You are on page 1of 6

pubs.acs.

org/JACS Communication

Helical Carbenium Ion: A Versatile Organic Photoredox Catalyst for


Red-Light-Mediated Reactions
Liangyong Mei, Jose ́ M. Veleta, and Thomas L. Gianetti*
Cite This: J. Am. Chem. Soc. 2020, 142, 12056−12061 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Red light has the advantages of low energy, less health risks, and high penetration depth through various media.
Herein, a helical carbenium ion (N,N′-di-n-propyl-1,13-dimethoxyquinacridinium (nPr-DMQA+) tetrafluoroborate) has been used as
an organic photoredox catalyst for photoreductions and photooxidations in the presence of red light (λmax = 640 nm). It has
Downloaded via UNIV FED FLUMINENSE on June 23, 2021 at 02:31:50 (UTC).

catalyzed red-light-mediated dual transition-metal/photo-redox-catalyzed C−H arylation and intermolecular atom-transfer radical
addition through oxidative quenching. Moreover, its potential in photooxidation catalysis has also been demonstrated by successful
applications in red-light-induced aerobic oxidative hydroxylation of arylboronic acids and benzylic C(sp3)−H oxygenation through
reductive quenching. Thus, a versatile organic photoredox catalyst (helical carbenium ion) for red-light-mediated photoredox
reactions has been developed.

O ver the past decade, photoredox catalysis has taken the


world of synthetic chemistry by a storm.1 Iridium and
ruthenium polypyridyl complexes such as 1a2 and 1b3 are
high-energy blue, green, or white light.1,4 Besides the health
hazards (photooxidative damage to the retina),8 blue light and
green light are also more prone to induce undesired mixtures
among the most powerful photoredox catalysts (PCs) due to due to the possible photon absorptions by reactants or
their unique properties such as stable and long-lived excited products.9 In contrast, red light (600−700 nm) has advantages
states and large redox windows, and they are effective as such as lower energy, fewer side reactions, and less health risks,
excited-state oxidants and reductants (Figure 1). However, and it is naturally abundant from sunlight. More importantly, it
penetrates turbid media, and great successes have been made
in its biology-related applications.10 Yet reports on red-light-
mediated synthetic methodologies still remain scarce, and only
a handful of examples can be found (Scheme 1a). Ferroud and
co-workers introduced red-light-induced Methylene Blue (2a)-
catalyzed aerobic photooxidation and photocyanation of
hydrazines (Scheme 1a, I).11 Though more applications of
2a in photocatalysis have been reported, they all employ blue,
green, or white light.4,12 Similarly, PCs 2b−2d have also
achieved some successes in red-light-mediated methodologies,
Figure 1. Commonly used photoredox catalysts. while no further applications were investigated (Scheme 1a,
II−IV). 13 Notably, Rovis and Campos have recently
introduced an elegant work on near-infrared light-induced
given the intrinsic disadvantages of transition-metal catalysts transformations in the absence of a PC, or in the presence of
such as high cost, low sustainability, and potential toxicity, 1b, 1d, or Rose Bengal via triplet fusion upconversion,14
researchers have been looking for inexpensive and environ- followed by the development of a series of Os(II)-based PCs
ment-friendly alternatives. In recent years, organic molecules for infrared photoredox catalysis.15 These important examples
such as acridiniums (e.g., 9-mesityl-10-methylacridinium 1c) all suggest that applying red light in photoredox catalysis is of
or xanthene dyes (e.g., Eosin Y, 1d) have proven to be efficient great significance.
alternatives (Figure 1).4 Despite the general merits of organic The helical carbenium ion, dimethoxyquinacridinium
PCs, there are disadvantages to overcome. For example, the (DMQA+), has been well-studied regarding its photophysical16
most widely used catalyst in this category, 1c, first introduced
by Fukuzumi5 and expanded by Nicewicz,4a,b,6 is mainly used
for reductive quenching. Similarly, the narrow redox window, Received: May 19, 2020
pH dependence, and susceptibility to bleaching have also Published: June 30, 2020
rendered 1d less effective.7 Therefore, developing more
versatile organic PCs is still highly desirable.
Despite the impressive progress in photoredox catalysis to
date, most light-induced reactions are accomplished under

© 2020 American Chemical Society https://dx.doi.org/10.1021/jacs.0c05507


12056 J. Am. Chem. Soc. 2020, 142, 12056−12061
Journal of the American Chemical Society pubs.acs.org/JACS Communication

Scheme 1. (a) Previous Work on Red-Light-Mediated particular, the peak absorption (λmax) around 620 nm16b makes
Methodologies and (b) a Versatile Organic PC in This our goal of taking advantage of low-energy red light feasible.
Work Initial reactivity studies began with a red-light-mediated dual
Pd/nPr-DMQA+-catalyzed C(sp2)−H arylation, which was first
reported by the Sanford group using Ru(bpy)32+ in the
presence of 26 W compact fluorescent light (CFL).21 In the
reported reaction, Ru(bpy)32+ (E1/2(Ru(III)/Ru(II*)) = −0.81
V vs SCE) acted as a photoreductant by reducing ArN2+ to aryl
radical (PhN2+/Ph• = −0.10 V vs SCE)22 through oxidative
quenching, followed by oxidation of Pd(III) to Pd(IV) to
regenerate Ru(bpy)32+ (E1/2(Ru(III)/Ru(II)) = +1.29 V vs
SCE). Given E1/2 (C•++/C+*) = −0.62 V and E1/2(C•++/C+) =
+1.32 V for 3 in MeOH, this reaction was deemed suitable to
display its photoreduction ability. To our delight, the reaction
between 1-([1,1′-biphenyl]-2-yl)pyrrolidin-2-one 4a and
benzenediazonium tetrafluoroborate 5a proceeded smoothly
in the presence of Pd(OAc)2 and 3 under red LED (λmax = 640
nm), affording the desired product 6a in 95% NMR yield after
4 h (Table S2). Consistent with Sanford’s work,21 in the
absence of 3, red light, or Pd(OAc)2, significantly lower yields
(≤25%) of 6a were observed (Table S2).
With these promising results in hand, we sought to expand
the substrate scope of this red-light-mediated dual Pd/nPr-
DMQA+-catalyzed C(sp2)−H arylation (Table 1). Using 4a as
a model substrate, electron-neutral, electron-deficient, and
electron-rich aryldiazonium salts 5a−5c were tested. Reactions
and electrochemical properties16b,17 and post-functionaliza-
tion16b,18 since its discovery by Laursen and co-workers.16a Table 1. Red-Light-Mediated Dual Pd/nPr-DMQA+-
Applications of DMQA+ have been primarily reserved for the Catalyzed C(sp2)−H Arylationa
biological arena.19 In 2005, the Lacour group revealed the first
and only example of photocatalysis using DMQA+, which
demonstrated the aerobic photooxidation of benzylamine in
the presence of N,N′-di-n-propyl-1,13-dimethoxy-
quinacridinium (nPr-DMQA+) tetrafluoroborate (3) under
photoirradiation (600 W lamp) to afford benzylimine in 15%
NMR yield at 70 °C.20 Based on the reported extraordinary
photophysical and electrochemical properties of 3, as well as
Lacour’s results on photocatalysis, we speculated that 3 could
serve as a versatile organic PC. Herein, we report the
photoredox properties of [nPr-DMQA+][BF4−] (3) and its
reactivity in a wide range of red-light-mediated reactions,
including reductive and oxidative quenching (Scheme 1b).
Before testing the photocatalytic reactivity of 3, we first
calculated the excitation energy (E0,0) and the excited-state
oxidation (E1/2 (C•++ /C+ *)) and reduction potentials
(E1/2(C+*/C•)) following literature protocols (Table S1, see
details in the Supporting Information).4a To our delight, 3
possesses moderate ground-state oxidation and reduction
potentials E1/2(C•++/C+) = +1.32 V and E1/2(C+/C•) =
−0.78 V vs the saturated calomel electrode (SCE), as well as
moderate excited-state oxidation and reduction potentials E1/2
(C•++/C+*) = −0.61 V and E1/2(C+*/C•) = +1.15 V in MeCN
(Table S1). Solvent effects on the properties of 3 were
examined by recording its absorption and emission spectra, as
well as cyclic voltammetric (CV) curves in four different
solvents (MeCN, DCM, DMF, and MeOH, Figures S1 and
S2). No significant solvatochromism or redox potential
differences were observed in these solvents (Table S1).
These features render 3 a mild reductant and oxidant whether
a reductive or oxidative quenching is involved during
photocatalysis. Moreover, the excited-state lifetime (τ = 5.5
ns)16b is comparable to those of other commonly used organic a
Reactions conducted on 0.2 mmol scale. Isolated yields shown. bThe
PCs (2−20 ns in general, albeit 5.1 μs for 4CzIPN).4 In reaction was run for 4 h.

12057 https://dx.doi.org/10.1021/jacs.0c05507
J. Am. Chem. Soc. 2020, 142, 12056−12061
Journal of the American Chemical Society pubs.acs.org/JACS Communication

proceeded smoothly, delivering the desired products 6a−6c in 87% yields when 7b−7g with different electron-donating
86−93% yields (Table 1, entry 1). Substrate 4b, containing groups or electron-withdrawing groups at the para position
pyridine as the directing group (DG), was also well-tolerated, were examined. Substrates 7h−7k with diverse substituents at
furnishing the corresponding products 6d−6f in 60−73% the ortho or meta position also provided the corresponding
yields in reactions with 5a−5c (entry 2). In addition, the phenols 8h−8k in 55−73% yields. 2-Naphthylboronic acid 7l
desired C−H arylated product 6g could be obtained in 57% was also suitable for this oxidative hydroxylation, providing
yield when 4c with pyrimidine as the DG reacted with 5a naphthalen-2-ol 8l in 65% yield. This red-light-induced nPr-
(entry 3). Further, oxime 6h was isolated in 64% yield when DMQA+-catalyzed aerobic oxidative hydroxylation shows that
4d was treated with 5a (entry 4). Notably, all the products 6 3 is an efficient PC for photocatalysis involving reductive
were obtained in relatively higher yields compared to the quenching.
previous work except for 6d.21 Furthermore, control experi- To further illustrate the generality of 3 as a PC, additional
ments in the absence of 3, red light, or Pd(OAc)2 resulted in red-light-induced transformations were investigated (Scheme
much lower yields for all substrates (Table S3). The successful 2). As another classic example of photocatalysis involving
application of 3 in this red-light-mediated C−H arylation
demonstrates that it is capable of engaging in oxidative Scheme 2. Additional Examples of Red-Light-Induced nPr-
quenching during photocatalysis. DMQA+-Catalyzed Reactionsa
Next, we turned our attention to photoinduced aerobic
oxidative hydroxylation of arylboronic acids, which is a well-
documented reaction via reductive quenching under white or
blue light.23 The reaction mainly involves the oxidation of
i
Pr2NEt (DIPEA) by an excited-state PC* to form the radical
cation iPr2NEt•+ (iPr2NEt/iPr2NEt•+ = +0.72 V vs SCE)24 and
the reduction of O2 by the PC•− to generate the superoxide
radical anion O2•− (O2/O2•− = −0.57 V vs SCE.).24a,25 With
E1/2(C+*/C•) = +1.18 V and E1/2(C+/C•) = −0.74 V in DMF,
3 should be competent to serve as an efficient PC for this
reaction. Gratifyingly, phenol 8a was obtained in 87% NMR
and 83% isolated yield when 7a was treated with DIPEA and 3
in the presence of air and red light for 24 h (Table S4). In
accordance with reported results,23 in the absence of PC or red
light, little or no conversion was observed (Table S4).
Substrate scope results for oxidative hydroxylation of
arylboronic acids are presented in Table 2. A wide range of a
Reactions conducted on 0.2 mmol scale. Isolated yields shown.
n +
Table 2. Pr-DMQA -Catalyzed Aerobic Oxidative
Hydroxylation of Arylboronic Acids under Red Lighta reductive quenching and O2•−, visible-light-mediated aerobic
benzylic C(sp3)−H oxygenation using oxygen as the oxidant
was extensively studied.26 When a tertiary amine 9 was treated
with 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) and 3 in the
presence of air and red light, amide 10 was isolated in 92%
yield (Scheme 2a). Atom-transfer radical addition (ATRA) of
organic halides to olefins serves as an atom-economical
approach for simultaneously forming C−C and C−X
bonds.27 As presented in Scheme 2b, in the presence of LiBr
and 3, a red-light-induced reaction between 4-nitrobenzyl
bromide 11 and styrene 12 was realized, affording the desired
adduct 13 in 59% yield (see Tables S5 and S6 for more
optimization and Scheme S1 for mechanism). Finally, 3 could
also act as a viable alternative for Glorius’s dual gold/photo-
redox-catalyzed C(sp)−H arylation of terminal alkyne 14 with
5a,28 providing the desired product 15 in 62% yield (Scheme
2c, see more optimization in Table S7). Control experiments
in the absence of 3, red light, or other reagents such as DBU,
a
Reactions conducted on 0.5 mmol scale. Isolated yields shown.
LiBr, or Au(PPh3)Cl were also performed (Scheme S2). The
reduced yields of 10, 13, and 15 further support the essential
roles of 3 and red light for these reactions.
arylboronic acids 7 with diverse useful functional groups such Based on previous work,21,23 two plausible catalytic cycles
as halide (7c), nitrile (7d), aldehyde (7e), ester (7f), involving oxidative or reductive quenching are proposed
carboxylic acid (7g), and nitro (7k) were well-tolerated, (Scheme 3). Scheme 3-I shows a proposed mechanism for
affording the desired phenols 8a−8l in moderate to high yields. dual Pd/nPr-DMQA+-catalyzed C(sp2)−H arylation.21 First,
Additionally, the substitution pattern on the phenyl ring or photoexcitation of 3 generates nPr-DMQA+*, which reduces 5
electronic properties of 7 did not have much influence over the to form an aryl radical and nPr-DMQA•++ through oxidative
reaction outcome. For example, 8b−8g were isolated in 41− quenching. Then, addition of the aryl radical to the Pd(II)
12058 https://dx.doi.org/10.1021/jacs.0c05507
J. Am. Chem. Soc. 2020, 142, 12056−12061
Journal of the American Chemical Society


pubs.acs.org/JACS Communication

Scheme 3. Plausible Catalytic Cycles for (I) C(sp2)−H AUTHOR INFORMATION


Arylation and (II) Aerobic Oxidative Hydroxylation Corresponding Author
Thomas L. Gianetti − Department of Chemistry and
Biochemistry, University of Arizona, Tucson, Arizona 85721,
United States; orcid.org/0000-0002-3892-3893;
Email: tgianetti@arizona.edu
Authors
Liangyong Mei − Department of Chemistry and Biochemistry,
University of Arizona, Tucson, Arizona 85721, United States
José M. Veleta − Department of Chemistry and Biochemistry,
University of Arizona, Tucson, Arizona 85721, United States
Complete contact information is available at:
https://pubs.acs.org/10.1021/jacs.0c05507

Notes
The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS
We are grateful for financial support from the University of
Arizona for this work. We thank Dr. Moutet from the Gianetti
intermediate A (generated by C−H activation of substrate 4) group for help in recording cyclic voltammetry, and Prof.
affords the Pd(III) species B, followed by a one-electron Jewett and Mr. Davis from the University of Arizona for their
oxidation with nPr-DMQA•++ to regenerate nPr-DMQA+ and assistance in recording emission spectra. We thank Profs.
form the Pd(IV) intermediate C. Finally, reductive elimination Tomat, Huxter, and Njardarson from the University of
of the intermediate C furnishes the arylated product 6. A Arizona, and Prof. Bergman from UC Berkeley for helpful
possible mechanism for nPr-DMQA+-catalyzed aerobic oxida- discussions.
tive hydroxylation is presented in Scheme 3-II.23 iPr2NEt is
first oxidized to an ammonium radical cation by the excited
state of 3, along with formation of a helicene radical nPr-
■ REFERENCES
(1) For selected reviews, see: (a) Prier, C. K.; Rankic, D. A.;
DMQA• through reductive quenching. Then, the helicene MacMillan, D. W. C. Visible Light Photoredox Catalysis with
radical, which has been observed and studied by Laursen17 and Transition Metal Complexes: Applications in Organic Synthesis.
our group,29 further reacts with oxygen to regenerate 3 and Chem. Rev. 2013, 113, 5322−5363. (b) Schultz, D. M.; Yoon, T. P.
form an O2•−. Finally, follow-up oxidative attack on substrates Solar Synthesis: Prospects in Visible Light Photocatalysis. Science
2014, 343, 1239176. (c) Skubi, K. L.; Blum, T. R.; Yoon, T. P. Dual
7 and hydrolysis afford phenols 8. Catalysis Strategies in Photochemical Synthesis. Chem. Rev. 2016,
In summary, we have disclosed a helical carbenium ion, [nPr- 116, 10035−10074. (d) Hopkinson, M. N.; Tlahuext-Aca, A.; Glorius,
DMQA+][BF4−] (3), which catalyzes photoreductions and F. Merging Visible Light Photoredox and Gold Catalysis. Acc. Chem.
photooxidations in the presence of low-energy red light. The Res. 2016, 49, 2261−2272. (e) Lang, X.; Zhao, J.; Chen, X.
role of 3 as an efficient PC in oxidative and reductive Cooperative Photoredox Catalysis. Chem. Soc. Rev. 2016, 45, 3026−
quenching was evaluated by transition-metal/nPr-DMQA+- 3038. (f) Twilton, J.; Le, C. C.; Zhang, P.; Shaw, M. H.; Evans, R. W.;
catalyzed C−H arylations and intermolecular ATRA (oxidative MacMillan, D. W. C. The Merger of Transition Metal and
quenching), as well as aerobic oxidative hydroxylation of Photocatalysis. Nat. Rev. Chem. 2017, 1, 0052. (g) Parasram, M.;
arylboronic acids and benzylic C(sp3)−H oxygenation Gevorgyan, V. Visible Light-Induced Transition Metal-Catalyzed
(reductive quenching). Eight diverse substrates were well- Transformations: Beyond Conventional Photosensitizers. Chem. Soc.
Rev. 2017, 46, 6227−6240. (h) Marzo, L.; Pagire, S. K.; Reiser, O.;
tolerated for red-light-mediated dual Pd/nPr-DMQA+-cata- König, B. Visible-Light Photocatalysis: Does It Make a Difference in
lyzed C(sp2)−H arylation. Meanwhile, with 12 different Organic Synthesis? Angew. Chem., Int. Ed. 2018, 57, 10034−10072.
arylboronic acids as the substrates, red-light-induced nPr- (i) Kelly, C. B.; Patel, N. R.; Primer, D. N.; Jouffroy, M.; Tellis, J. C.;
DMQA+-catalyzed aerobic oxidative hydroxylation proceeded Molander, G. A. Preparation of Visible-Light-Activated Metal
smoothly as well. The successful applications of 3 in these red- Complexes and Their Use in Photoredox/Nickel Dual Catalysis.
light-mediated reactions have established its role as a versatile Nat. Protoc. 2017, 12, 472−492.
organic PC, which can serve as a lower-energy and therefore (2) Lowry, M. S.; Goldsmith, J. I.; Slinker, J. D.; Rohl, R.; Pascal, R.
milder option for current white-, blue-, or green-light-mediated A.; Malliaras, G. G.; Bernhard, S. Single-Layer Electroluminescent
photocatalysis. Further investigations on the applications of 3 Devices and Photoinduced Hydrogen Production from an Ionic
toward more challenging photoredox catalysis are in progress. Iridium(III) Complex. Chem. Mater. 2005, 17, 5712−5719.


(3) (a) Juris, A.; Balzani, V.; Belser, P.; von Zelewsky, A.
Characterization of the Excited State Properties of Some New
ASSOCIATED CONTENT Photosensitizers of the Ruthenium (Polypyridine) Family. Helv. Chim.
*
sı Supporting Information
Acta 1981, 64, 2175−2182. (b) Kalyanasundaram, K. Photophysics,
Photochemistry and Solar Energy Conversion with Tris(Bipyridyl)-
The Supporting Information is available free of charge at Ruthenium(II) and Its Analogues. Coord. Chem. Rev. 1982, 46, 159−
https://pubs.acs.org/doi/10.1021/jacs.0c05507. 244.
(4) (a) Romero, N. A.; Nicewicz, D. A. Organic Photoredox
Experimental procedures, characterization, and spectral Catalysis. Chem. Rev. 2016, 116, 10075−10166. (b) Nicewicz, D. A.;
data (PDF) Nguyen, T. M. Recent Applications of Organic Dyes as Photoredox

12059 https://dx.doi.org/10.1021/jacs.0c05507
J. Am. Chem. Soc. 2020, 142, 12056−12061
Journal of the American Chemical Society pubs.acs.org/JACS Communication

Catalysts in Organic Synthesis. ACS Catal. 2014, 4, 355−360. Light. RSC Adv. 2016, 6, 59269−59272. (b) Matsuzaki, K.; Hiromura,
(c) Shang, T. Y.; Lu, L. H.; Cao, Z.; Liu, Y.; He, W. M.; Yu, B. Recent T.; Tokunaga, E.; Shibata, N. Trifluoroethoxy-Coated Subphthalo-
Advances of 1,2,3,5-Tetrakis(carbazol-9-yl)-4,6-dicyanobenzene cyanine Affects Trifluoromethylation of Alkenes and Alkynes Even
(4CzIPN) in Photocatalytic Transformations. Chem. Commun. under Low-Energy Red-Light Irradiation. ChemistryOpen 2017, 6,
2019, 55, 5408−5419. (d) Phelan, J. P.; Lang, S. B.; Compton, J. 226−230. (c) Yerien, D. E.; Cooke, M. V.; García Vior, M. C.; Barata-
S.; Kelly, C. B.; Dykstra, R.; Gutierrez, O.; Molander, G. A. Redox- Vallejo, S.; Postigo, A. Radical Fluoroalkylation Reactions of
Neutral Photocatalytic Cyclopropanation via Radical/Polar Cross- (Hetero)Arenes and Sulfides under Red Light Photocatalysis. Org.
over. J. Am. Chem. Soc. 2018, 140, 8037−8047. Biomol. Chem. 2019, 17, 3741−3746.
(5) Fukuzumi, S.; Kotani, H.; Ohkubo, K.; Ogo, S.; Tkachenko, N. (14) Ravetz, B. D.; Pun, A. B.; Churchill, E. M.; Congreve, D. N.;
V.; Lemmetyinen, H. Electron-Transfer State of 9-Mesityl-10- Rovis, T.; Campos, L. M. Photoredox Catalysis Using Infrared Light
Methylacridinium Ion with a Much Longer Lifetime and Higher via Triplet Fusion Upconversion. Nature 2019, 565, 343−346.
Energy Than That of the Natural Photosynthetic Reaction Center. J. (15) Ravetz, B. D.; Tay, N. E. S.; Joe, C. L.; Sezen-Edmonds, M.;
Am. Chem. Soc. 2004, 126, 1600−1601. Schmidt, M. A.; Tan, Y.; Janey, J. M.; Eastgate, M. D.; Rovis, T. Spin-
(6) For selected examples, see: (a) Hamilton, D. S.; Nicewicz, D. A. Forbidden Excitation Enables Infrared Photoredox Catalysis.
Direct Catalytic Anti-Markovnikov Hydroetherification of Alkenols. J. ChemRxiv 2020, 12124215.
Am. Chem. Soc. 2012, 134, 18577−18580. (b) Romero, N. A.; (16) (a) Laursen, B. W.; Krebs, F. C. Synthesis of a
Margrey, K. A.; Tay, N. E.; Nicewicz, D. A. Site-Selective Arene C-H Triazatriangulenium Salt. Angew. Chem., Int. Ed. 2000, 39, 3432−
Amination via Photoredox Catalysis. Science 2015, 349, 1326−1330. 3434. (b) Delgado, I. H.; Pascal, S.; Wallabregue, A.; Duwald, R.;
(7) (a) Neumann, M.; Füldner, S.; König, B.; Zeitler, K. Metal-Free, Besnard, C.; Guénée, L.; Nançoz, C.; Vauthey, E.; Tovar, R. C.;
Cooperative Asymmetric Organophotoredox Catalysis with Visible Lunkley, J. L.; Muller, G.; Lacour, J. Functionalized Cationic
Light. Angew. Chem., Int. Ed. 2011, 50, 951−954. (b) Hari, D. P.; [4]Helicenes with Unique Tuning of Absorption, Fluorescence and
König, B. Synthetic Applications of Eosin Y in Photoredox Catalysis. Chiroptical Properties up to the Far-Red Range. Chem. Sci. 2016, 7,
Chem. Commun. 2014, 50, 6688−6699. (c) Srivastava, V.; Singh, P. P. 4685−4693.
Eosin Y Catalysed Photoredox Synthesis: A Review. RSC Adv. 2017, (17) Sørensen, T. J.; Nielsen, M. F.; Laursen, B. W. Synthesis and
7, 31377−31392. Stability of N,N′-Dialkyl-1,13-Dimethoxyquinacridinium (DMQA+):
(8) (a) Ham, W. T.; Mueller, H. A.; Sliney, D. H. Retinal Sensitivity A [4]Helicene with Multiple Redox States. ChemPlusChem 2014, 79,
to Short Wavelength Light. Nature 1976, 260, 153−155. (b) Revell, 1030−1035.
V. L.; Skene, D. J. Light-Induced Melatonin Suppression in Humans (18) (a) Mei, L.; Veleta, J. M.; Bloch, J.; Goodman, H. J.; Pierce-
with Polychromatic and Monochromatic Light. Chronobiol. Int. 2007, Navarro, D.; Villalobos, A.; Gianetti, T. L. Tunable Carbocation-
24, 1125−1137. (c) Kuse, Y.; Ogawa, K.; Tsuruma, K.; Shimazawa, Based Redox Active Ambiphilic Ligands: Synthesis, Coordination and
M.; Hara, H. Damage of Photoreceptor-Derived Cells in Culture Characterization. Dalton Trans. 2020, DOI: 10.1039/D0DT00419G.
Induced by Light Emitting Diode-Derived Blue Light. Sci. Rep. 2015, (b) Duwald, R.; Pascal, S.; Bosson, J.; Grass, S.; Besnard, C.; Bürgi,
4, 5223. T.; Lacour, J. Enantiospecific Elongation of Cationic Helicenes by
(9) Szaciłowski, K.; Macyk, W.; Drzewiecka-Matuszek, A.; Brindell, Electrophilic Functionalization at Terminal Ends. Chem. - Eur. J.
M.; Stochel, G. Bioinorganic Photochemistry: Frontiers and 2017, 23, 13596−13601.
Mechanisms. Chem. Rev. 2005, 105, 2647−2694. (19) (a) Bosson, J.; Gouin, J.; Lacour, J. Cationic Triangulenes and
(10) For selected examples, see: (a) Fülöp, A.; Peng, X.; Greenberg, Helicenes: Synthesis, Chemical Stability, Optical Properties and
M. M.; Mokhir, A. A. Nucleic Acid-Directed, Red Light-Induced Extended Applications of These Unusual Dyes. Chem. Soc. Rev. 2014,
Chemical Reaction. Chem. Commun. 2010, 46, 5659−5661. 43, 2824−2840. (b) Kel, O.; Fürstenberg, A.; Mehanna, N.; Nicolas,
(b) Zhang, H.; Trout, W. S.; Liu, S.; Andrade, G. A.; Hudson, D. C.; Laleu, B.; Hammarson, M.; Albinsson, B.; Lacour, J.; Vauthey, E.
A.; Scinto, S. L.; Dicker, K. T.; Li, Y.; Lazouski, N.; Rosenthal, J.; Chiral Selectivity in the Binding of [4]Helicene Derivatives to
Thorpe, C.; Jia, X.; Fox, J. M. Rapid Bioorthogonal Chemistry Turn- Double-Stranded DNA. Chem. - Eur. J. 2013, 19, 7173−7180.
on through Enzymatic or Long Wavelength Photocatalytic Activation (20) Nicolas, C.; Herse, C.; Lacour, J. Catalytic Aerobic Photo-
of Tetrazine Ligation. J. Am. Chem. Soc. 2016, 138, 5978−5983. oxidation of Primary Benzylic Amines Using Hindered Acridinium
(c) Carling, C. J.; Olejniczak, J.; Foucault-Collet, A.; Collet, G.; Viger, Salts. Tetrahedron Lett. 2005, 46, 4605−4608.
M. L.; Nguyen Huu, V. A.; Duggan, B. M.; Almutairi, A. Efficient Red (21) Kalyani, D.; McMurtrey, K. B.; Neufeldt, S. R.; Sanford, M. S.
Light Photo-Uncaging of Active Molecules in Water upon Assembly Room-Temperature C-H Arylation: Merger of Pd-Catalyzed C-H
into Nanoparticles. Chem. Sci. 2016, 7, 2392−2398. Functionalization and Visible-Light Photocatalysis. J. Am. Chem. Soc.
(11) (a) Cocquet, G.; Ferroud, C.; Guy, A. A Mild and Efficient 2011, 133, 18566−18569.
Procedure for Ring-Opening Reactions of Piperidine and Pyrrolidine (22) (a) Allongue, P.; Delamar, M.; Desbat, B.; Fagebaume, O.;
Derivatives by Single Electron Transfer Photooxidation. Tetrahedron Hitmi, R.; Pinson, J.; Savéant, J. M. Covalent Modification of Carbon
2000, 56, 2975−2984. (b) Cocquet, G.; Ferroud, C.; Simon, P.; Surfaces by Aryl Radicals Generated from the Electrochemical
Taberna, P. L. Single Electron Transfer Photoinduced Oxidation of Reduction of Diazonium Salts. J. Am. Chem. Soc. 1997, 119, 201−
Piperidine and Pyrrolidine Derivatives to the Corresponding Lactams. 207. (b) Andrieux, C. P.; Pinson, J. The Standard Redox Potential of
J. Chem. Soc. Perkin Trans. 2 2000, 1147−1153. the Phenyl Radical/Anion Couple. J. Am. Chem. Soc. 2003, 125,
(12) For selected examples, see: (a) Pitre, S. P.; McTiernan, C. D.; 14801−14806.
Ismaili, H.; Scaiano, J. C. Metal-Free Photocatalytic Radical (23) For selected examples, see: (a) Zou, Y. Q.; Chen, J. R.; Liu, X.
Trifluoromethylation Utilizing Methylene Blue and Visible Light P.; Lu, L. Q.; Davis, R. L.; Jørgensen, K. A.; Xiao, W. J. Highly
Irradiation. ACS Catal. 2014, 4, 2530−2535. (b) Kalaitzakis, D.; Efficient Aerobic Oxidative Hydroxylation of Arylboronic Acids:
Kouridaki, A.; Noutsias, D.; Montagnon, T.; Vassilikogiannakis, G. Photoredox Catalysis Using Visible Light. Angew. Chem., Int. Ed.
Methylene Blue as a Photosensitizer and Redox Agent: Synthesis of 5- 2012, 51, 784−788. (b) Pitre, S. P.; McTiernan, C. D.; Ismaili, H.;
Hydroxy-1H-Pyrrol-2(5H)-Ones from Furans. Angew. Chem., Int. Ed. Scaiano, J. C. Mechanistic Insights and Kinetic Analysis for the
2015, 54, 6283−6287. (c) Jiang, H.; Mao, G.; Wu, H.; An, Q.; Zuo, Oxidative Hydroxylation of Arylboronic Acids by Visible Light
M.; Guo, W.; Xu, C.; Sun, Z.; Chu, W. Synthesis of Dibenzocyclo- Photoredox Catalysis: A Metal-Free Alternative. J. Am. Chem. Soc.
ketones by Acyl Radical Cyclization from Aromatic Carboxylic Acids 2013, 135, 13286−13289. (c) Yu, X.; Cohen, S. M. Photocatalytic
Using Methylene Blue as a Photocatalyst. Green Chem. 2019, 21, Metal-Organic Frameworks for the Aerobic Oxidation of Arylboronic
5368−5373. Acids. Chem. Commun. 2015, 51, 9880−9883. (d) Xie, H. Y.; Han, L.
(13) (a) Lee, J.; Papatzimas, J. W.; Bromby, A. D.; Gorobets, E.; S.; Huang, S.; Lei, X.; Cheng, Y.; Zhao, W.; Sun, H.; Wen, X.; Xu, Q.
Derksen, D. J. Thiaporphyrin-Mediated Photocatalysis Using Red L. N-Substituted 3(10H)-Acridones as Visible-Light, Water-Soluble

12060 https://dx.doi.org/10.1021/jacs.0c05507
J. Am. Chem. Soc. 2020, 142, 12056−12061
Journal of the American Chemical Society pubs.acs.org/JACS Communication

Photocatalysts: Aerobic Oxidative Hydroxylation of Arylboronic


Acids. J. Org. Chem. 2017, 82, 5236−5241.
(24) (a) Pavlishchuk, V. V.; Addison, A. W. Conversion Constants
for Redox Potentials Measured versus Different Reference Electrodes
in Acetonitrile Solutions at 25 °C. Inorg. Chim. Acta 2000, 298, 97−
102. (b) Barbante, G. J.; Kebede, N.; Hindson, C. M.; Doeven, E. H.;
Zammit, E. M.; Hanson, G. R.; Hogan, C. F.; Francis, P. S. Control of
Excitation and Quenching in Multi-Colour Electrogenerated Chem-
iluminescence Systems through Choice of Co-Reactant. Chem. - Eur. J.
2014, 20, 14026−14031. (c) Schwarz, J.; König, B. Metal-Free,
Visible-Light-Mediated, Decarboxylative Alkylation of Biomass-
Derived Compounds. Green Chem. 2016, 18, 4743−4749.
(25) Wood, P. M. The Redox Potential of the System Oxygen-
Superoxide. FEBS Lett. 1974, 44, 22−24.
(26) For selected examples, see: (a) Finney, L. C.; Mitchell, L. J.;
Moody, C. J. Visible Light Mediated Oxidation of Benzylic sp3 C-H
Bonds Using Catalytic 1,4-Hydroquinone, or Its Biorenewable
Glucoside, Arbutin, as a Pre-Oxidant. Green Chem. 2018, 20, 2242−
2249. (b) Zhang, Y.; Riemer, D.; Schilling, W.; Kollmann, J.; Das, S.
Visible-Light-Mediated Efficient Metal-Free Catalyst for α-Oxygen-
ation of Tertiary Amines to Amides. ACS Catal. 2018, 8, 6659−6664.
(c) Ren, L.; Yang, M. M.; Tung, C. H.; Wu, L. Z.; Cong, H. Visible-
Light Photocatalysis Employing Dye-Sensitized Semiconductor:
Selective Aerobic Oxidation of Benzyl Ethers. ACS Catal. 2017, 7,
8134−8138.
(27) For selected examples, see: (a) Nguyen, J. D.; Tucker, J. W.;
Konieczynska, M. D.; Stephenson, C. R. J. Intermolecular Atom
Transfer Radical Addition to Olefins Mediated by Oxidative
Quenching of Photoredox Catalysts. J. Am. Chem. Soc. 2011, 133,
4160−4163. (b) Wallentin, C. J.; Nguyen, J. D.; Finkbeiner, P.;
Stephenson, C. R. J. Visible Light-Mediated Atom Transfer Radical
Addition via Oxidative and Reductive Quenching of Photocatalysts. J.
Am. Chem. Soc. 2012, 134, 8875−8884. (c) Pirtsch, M.; Paria, S.;
Matsuno, T.; Isobe, H.; Reiser, O. [Cu(Dap)2Cl] as an Efficient
Visible-Light-Driven Photoredox Catalyst in Carbon-Carbon Bond-
Forming Reactions. Chem. - Eur. J. 2012, 18, 7336−7340. (d) Paria,
S.; Pirtsch, M.; Kais, V.; Reiser, O. Visible-Light-Induced
Intermolecular Atom-Transfer Radical Addition of Benzyl Halides
to Olefins: Facile Synthesis of Tetrahydroquinolines. Synthesis 2013,
45, 2689−2698. (e) Rawner, T.; Lutsker, E.; Kaiser, C. A.; Reiser, O.
The Different Faces of Photoredox Catalysts: Visible-Light-Mediated
Atom Transfer Radical Addition (ATRA) Reactions of Perfluoroalkyl
Iodides with Styrenes and Phenylacetylenes. ACS Catal. 2018, 8,
3950−3956.
(28) Tlahuext-Aca, A.; Hopkinson, M. N.; Sahoo, B.; Glorius, F.
Dual Gold/Photoredox-Catalyzed C(sp)-H Arylation of Terminal
Alkynes with Diazonium Salts. Chem. Sci. 2016, 7, 89−93.
(29) Shaikh, A. C.; Hossain, M.; Moutet, J.; Veleta, J. M.; Bloch, J.;
Andrei, V.; Gianetti, T. L. Stable Helicene Radicals : Synthesis,
Structure, Physical Properties, and Photocatalysis. ChemRxiv 2020,
12408245.

12061 https://dx.doi.org/10.1021/jacs.0c05507
J. Am. Chem. Soc. 2020, 142, 12056−12061

You might also like