 4-FPh, 4-CF3Ph, 4-ClPh, 4-BrPh, 4-MePh, 4-nBuPh,
4-tBuPh, 4-MeOPh, 3-MePh, 3-MeOPh, 3,5-Me2Ph, 2-naphthyl, 2-thienyl N N
N N
N H H N
BF4 BF4
N 549
blue LED

O
THIQ-2
THIQ-2 R2
R1 +
O 547
color EDA complex
O
R2 HO H
R1 C3
R2
O R1
549 O 548

THIQ-2•HBF4 THIQ-2•HBF4
BF4

Ph Ph
N N BF4 THIQ-2
N N Ph Ph
N N N N
H H N N
N N
O H H
R2
R1 O
O SET R2
R1
O
THIQ-2 THIQ-2

Scheme 103 Enantioselective visible-light photocatalytic reduction of -ketoketimines and 1,2-diketones.

Photocatalysis with chiral media


The formerly elusive enantioselective [2+2] photocycloaddition of 1-bromoacenaphthylene has
been achieved by Su et al. using the strategy of cage-confined catalysis under blue LED irradiation,
max = 450 nm, (Scheme 104).148 The supramolecular photoactive catalyst was established based on
the chiral Δ/Λ-[Pd6(RuL3)8]28+ metal-organic cage (Δ/Λ-MOC-16).
Organic & Biomolecular Chemistry Page 78 of 83

78

Br Br Br View Article Online


DOI: 10.1039/D0OB00759E

10% -MOC-16 10% -MOC-16

Organic & Biomolecular Chemistry Accepted Manuscript


blue LED, 2 h blue LED, 2 h Br Br
550 551
90% conversion, 88% ee 92% conversion, 86% ee

2+ 2+
N
N

HN
N
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

N
NH

N
N
H N N N
N N N N Ru
Ru
N N N N
N N N H N
N

NH
N
N
HN
-552 -552
N
N Pd(BF4)2
Pd(BF4)2

-MOC-16 -MOC-16

Scheme 104 Eantioselective [2+2]cycloaddition with metal-organic cage-confine supramolecular catalysis

13. Conclusion

Photoreaction, especially enabled by UV irradiation, has been a compelling subject in organic


chemistry for more than a century. It was not until almost half a century ago, however, that chemists
began to overcome the complicated stereoselectivity of photochemical reactions and succeed in the
effective synthesis of elaborate organic molecules with multiple stereocenters. The difficulties with
this enantioselective photoreaction stem from a limited understanding of the higher energy excited
states of UV-irradiated substrates and the scarcity of effective enantioselective catalysis. With the
most recent progress in asymmetric catalysis and the emergence of visible-light photocatalysis, a
number of new methods have arisen to fill the gaps in this field. The enantioselective synthesis
enabled by visible-light photocatalysis clearly has been advancing significantly over the last decade
or so. In this review, we have summarized a great deal of this substantial progress.
Despite these rapid advances, some challenges remain to be tackled, and certain objectives can
be anticipated and accomplished, as follows. First is the development of more effective catalysts with
Page 79 of 83 Organic & Biomolecular Chemistry

79

a lesser quantity requirement, along with improved chemoselectivity and increased enantioselectivity View Article Online
DOI: 10.1039/D0OB00759E
of visible-light photocatalysis reactions. With the discovery of new photocatalysts and an improved
understanding of photocatalytic mechanisms, a second objective is the development of more

Organic & Biomolecular Chemistry Accepted Manuscript


advanced methods that further enrich the field. As a third objective, future strategies in this field could
involve the cascade reaction or one-pot operation of multiple bonds and chiral centers with
enantioselectivity by visible-light photocatalysis. New multicomponent visible-light photocatalysis
for enantioselective reactions should be a fourth objective. A fifth aim should be a better
understanding of the photophysical properties of reactants, such as transient electron donor–acceptor
complexes, as a step toward discovering more photocatalyst-free enantioselective photocatalysis
reactions. Finally, with the development of more effective asymmetric catalysis that is
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

environmentally friendly and convenient to operate, or of a continuous flow process, further


innovations can become possible that integrate this technology with visible-light photocatalytic
reactions.
In short, enantioselective photocatalysis will undoubtedly continue to expand and thrive in
decades to come while offering benefits for advanced synthetic chemistry.

Abbreviation list

CFL compact fluorescent light


CAN cerium ammonium nitrate
CPA chiral phosphoric acid
DIPEA diisopropylethylamine
DPZ A chromophore (DPZ) photosensitizer derived from dicyanopyrazine.
Usually refers to 5,6-bis (5-methoxythien-2-yl) pyrazine-2,3-dicarbonitrile.
EDA electron donor-acceptor complex
HAT hydrogen atom transfer
HE Hantzsch ester. It refers to a series of dialkyl 1,4-dihydro-2,6-dialkyl-3,5-
pyridinedicarboxylate. Unless otherwise indicated with different substitutes or
esters, it refers to diethyl 1,4-dihydro-2,6-dimethyl-3,5-pyridinedicarboxylate.
ISC intersystem crossing
LED light emitting diode
MTBE methyl tert-butyl ether
NHC N-Heterocyclic carbenes
NMA N-methyl acetamide
PC photocatalysis
SET single electron transfer
SOMO singly occupied molecular orbital
SPINOL Refers to a series of axially chiral 1,1′-spirobiindane-7,7′-diols
SPINOL-CPA SPINOL-derived chiral phosphoric acids
Organic & Biomolecular Chemistry Page 80 of 83

80

TFA trifluoroacetic acid View Article Online


DOI: 10.1039/D0OB00759E

Organic & Biomolecular Chemistry Accepted Manuscript


Conflicts of interest
There are no conflicts to declare.

Acknowledgments

Financial support from the Ministry of Science and Technology (MOST, Taiwan), is gratefully
acknowledged. We thank Simon Partridge, Ph.D., from Edanz Group for editing the English text of
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

a draft of this manuscript.

References

1 For reviews, see: (a) M. Silvi, P. Melchiorre, Nature, 2018, 554, 41–49. (b) G. E. M. Crisenza, D. Mazzarella, and P.
Melchiorre, J. Am. Chem. Soc., 2020, 142, 5461–5476.
2 (a) E. Meggers, Chem. Commun., 2015, 51, 3290–3301. (b) C. Wang, and Z. Lu, Org. Chem. Front., 2015, 2, 179–

190. (c) C. Jiang, W. Chen, W.-H. Zheng, and H. Lu, Org. Biomol. Chem., 2019, 17, 8673–8689.
3 S. Li, S.-H. Xiang, B. Tan, Chin. J. Chem. 2020, 38, 213–214.
4 T. P. Yoon, Acc. Chem. Res., 2016, 49, 2307–2315.
5 Y. Xu, M. L. Conner, and M. K. Brown, Angew. Chem. Int. Ed., 2015, 54, 11918–11928.
6 (a) E. Meggers, Angew. Chem. Int. Ed., 2017, 56, 5668–5675. (b) A. G. Amador, and T. P. Yoon, Angew. Chem. Int.

Ed., 2016, 55, 2304–2306. (c) J. Ma, X. Zhang, X. Huang, S. Luo, E. Meggers, Nat. Protoc., 2018, 13, 605–632.
7 C. Brenninger, J. D. Jolliffe, and T. Bach, Angew. Chem. Int. Ed., 2018, 57, 14338–14349.
8 R. Brimioulle, D. Lenhart, M. M. Maturi, and T. Bach, Angew. Chem. Int. Ed., 2015, 54, 3872–3890.
9 Y.-Q. Zou, F. M. Hoörmann, and T. Bach, Chem. Soc. Rev., 2018, 47, 278–290.
10 X. Huang, and E. Meggers, Acc. Chem. Res., 2019, 52, 833−847.
11 Q. Shi, and J. Ye, Angew. Chem. Int. Ed., 2020, 59, 4998–5001.
12 B. A. Sandoval, and T. K. Hyster Curr. Opin. Chem. Biol., 2020, 55, 45–51.
13 D. A. Nicewicz, D. W. C. MacMillan, Science, 2008, 322, 77–80.
14 D. A. Nagib, M. E. Scott, and D. W. C. MacMillan, J. Am. Chem. Soc., 2009, 131, 10875–10877.
15 G. Cecere, C. M. König, J. L. Alleva, and D. W. C. MacMillan, J. Am. Chem. Soc., 2013, 135, 11521−11524.
16 H.-W. Shih, M. N. Vander Wal, R. L. Grange, D. W. C. MacMillan, J. Am. Chem. Soc., 2010, 132, 13600–13603.
17 M. Neumann, S. Füldner, B. König, K. Zeitler, Angew. Chem. Int. Ed., 2011, 50, 951–954.
18 M. Neumann, K. Zeitler, Org. Lett., 2012, 14, 2658–2661.
19 M. Cherevatskaya, M. Neumann, S. Füldner, C. Harlander, S. Kümmel, S. Dankesreiter, A. Pfitzner, K. Zeitler, and

B. König, Angew. Chem. Int. Ed., 2012, 51, 4062.


20 B. Schweitzer-Chaput, M. A. Horwitz, E. de P. Beato, and P. Melchiorre, Nat. Chem., 11, 2019, 129–135
21 Y. Luo, Z.-Y. Xu, H. Wang, X.-W. Sun, Z.-T. Li, and D.-W. Zhang, ACS Macro Lett., 2020, 9, 90−95.
22 X. Kang, X.Wu, X. Han, C. Yuan, Y. Liu, and Y. Cui, Chem. Sci., 2020, 11, 1494–1502.
23 H.-S. Yoon, Z.-H. Ho, J. Jang, H.-J. Lee, S.-J. Kim, H.-Y. Jang, Org. Lett., 2012, 14, 3272–3275.
24 E. Arceo, I. D. Jurberg, A. Álvarez-Fernández, P. Melchiorre, Nat. Chem., 2013, 5, 750 – 756.
25 P. Riente, A. M. Adams, J. Albero, E. Palomares and M. A. Pericàs, Angew. Chem. Int. Ed., 2014, 53, 9613 –9616.
26 E. Arceo, A. Bahamonde, G. Bergonzini, P. Melchiorre Chem. Sci., 2014, 5, 2438–2442
27 M. Silvi, E. Arceo, I. D. Jurberg, C. Cassani, P. Melchiorre, J. Am. Chem. Soc, 2015, 137, 6120−6123.
28 A. Bahamonde, P. Melchiorre, J. Am. Chem. Soc., 2016, 138, 8019−8030.
29 A. Gualandi, M. Marchini, L. Mengozzi, M. Natali, M. Lucarini, P. Ceroni, and P. G. Cozzi, ACS Catal., 2015, 5,

5927−5931.
30 M. T. Pirnot, D. A. Rankic, D. B. C. Martin, D. W. C. MacMillan, Science, 2013, 339, 1593–1596.
31 Y. Zhu, L. Zhang, and S. Luo, J. Am. Chem. Soc., 2014, 136, 14642−14645.
32 W.-Z. Zhang, Y. Zhu, L. Zhang, and S. Luo, Chin. J. Chem., 2018, 36, 716-722.
33 E. Larionov, M. M. Mastandrea, and M. A. Pericàs, ACS Catal., 2017, 7, 7008−7013.
34 D. Wang, L. Zhang, S. Luo, Org. Lett., 2017, 19, 4924−4927.
35 M. Silvi, C. Verrier, Y. P. Rey, L. Buzzetti, and P. Melchiorre, Nat. Chem. 2017, 9, 868-873.
Page 81 of 83 Organic & Biomolecular Chemistry

81

View Article Online


DOI: 10.1039/D0OB00759E
36 C. Verrier, N. Alandini, C. Pezzetta, M. Moliterno, L. Buzzetti, Ha. B. Hepburn, A. Vega-Peñaloza, M. Silvi, and P.
Melchiorre, ACS Catal., 2018, 8, 1062−1066.
37 A. G. Capacci, J. T. Malinowski, N. J. McAlpine, J. Kuhne, and D. W. C. MacMillan, Nat. Chem. 2017, 9, 1073-

Organic & Biomolecular Chemistry Accepted Manuscript


1077.
38 F. M. Hörmann, T. S. Chung, E. Rodriguez, M. Jakob, and T. Bach, Angew. Chem. Int. Ed., 2018, 57, 827 –831
39 S. Stegbauer, C. Jandl, and T. Bach, Angew. Chem. Int. Ed., 2018, 57, 14593–14596.
40 Z.-Y. Cao, T. Ghosh, and P. Melchiorre, Nat. Commun., 2018, 9, 3274.
41 P. Bonilla, Y. P. Rey, C. M. Holden, and P. Melchiorre, Angew. Chem. Int. Ed., 2018, 57, 12819–12823.
42 J. Rostoll-Berenguer, G. Blay, M. C. Muñoz, J. R. Pedro, and C. Vila, Org. Lett., 2019, 21, 6011−6015.
43 C.-L. Dong, X. Ding, L.-Q. Huang, Y.-H. He, and Z. Guan, Org. Lett., 2020, 22, 1076–1080.
44 L. A. Perego, P. Bonilla, and P. Melchiorre, Adv. Synth. Catal., 2020, 362, 302 – 307.
45 D. Spinnato, B. Schweitzer-Chaput, G. Goti, M. Ošeka, and P. Melchiorre, Angew. Chem. Int. Ed., 2020, xxxx.

10.1002/anie.201915814.
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

46 J. Du, K. L. Skubi, D. M. Schultz, T. P. Yoon, Science, 2014, 344, 392–396.


47 (a) C. Müller, A. Bauer, T. Bach, Angew. Chem. Int. Ed., 2009, 48, 6640–6642; (b) M. M. Maturi, M. Wenninger, R.

Alonso, A. Bauer, A. Pöthig, E. Riedle, T. Bach, Chem. Eur. J., 2013, 19, 7461–7472; (c) C. Müller, A. Bauer, M. M.
Maturi, C. Cuquerella, M. A. Miranda, T. Bach, J. Am. Chem. Soc., 2011, 133, 16689–16697; (d) H. Guo, E.
Herdtweck, T. Bach, Angew. Chem. Int. Ed., 2010, 49, 7782–7785. (e) R. Brimioulle, T. Bach, Science, 2013, 342, 840-
843.
48 (a) T. R. Blum, Z. D. Miller, D. M. Bates, I. A. Guzei, T. P. Yoon, Science, 2016, 354, 1391–1395.
49 Z. D. Miller, B. J. Lee, and T. P. Yoon, Angew. Chem. Int. Ed., 2017, 56, 11891–11895.
50 M. E. Daub, H. Jung, B. J. Lee, J. Won, M.-H. Baik, and T. P. Yoon, J. Am. Chem. Soc., 2019, 141, 9543−9547.
51 L. R. Espelt, I. S. McPherson, E. M. Wiensch, T. P. Yoon, J. Am. Chem. Soc., 2015, 137, 2452−2455.
52 C.-X. Ye, Y. Y. Melcamu, H.-H. Li, J.-T. Cheng, T.-T. Zhang, Y.-P. Ruan, X. Zheng, X. Lu, P.-Q. Huang, Nat.

Commun., 2018, 9, 410.


53 H. Liang, G.-Q. Xu, Z.-T. Feng, Z.-Y. Wang, and P.-F. Xu, J. Org. Chem., 2019, 84, 60−72.
54 F. Li, D. Tian, Y. Fan, R. Lee, G. Lu, Y. Yin, B. Qiao, X. Zhao, Z. Xiao, Z. Jiang, Nat. Commun., 2019, 10, 1774.
55 J. N. Arokianathar, A. B. Frost, A. M. Z. Slawin, D. Stead, and A. D. Smith, ACS Catal., 2018, 8, 1153−1160.
56 Ł. Woźniak, J. J. Murphy, P. Melchiorre, J. Am. Chem. Soc. 2015, 137, 5678−5681.
57 L. J. Rono, H. G. Yayla, D. Y. Wang, M. F. Armstrong, R. R. Knowles, J. Am. Chem. Soc., 2013, 135, 17735−17738.
58 For recent review: P.-Z. Wang, J.-R. Chen, W.-J. Xiao, Org. Biomol. Chem., 2019, 17, 6936–6951.
59 D. Uraguchi, N. Kinoshita, T. Kizu, T. Ooi, J. Am. Chem. Soc. 2015, 137, 13768−13771.
60 Z. Yang, H. Li, S. Li, M.-T. Zhang, and S. Luo. Org. Chem. Front., 2017, 4, 1037–1041.
61 J. Li, M. Kong, B. Qiao, R. Lee, X. Zhao, Z. Jiang, Nat. Commun., 2018, 9, 2445.
62 X. Liu, Y. Liu, G. Chai, B. Qiao, X. Zhao, and Z. Jiang, Org. Lett., 2018, 20, 6298−6301.
63 L. Bu, J. Li, Y. Yin, B. Qiao, G. Chai, X. Zhao, and Z. Jiang, Chem. Asian J. 2018, 13, 2382–2387.
64 K. Liang, X. Tong, T. Li, B. Shi, H. Wang, P. Yan, C. Xia, J. Org. Chem., 2018, 83, 10948−10958.
65 Y. Liu, X. Liu, J. Li, X. Zhao, B. Qiao, and Z. Jiang, Chem. Sci., 2018, 9, 8094–8098.
66 Y. Yin, Y. Dai, H. Jia, J. Li, L. Bu, B. Qiao, X. Zhao, and Z. Jiang, J. Am. Chem. Soc., 2018, 140, 6083−6087.
67 T. Shao, Y. Li, N. Ma, C. Li, G. Chai, X. Zhao, B. Qiao, and Z. Jiang, iScience, 2019, 16, 410–419.
68 B. Qiao, C. Li, X. Zhao, Y. Yin, and Z. Jiang, Chem. Commun., 2019, 55, 7534–7537.
69 M. Hou, L. Lin, X. Chai, X. Zhao, B. Qiao, and Z. Jiang, Chem. Sci., 2019, 10, 6629–6634.
70 R. S. J. Proctor, H. J. Davis, and R. J. Phipps, Science, 2018, 360, 419–422.
71 K. Cao, S. M. Tan, R. Lee, S. Yang, H. Jia, X. Zhao, B. Qiao, and Z. Jiang, J. Am. Chem. Soc., 2019, 141,

5437−5443.
72 W. Xiong, S. Li, B. Fu, J. Wang, Q.-A. Wang, and W. Yang, Org. Lett., 2019, 21, 4173−4176.
73 C. B. Roos, J. Demaerel, D. E. Graff, R. Knowles, J. Am. Chem. Soc. 2020, 142, 5974–5979.
74 For a review, see: L. Zhang, and E. Meggers, Chem. Asian J., 2017, 12, 2335–2342.
75 H. Huo, X. Shen, C. Wang, L. Zhang, P. Röse, L.-A. Chen, K. Harms, M. Marsch, G. Hilt, E. Meggers, Nature, 2014,

515, 100–103.
76 C. Wang, Y. Zheng, H. Huo, P. Rcse, L. Zhang, K. Harms, G. Hilt, E. Meggers, Chem. Eur. J. 2015, 21, 7355–7359.
77 H. Huo, C. Wang, K. Harms, E. Meggers, J. Am. Chem. Soc., 2015, 137, 9551−9554.
78 Y. Tan, W. Yuan, L. Gong, E. Meggers, E. Angew. Chem. Int. Ed. 2015, 54, 13045–13048.
79 H. Huo, K. Harms, E. Meggers, J. Am. Chem. Soc., 2016, 138, 6936−6939.
80 C. Wang, J. Qin, X. Shen, R. Riedel, K. Harms, and E. Meggers, Angew. Chem. Int. Ed., 2016, 55, 685–688.
81 J. Ma, K. Harms, and E. Meggers, Chem. Commun., 2016, 52, 10183–10186.
82 X. Shen, K. Harms, M. Marsch, and E. Meggers Chem. Eur. J., 2016, 22, 9102 – 9105.
83 H. Huo, X. Huang, X. Shen, K. Harms, E. Meggers, Synlett, 2016, 27, 749–753.
84 G.-Q. Xu, H. Liang, J. Fang, Z.-L. Jia, J.-Q. Chen and P.-F. Xu, Chem. Asian J. 2016, 11, 3355–3358.
Organic & Biomolecular Chemistry Page 82 of 83

82

View Article Online


DOI: 10.1039/D0OB00759E
85 S.-X. Lin, G.-J. Sun, and Q. Kang Chem. Commun., 2017, 53, 7665–7668.
86 G. J. Choi, Q. Zhu, D. C. Miller, C. J. Gu, and R. R. Knowles, Nature, 2016, 539, 268–271.
87 Z. Zhou, Y. Li, B. Han, L. Gong, and E. Meggers, Chem. Sci., 2017, 8, 5757–5763.

Organic & Biomolecular Chemistry Accepted Manuscript


88 X. Huang, S. Luo, O. Burghaus, R. D. Webster, K. Harms, and E. Meggers, Chem. Sci., 2017, 8, 7126–7131
89 T. Deng, G. K. Thota, Y. Li, and Q. Kang, Org. Chem. Front., 2017, 4, 573–577.
90 S.-W. Li, J. Gong, and Q. Kang, Org. Lett., 2017, 19, 1350−1353.
91 F. F. de Assis, X. Huang, M. Akiyama, R. A. Pilli, and E. Meggers, J. Org. Chem., 2018, 83, 10922−10932.
92 J. Ma, A. R. Rosales, X. Huang, K. Harms, R. Riedel, O. Wiest, and E. Meggers, J. Am. Chem. Soc., 2017, 139,

17245−17248.
93 X. Shen, H. Huo, C. Wang, B. Zhang, K. Harms, and E. Meggers, Chem. Eur. J., 2015, 21, 9720–9726.
94 X. Zhang, J. Qin, X. Huang, and E. Meggers, Org. Chem. Front., 2018, 5, 166–170.
95 X. Zhang, J. Qin, X. Huang, and E. Meggers, Eur. J. Org. Chem. 2018, 571–577.
96 X. Huang, T. R. Quinn, K. Harms, R. D. Webster, L. Zhang, O.Wiest, and E. Meggers, J. Am. Chem. Soc., 2017, 139,
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

9120−9123.
97 R. Alonso and T. Bach, Angew. Chem. Int. Ed. 2014, 53, 4368–4371.
98 A. Tröster, R. Alonso, A and Bauer, T. Bach, J. Am. Chem. Soc. 2016, 138, 7808−7811.
99 M. M. Maturi and T. Bach, Angew. Chem. Int. Ed., 2014, 53, 7661 – 7664.
100 N.Vallavoju, S. Selvakumar, S. Jockusch, M. P. Sibi, J. Sivaguru, Angew. Chem. Int. Ed. 2014, 53, 5604 –5608.
101 K. L. Skubi, J. B. Kidd, H. Jung, I. A. Guzei, M.-H. Baik, and T. P. Yoon, J. Am. Chem. Soc., 2017, 139,

17186−17192.
102 P. D. Morse, T. M. Nguyen, C. L. Cruz, and D. A. Nicewicz, Tetrahedron, 2018, 74, 3266–3272.
103 X.-F.Tang , S.-H. Feng, Y.-K. Wang, F. Yang, Z.-H. Zheng, J.-N. Zhao, Y.-F. Wu, H. Yin, G.-Z. Liu, Q.-W. Meng

Tetrahedron, 2018, 74, 3624–3633.


104 X.-F. Tang, J.-N. Zhao, Y.-F. Wu, S.-H. Feng, F. Yang ,Z.-Y. Yu, and Q.-W. Meng, Adv Synth Catal., 2019, 361,

5245–5252.
105 J. Ma, J. Lin, L. Zhao, K. Harms, M. Marsch, X. Xie, and E. Meggers, Angew Chem Int Ed, 2018, 57, 11193–11197.
106 Y. Li, K. Zhou, Z. Wen, S. Cao, X. Shen, M. Lei, and L. Gong, J. Am. Chem. Soc., 2018, 140, 15850−15858.
107 X. Huang, J. Lin, T. Shen, K. Harms, M. Marchini, P. Ceroni, and E. Meggers, Angew. Chem. Int. Ed., 2018, 57,

5454–5458.
108 G. Bergonzini, C. S. Schindler, C.-J. Wallentin, E. N. Jacobsen and C. R. J. Stephenson, Chem. Sci., 2014, 5, 112–

116.
109 For reviews, see: (a) D. Qian and J. Sun Chem. Eur. J., 2019, 25, 3740–3751; (b) K. Brak and E. N. Jacobsen,

Angew. Chem., Int. Ed., 2013, 52, 534–561; (c) R. J. Phipps, G. L. Hamilton and F. D. Toste, Nat. Chem., 2012, 4, 603–
614.
110 J. C. Tellis, D. N. Primer, G. A. Molander, Science, 2014, 345, 433–436.
111 For a mini-review of , see: Photoredox/Nickel Dual Catalysis, see: J. C. Tellis, C. B. Kelly, D. N. Primer, M.

Jouffroy, N. R. Patel, and G. A. Molander, Acc. Chem. Res., 2016, 49, 1429−1439.
112 O. Gutierrez, J. C. Tellis, D. N. Primer, G. A. Molander, and M. C. Kozlowski, J. Am. Chem. Soc., 2015, 137,

4896−4899.
113 Z. Zuo, H. Cong, W. Li, J. Choi, G. C. Fu, and D. W. C. MacMillan, J. Am. Chem. Soc., 2016, 138, 1832−1835.
114 I. Perepichka, S. Kundu, Z. Hearne, and C.-J. Li, Org. Biomol. Chem., 2015, 13, 447–451.
115 W. Ding, L.-Q. Lu, Q.-Q. Zhou, Y. Wei, J.-R. Chen, and W.-J. Xiao, J. Am. Chem. Soc., 2017, 139, 63−66.
116 J. Liu, W. Ding, Q.-Q. Zhou, D. Liu, L.-Q. Lu, and W.-J. Xiao, Org. Lett., 2018, 20, 461−464.
117 D. Wang, N. Zhu, P. Chen, Z. Lin, and G. Liu, J. Am. Chem. Soc., 2017, 139, 15632−15635.
118 M.-M. Li, Y. Wei, J. Liu, H.-W. Chen, L.-Q. Lu, and W.-J. Xiao, J. Am. Chem. Soc., 2017, 139, 14707−14713.
119 Q.-Q. Zhou, F.-D. Lu, D. Liu, L.-Q. Lu, and W.-J. Xiao, Org. Chem. Front., 2018, 5, 3098–3102.
120 W. Sha, L. Deng, S. Ni, H. Mei, J. Han, and Y. Pan, ACS Catal. 2018, 8, 7489−7494.
121 H. Yu, S. Dong, Q. Yao, L. Chen, D. Zhang, X. Liu, and X. Feng, Chem. Eur. J., 2018, 24, 19361–19367.
122 X. Shen, Y. Li, Z. Wen, S. Cao, X. Hou, and L. G. Chem. Sci., 2018, 9, 4562–4568.
123 H.-H. Zhang, J.-J. Zhao, and S. Yu, J. Am. Chem. Soc., 2018, 140, 16914−16919.
124 A. Casado-Sánchez, P. Domingo-Legarda, S. Cabrera, and J. Alemán, Chem. Commun., 2019, 55, 11303–11306.
125 Q. Guo, M. Wang, Q. Peng, Y. Huo, Q. Liu, R. Wang, and Z. Xu, ACS Catal., 2019, 9, 4470−4476.
126 H. Mitsunuma, S. Tanabe, H. Fuse, K. Ohkubo, and M. Kanai, Chem. Sci., 2019, 10, 3459–3465.
127 L. Pitzer, F. Schäfers, and F. Glorius, Angew. Chem. Int. Ed., 2019, 58, 8572 – 8576.
128 K. Zhang, L.-Q. Lu, Y. Jia, Y. Wang, F.-D. Lu, F. Pan, and W.-J. Xiao, Angew. Chem. Int. Ed., 2019, 58, 13375-

13379.
129 B. Han, Y. Li, Y. Yu, L. Gong, Nat. Commun., 2019, 10, 3804.
130 Schwarz, J. L.; Huang, H.-M., Paulisch, T. O.; Glorius, F. ACS Catal., 2020, 10, 1621–1627.
131 Y. Zhang, Y. Sun, B. Chen, M. Xu, C. Li, D. Zhang, and G. Zhang, Org. Lett., 2020, 22, 1490–1494.
Page 83 of 83 Organic & Biomolecular Chemistry

83

View Article Online


DOI: 10.1039/D0OB00759E
132 P. Fan, Y. Lan, C. Zhang, and C. Wang J. Am. Chem. Soc. 2020, 142, 2180-2186.
133 Daniel A. DiRocco and Tomislav Rovis, J. Am. Chem. Soc., 2012, 134, 8094−8097.
134 L. Dai, Z.-H. Xia, Y.-Y. Gao, Z.-H. Gao*, and S. Ye, Angew. Chem. Int. Ed., 2019, 58, 18124-18130.

Organic & Biomolecular Chemistry Accepted Manuscript


135 K. Lauder, and D. Castagnolo, Synlett, 2020, 31, 737-744.
136 M. A. Emmanuel, N. R. Greenberg, D. G. Oblinsky, and T. K. Hyster, Nature, 2016, 540, 414-417.
137 (a) K. F. Biegasiewicz, S. J. Cooper, M. A. Emmanuel, D. C. Miller, and T. K. Hyster Nat. Chem. 2018, 10, 770–

775. (b) B. A. Sandoval, S. I. Kurtoic, M. M. Chung, K. F. Biegasiewicz, T. K. Hyster, Angew. Chem. Int. Ed. 2019,
58, 8714-8718. (c) K. F. Biegasiewicz, S. J. Cooper, X. Gao, D. G. Oblinsky, J. H. Kim, S. E. Garfinkle, L. A. Joyce, B.
A. Sandoval, G. D. Scholes, T. K. Hyster, Science, 2019, 364, 1166–1169. (d) M. J. Black, K. F. Biegasiewicz, A. J.
Meichan, D. G. Oblinsky, B. Kudisch, G. D. Scholes, and T. K. Hyster, Nat. Chem., 2020, 12, 71–75.
138 S. Choudhury, J.-O. Baeg, N.-J. Park and R. K. Yadav, Green Chem., 2014, 16, 4389–4400.
139 M. Rauch, S. Schmidt, I. W. C. E. Arends, K. Oppelt, S. Kara and F. Hollmann Green Chem., 2017, 19, 376–379
140 S. H. Lee, D. S. Choi, M. Pesic, Y. W. Lee, C. E. Paul, F. Hollmann, and C. B. Park, Angew. Chem. Int. Ed., 2017,
Published on 21 May 2020. Downloaded by Uppsala University on 5/21/2020 5:45:44 PM.

56, 8681–8685.
141 Q. Yang, F. Zhao, N. Zhang, M. Liu, H. Hu, J. Zhang, and S. Zhou, Chem. Commun., 2018, 54, 14065–14068.
142 X. Guo, Y. Okamoto, M. R. Schreier, T. R. Ward, and O. S. Wenger, Chem. Sci., 2018, 9, 5052–5056.
143 J. Yoon, S. H. Lee, F. Tieves, M. Rauch, F. Hollmann, and C. B. Park, ACS Sustainable Chem. Eng., 2019, 7,

5632−5637.
144 X. Ding, C.-L. Dong, Z. Guan, and Y.-H. He, Angew. Chem. Int. Ed., 2019, 58, 118–124.
145 W. Zhang, E. F. Fueyo, F. Hollmann, L. L. Martin, M. Pesic, R. Wardenga, M. Höhne, and S. Schmidt, Eur. J. Org.

Chem., 2019, 80–84.


146 Y. Nakano, M. J. Black, A. J. Meichan, B. A. Sandoval, M. M. Chung, K. F. Biegasiewicz, T. Zhu, T. K. Hyster,

Angew. Chem. Int. Ed., 2020, 10.1002/anie.202003125.


147 L. Lin, X. Bai, X. Ye, X. Zhao, C.-H. Tan, and Z. Jiang, Angew. Chem. Int. Ed., 2017, 56, 13842–13846.
148 J. Guo, Y.-Z. Fan, Y.-L. Lu, S.-P. Zheng, C.-Y. Su, Angew Chem. Int. Ed., 2020, 59, 10.1002/ange.201916722

You might also like