You are on page 1of 12

Article

pubs.acs.org/accounts

The Prowess of Photogenerated Amine Radical Cations in Cascade


Reactions: From Carbocycles to Heterocycles
Published as part of the Accounts of Chemical Research special issue “Photoredox Catalysis in Organic
Chemistry”.
Scott A. Morris, Jiang Wang, and Nan Zheng*
Department of Chemistry and Biochemistry, University of Arkansas, Fayetteville, Arkansas 72701, United States

CONSPECTUS: Cascade reactions represent a class of ideal organic reactions because they
empower efficiency, elegance, and novelty. However, development of cascade reactions
remains a daunting task for synthetic chemists. Radicals are known to be well suited for
cascade reactions. Compared with widely used carbon-based radicals, nitrogen-based radicals,
such as neutral aminyl radicals and protonated aminyl radicals (amine radical cations), are
underutilized, although they are behind some notable synthetic methods such as the
Hofmann−Löffler−Freytag reaction. The constraint on their usage is generally attributed to
the limited number of available stable precursors. Since amine radical cations offer increased
reactivity and selectivity in chemical transformations compared with neutral aminyl radicals,
their generation is of utmost importance. Recently, a surge of reports has been revealed using
visible light photoredox catalysis. It has been demonstrated that amines can act as an electron
donor in a reductive quenching cycle while the amine itself is oxidized to the amine radical
cation. Although a number of methods exist to generate amine radical cations, the
photochemical formation of these species offers many practical advantages.
In this Account, we discuss our journey to the development of annulation reactions with various π-bonds and electrophilic
addition reactions to alkenes using photogenerated amine radical cations. Various carbocycles and heterocycles are produced by
these reactions. In our annulation work, we first show that single electron photooxidation of cyclopropylanilines to the amine
radical cations triggers ring opening of the strained carbocycle, producing distonic radical cations. These odd-electron species are
shown to react with alkenes and alkynes to yield the corresponding cyclopentanes and cyclopentenes in an overall redox neutral
process. Further development of this annulation reaction allows us to achieve the [4 + 2] annulation of cyclobutylanilines with
alkynes. In our work on electrophilic addition reactions to alkenes, we reveal that photogenerated amine radical cations are
capable of undergoing the electrophilic addition reactions to alkenes to form a variety of indoles and indolines. This chemistry
represents a rare oxidative C−N bond-forming reaction using visible light. Conclusions drawn from observational results and
proposed mechanisms are outlined in this Account. Additionally, open discussion of our successes and deficiencies in our
experiences will give readers helpful insights as to how these species tend to react.
The overall utility of photogenerated amine radical cations has yet to reach its full potential. With our current results, we
anticipate more new transformations can still be derived from the ring opening processes of cyclopropylanilines and
cyclobutylanilines under visible light photocatalysis. Additionally, since utilizing photogenerated amine radical cations in C−N
bond-forming reactions has practically been absent in literature, we are confident more new reactions have yet been exploited.

■ INTRODUCTION
Compared with carbon radicals, which are considered to be
The chemistry of amine radical cations is usually determined by
the functionality around its periphery. For instance, although
mainstream in synthetic chemists’ toolbox, nitrogen-centered deprotonation is a major pathway, amine radical cations have
radicals are rather underutilized. The limited number of been shown to undergo irreversible C−C bond cleavage,
available precursors often constrains their wide use in organic electrophilic addition, and hydrogen atom abstraction, which
synthesis.1 Synthetically, amine radical cations (protonated are all determined by the nitrogen atom’s substituents.4
aminyl radicals) are typically preferred over neutral aminyl Our group has recently been intrigued by the synthetic
radicals in large part due to their increased electrophilicity, their potential of photogenerated amine radical cations, particularly
ability to add to alkenes, and their propensity to undergo their use in cascade reactions. Chemists have generated amine
synthetically useful transformations, such as the formation of radical cations using various approaches, including the use of
cyclic amines as seen in the Hofmann−Lö ffler−Freytag UV light with a photosensitizer,5 strong stoichiometric
reaction.2 In addition, amine radical cations show heightened
reactivity and selectivity over neutral aminyl radicals, adding Received: May 29, 2016
further credence to their importance in chemical synthesis.3

© XXXX American Chemical Society A DOI: 10.1021/acs.accounts.6b00263


Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

oxidants,6 enzyme oxidation,7 anodic oxidation,8,9 and visible yielded an unexpected adduct which, upon further analysis, was
light photochemistry.10 We specifically target the formation of determined to be the endoperoxide adduct derived from a
amine radical cations by direct one-electron oxidation of the [3 + 2] annulation with oxygen (Scheme 2A). We speculated
corresponding amines using visible light photoredox catalysis.
This approach overcomes some commonly observed short- Scheme 2. Initial Discovery of the [3 + 2] Annulation Using
comings in other methods, such as instability of precursors, use Cyclopropylanilines
of toxic reagents (e.g., Bu3SnH), and harsh conditions (e.g.,
strongly acidic conditions) and the need for special apparatus
(e.g., quartz).
Visible light photoredox catalysis has been a subject of great
importance recently, largely due to its ability to generate odd-
electron species such as carbon- and nitrogen-centered radicals
from stable organic substrates with full valence electrons.11
Amines are typically used as stoichiometric sacrificial electron
donors to generate the highly reducing species that initializes
radical reactions with organic substrates, and amine radical
cations are concomitantly formed. This process is referred to as
a reductive quenching cycle, one of the three main cycles for
quenching the photoexcited state (Scheme 1).4,12 However, that, as anticipated, the N−N bond cleavage would occur first
to produce cyclopropylaniline, which could then undergo the
[3 + 2] annulation with oxygen. To verify this hypothesis, we
Scheme 1. Photogenerated Amine Radical Cations Using a
prepared cyclopropylaniline and subjected it to the identical
Reductive Quenching Cycle
reactions conditions, which yielded the same endoperoxide
(Scheme 2B), giving further credence to the proposed [3 + 2]
annulation pathway.
In order for the annulation to occur, the highly strained
cyclopropane ring would need to open first, react with the π
bond of the reaction partner, and then close to complete the
transformation. The key to this annulation reaction centers on
the well-established ring opening reaction of the radical cation
of cyclopropylanilines, which would give rise to the
corresponding distonic radical cation capable of participating
in the annulation in a stepwise fashion (Scheme 3).

Scheme 3. Cyclopropylaniline Ring Opening to Generate a


Reactive Distonic Radical Cation

chemists have only recently exploited amines’ dual role as the


sacrificial electron donor and a substrate, which allows the
amine to not only reduce the photocatalyst but also serve as a
reactive substrate for useful transformations. In our attempts to
exploit amines’ dual role, we were able to form a variety of
carbocycles and heterocycles using cascade processes initialized Since our group’s goal was to exploit the reactivity of
by amine radical cations. Our first report, focusing on photogenerated amine radical cations, this system seemed
annulation reactions of cyclopropyl- and cyclobutyl-anilines appropriate for further study and examination. Furthermore,
with alkenes and alkynes, reveals a redox-neutral pathway although distonic radical cations are well-established species in
centered on a ring-opening process to form a reactive distonic the gas phase, their reactivity in solution phase is not well
ion. Our second usage afforded indoles and indolines by understood. From the limited published data, they exhibit
utilizing photogenerated amine radical cations as electrophilic unique reactivity distinct from conventional radical cations and
addition partners to alkenes. This work was carried out using related even-electron ions.14 We rationalized that this distonic
benign aerobic reaction conditions and stands in contrast to radical cation possesses a dual functionality in which the
similar indole syntheses. nucleophilic radical has the capacity to act as a nucleophile with

■ C−C BOND CLEAVAGE AS A POWERFUL TOOL


FOR THE CONSTRUCTION OF CARBOCYCLES
acceptors, such as alkenes in a Giese reaction,15 while the
electrophilic iminium ion could engage in a subsequent ring
closure as an electrophile with the resulting radical from the
Giese reaction to complete the transformation.16


Our group’s initial work in photoredox-catalyzed C−C and C−
N bond formation tandems began with [3 + 2] annulation
reactions of cyclopropylanilines with olefins. The discovery of [3 + 2] ANNULATION USING ALKENES AND
this chemistry began with a serendipitous result stemming from CYCLOPROPYLANILINES
our previous work in photoinduced N−N bond cleavage of We immediately decided to test this reaction system using
aromatic hydrazines.13 styrene in the presence of oxygen to generate the
In what appeared to be a trivial transformation, photo- corresponding carbocycle.17 Our initial results, obtained by
cleavage of the aromatic hydrazine bearing a cyclopropyl group performing the reaction with 2 mol% Ru(bpz)3(PF6)2 in
B DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

nitromethane, gave modest amounts of annulation product 2 observed in the monocyclic case. When we tested our
(21%) although 100% conversion of 1 was observed hypothesis using styrene and an angular-methyl-bearing bicycle,
(Table 1, entry 1). Presumably, the annulation of the distonic we were pleased to isolate the annulation product 3 as a 4:1
mixture of diastereomers in 77% combined yield, which
Table 1. Reaction Optimization for the [3 + 2] Annulation presumably proceeds through the distonic radical cation int 1
(Scheme 5). This result stands in direct contrast to the
monocyclic case in which tertiary amines failed to react. We
reasoned that this phenomenon was likely due to the higher
ring strain built in fused bicyclic cyclopropanes.
Carbocycles 4 and 5 were acquired using the optimized
conditions, both of which display variability on the bicycle,
entry conditions catalyst time (h) yield of 2 (%)
albeit low diastereoselectivity. However, in order to further
1 open to air Ru(bpz)3(PF6)2 12 21 enhance the diastereoselectivity, an angular tert-butyl-bearing
2 degassed Ru(bpz)3(PF6)2 3 96 bicyclic cyclopropylaniline was subjected to the annulation
3 degassed Ru(bpy)3(PF6)2 12 79 conditions with the para-methoxy analogue of styrene, which
4 degassed Ir(ppy)2(dtbbpy)PF6 12 73 yielded fused 5,5-carbocycle 6 as a single isomer in 30% yield.
We attributed the low yield to a sluggish ring opening process
radical cation with oxygen was competitive against with styrene, or slow addition to the iminium ion to close the ring, since we
resulting in low yield of 2. Further optimization revealed that recovered roughly 50% of the starting material.
the desired [3 + 2] annulation required degassed conditions, To justify the observed diastereoselectivity in the [3 + 2]
which yielded 2 as a single regioisomer in 96% yield, albeit as a annulation using bicyclic cyclopropylanilines and alkenes, we
1:1 mixture of diastereomers (entry 2). We also noted that applied the Beckwith−Houk model, which is generally used to
under the optimized conditions, other photoredox catalysts predict the diastereoselectivity of the cyclization of hexenyl
failed to match the productivity compared with Ru(bpz)32+ radicals (Figure 1).19 We determined that if the hydrogen on
(entries 3 and 4), which bolsters an impressive 1.31 V the radical carbon is axial to the R1 on the iminium, then the
reduction potential for its excited state (Ru(bpz)32+*/Ru- 1,3-diaxial interaction between R1 and R2 is obviated.
(bpz)3+ versus Ag/AgCl).
We subsequently developed a substrate scope that took into
consideration both olefin and N-cyclopropylaniline alterations,
■ REACTION EXPANSION: UTILIZING TERMINAL
ALKYNES, ENYNES, AND DIYNES
which is summarized in Scheme 4. Following this report, our group revealed further improvements
As anticipated, we were able to generate a variety of to the substrate scope, documented in subsequent accounts.20
annulation products in moderate to high yields (40−87%). In these instances, we aimed to eliminate diastereoselectivity by
Although the highest reaction efficiency was observed using using alkynes as the annulation partner to yield the
styrene as the olefin component, we were delighted to find that corresponding cyclopentenes. We were also intrigued by the
acrylonitrile and 1-phenyl-1,3-butadiene participated in the possibility to introduce an alkene functionality into the
annulation as well. We discovered that internal alkenes failed to carbocycle, thus allowing for further elaboration into other
react under the prescribed conditions, presumably due to steric useful products. Since alkynes innately have lower reactivity
hindrance, so it was not surprising that the conjugated diene’s toward the radical addition than the alkene counterparts, a
terminal double bond was more reactive in the [3 + 2] stronger light was shown to be necessary for an efficient
annulation. annulation.21 As expected, however, the optimized conditions
Our next goal was to test the viability of using bicyclic were identical to those described in our first annulation report.
cyclopropanes, readily synthesized via Kulinkovich−de Meijere Initially, we found that terminal alkynes bearing radical-
reaction of amide precursors,9,18 to give the corresponding 5,5- stabilizing groups, such as the phenyl group, and electron-
and 6,5-fused carbocycles. Furthermore, we rationalized that withdrawing groups were effective partners in these trans-
conformational bias imposed by the constraint of bicyclic formations. However, we were particularly impressed when a
cyclopropanes could improve the poor diastereoselectivity variety of diynes and enynes reacted well with cyclopropylani-

Scheme 4. Summarized Substrate Scope Using Monocyclic Cyclopropylanilines

C DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

Scheme 5. Summarized Substrate Scope Using Bicyclic Cyclopropylanilines

reactivity was more complicated, because the olefin’s steric


bulk influenced which π system (alkene or alkyne) participated
in the annulation. We found that when dealing with sterically
unhindered alkenes, the alkene moiety was the major reacting
site whereas the alkyne moiety was favored in the cases
involving sterically hindered alkenes. In certain instances when
distinguishing between the alkene and alkyne by steric bulk or
electronic factors became less clear, poor selectivity was often
Figure 1. Diastereoselectivity model for bicyclic cyclopropylamines. observed, resulting in a complicated mixture of both predictable
and unidentified annulation products. We also noted that, as
lines to give added functionality to the annulation product expected from literature precedence22 and our previous
(Scheme 6). The diynes, whether symmetrical or asymmetrical, annulation report using alkenes,17 tertiary amines failed to
exhibited excellent regioselectivity and provided the highest react in this transformation, likely due to a slower ring opening
yields when bearing electron-withdrawing groups. Enyne process.

Scheme 6. Expansion of the [3 + 2] Annulation Using Various Alkynes

D DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

Scheme 7. Rationale for Observed Regioselectivity Using Diynes

Scheme 8. Utilizing Substituted Cyclopropanes in the [3 + 2] Annulation

Scheme 9. Oxidative Cleavage of Requisite Aryl Group

Since regioselectivity became a focused point of discussion We started our study by subjecting α-methylcyclopropylaniline
with diynes and enynes, we decided to probe further into this to the optimized conditions with phenylacetylene. After 15 h,
phenomenon. Using diynes as a representative, we performed we were delighted to observe the α-substituted annulation
DFT calculations on model systems such as vinyl radicals 9 and product 11, albeit in only 33% yield (Scheme 8). We next
10 that reflected the proposed intermediates 7 and 8, turned our attention to β-substituted cyclopropylanilines. Both
respectively (Scheme 7). As expected, resonance-stabilized methyl- and phenyl-substitutions were amenable to the
vinyl radical 9, which led to the major regioisomer, was found annulation, yielding carbocycles 12 and 13 in 46% and 30%
to be 10.5 kcal/mol more stable than vinyl radical 10, which led yields, respectively. Additionally, we note that the observed
to the minor regioisomer.20b regioselectivity of the final product is derived from
We then shifted our focus to studying substituted cyclo- intermediates similar to int 2, which possess a stabilized carbon
propylanilines, focusing specifically on the cyclopropane ring. radical.
We realized that decorating the cyclopropane could result in a The N-aryl substituent quickly emerged to be a critical
larger substrate scope and further improve the utility of this constituent for these transformations as they were shown to
chemistry, since, in our previous reports, we had only examined lower the redox potential of amines.23 Cyclic voltammetry
modified cyclopropanes in the bicyclic examples with alkenes.17 experiments on both N-cyclopropylaniline and an N-
E DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

alkylcyclopropylamine revealed roughly a 0.6 V lower oxidation Scheme 11. Key Physical Data of Cyclobutylaniline and
peak potential when the nitrogen atom bears a phenyl rather Cyclopropylaniline
than an alkyl group.17 We realized that this posed a limitation
to our chemistry; therefore, we set our final focus on the
removal of the requisite aryl group. Oxidative cleavage has
shown to be effective in the removal of electron-rich N-aryl
groups, typically being mediated by strong oxidants like ceric
ammonium nitrate (CAN), periodic acid, or trichloroisocyanu-
ric acid (TCCA).24 We synthesized a variety of cyclopropyl-
amines bearing a removable aryl group in hopes they would
participate in the annulation reaction. Among the groups
screened, the para-methoxy phenyl (PMP) group was found to
be the best, being capable of undergoing the annulation
reaction while also being removable. Indeed, we were excited to
see that the annulation reaction proceeded in 47% yield to give 11B).27 On the other hand, it was reported that fragmentation
the corresponding PMP-bearing cyclopentene 14, which after of amine radical cations was significantly faster than their
much screening was cleaved using CAN/H2SO4 and then neutral aminyl radical counterpart.28 Since amine radical cations
acylated to give amide 15 in 68% yield over 2 steps (Scheme 9). were the proposed intermediate to undergo the ring opening,
In accordance with the data presented in these reports, an we felt that the ring opening’s rate might be sufficient for the
overall redox-neutral pathway was proposed (Scheme 10). annulation to occur.
With these in mind, we used 4-tert-butyl-N-cyclobutylaniline
Scheme 10. Proposed Mechanism for the [3 + 2] Annulation 16 and phenylacetylene as the model substrates to examine the
proposed [4 + 2] annulation (Table 2).29 The catalyst system

Table 2. Reaction Optimization for the [4 + 2] Annulation

time GC yield of 17
entry catalyst conditions (h) (%)
1 Ru(bpz)3(PF6)2 CH3NO2, 2 16 30
LED
2 Ru(bpz)3(PF6)2 MeOH, 2 LED 16 37
3 Ir(ppy)2(dtbbpy)PF6 MeOH, 2 LED 12 97
4 Ir(ppy)2(dtbbpy)PF6 MeOH, 1 LED 16 27

Upon irradiation with visible light, the photocatalyst gets


excited to its triplet state, which is capable of oxidizing the optimized for the [3 + 2] annulation, 2 mol% Ru(bpz)3(PF6)2
cyclopropylaniline to the corresponding amine radical cation. in nitromethane under nitrogen, was initially examined. The
Subsequent C−C bond cleavage results in the distonic radical desired [4 + 2] annulation product 17 was obtained in 30%
cation, which adds via a radical pathway to the π bond, whether yield after irradiatioin for 16 h by two 18 W LED light bulbs
alkene or alkyne. The resulting alkyl or vinyl radical then (entry 1). Encouraged by this success, we underwent systematic
cyclizes onto the iminium ion to yield another amine radical screening of the catalyst system. Methanol was found to be a
cation that can readily be reduced by the Ru+ complex, yielding better solvent than nitromethane (entry 2). Switching the
both the annulation product and starting photocatalyst.


catalyst to Ir(ppy)2(dtbbpy)PF6 made the largest impact on the
annulation, affording product 17 in 97% yield (entry 3). The
[4 + 2] ANNULATION: OPENING CYCLOBUTANE use of two lights was critical because they provided more
USING VISIBLE LIGHT photon flux and higher temperature. By comparison, irradiation
Encouraged by the successful results of the [3 + 2] annulation with one LED light furnished product 17 in 27% yield (entry
of cyclopropylanilines with various π bonds, we questioned 4). The need for stronger light in the [4 + 2] annulation was
whether cyclobutylanilines could participate in the annulation consistent with the premise that the amine radical cation of
reaction in an analogous manner to produce the [4 + 2] cyclobutylanilines undergoes the ring opening more slowly than
annulation products. Similar ring strain between cyclobutane that of cyclopropylanilines (vide supra).
and cyclopropane (27.5 vs 26.7 kcal/mol)25 as well as oxidation We next established the substrate scope by varying
peak potential between cyclopropyl- and cyclobutyl-aniline lent substituents on the N-aryl group and alkyne. Substituents
credence to this idea, as shown in Scheme 11A. However, ring with various electronic and steric characters were generally
opening of cyclobutylcarbinyl radicals was reported to be tolerated, and heterocycles showed no ill effects toward the
significantly slower than that of cyclopropylcarbinyl radicals.26 annulation (Scheme 12). With respect to alkynes, we also
The same phenomenon was observed with respect to observed a notable difference in the substrate scope between
cyclobutyl-n-propylaminyl and cyclopropyl-n-propylaminyl rad- the [3 + 2] and [4 + 2] annulation. Unlike the [3 + 2]
icals, whereas the rate constant of the latter was determined to annulation, the [4 + 2] annulation was not limited to the
be more than 100 times larger than the former (Scheme annulation with terminal alkynes, and internal alkynes also
F DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

Scheme 12. [4 + 2] Annulation Scope of Monocyclic


Cyclobutylanilines
■ AMINE RADIAL CATIONS AS A WAY TO FORM
SUBSTITUTED HETEROCYCLES
One of the hallmark features of carbon radicals is their ability to
engage in inter- and intramolecular addition to various types of
π bonds in cascade reactions. We wondered whether photo-
generated amine radical cations could mimic carbon radicals in
these types of cascade reactions leading to indoles and
indolines, two important motifs in medicinal chemistry. Similar
transformations have been successfully realized using metals,
such as Pd. Our approach was centered on a C−N bond
forming photoredox cascade in which an Umpolung reaction of
anilines was employed. The amine radical cations of anilines
acted as an electrophile, while alkenes functioned as a
nucleophile. Thus, it was mechanistically distinct from the
Pd-catalyzed process and offered advantages, including lower
reaction temperatures. More importantly, our approach utilized
the cascade reaction initialized by the photogenerated amine
radical cations, which could furnish fused indoles and indolines.
Initially, we designed styrene 18 bearing an ortho
sulfonamide as the substrate to test the cascade reaction, as
reacted with predictable regioselectivity based on the
shown in Scheme 15A. The sulfonamide group was chosen as
substituent’s ability to stabilize the vinyl radical. We attributed
the nucleophile because it was not photo-oxidizable.32 The
this difference to cyclobutylanilines’ increased stability toward
alkene moiety would be oxidized to the electrophilic alkene
light and the primary carbon radical of the distonic radical
radical cation by single electron photooxidants, such as the
cation derived from cyclobutylanilines being more nucleo-
photoexcited Ru(bpz)32+*. The targeted C−N bond formation
philic.30
would be realized via nucleophilic capture of the radical cation
We sought to further expand the substrate scope by studying
by the sulfonamide group. Instead of yielding the anticipated α-
cis fused 5,4-membered and 6,4-membered bicyclic cyclo-
phenyl indole, dimerization adduct 19 was observed. We
butylanilines, which were readily prepared in four steps. These
immediately revised our strategy and decided to attempt an
cyclobutylanilines generally gave higher yields of the annulation
Umpolung approach in which an electron-rich amine, such as
products than the monocyclic ones, and modest to good
aniline, would be oxidized to the amine radical cation by the
diastereoselectivities were achieved (Scheme 13). The opening
photoexcited Ru(bpz)32+*. The amine radical cation would
of bicyclic cyclobutylanilines was completely regioselective,
then act as an electrophile in the addition reaction to the
producing the more stable carbon radical.


alkene. The use of amine radical cations in electrophilic
addition reactions to alkenes is an established pathway and has
[4 + 2] ANNULATION IN FLOW been applied to the synthesis of many interesting molecules.33
In line with other reported photochemistries, we encountered The sulfonamide group was replaced by an aniline, as shown in
long reaction time and difficulty in scale up during our studies Scheme 15B with compound 20, which resulted in no reaction.
of the [4 + 2] annulation. Flow is often considered to be the de This was a puzzling observation, since cyclic voltammetry
facto solution to these issues. We applied flow to the [4 + 2] experiments revealed an oxidation peak potential of 0.9 V for
annulation of cyclobutylanilines with alkenes, alkynes, and 20, which suggested that 20 should be readily oxidized by the
diynes.31 The reaction time was often cut by half in flow photoexcited Ru(bpz)3(PF6)2 (1.31 V reduction potential for
(Scheme 14A). A gram scale annulation was also successfully Ru(bpz)32+*/Ru(bpz)3+ vs Ag/AgCl).34,35 Nonetheless, since
achieved in flow with a slightly lower yield (Scheme 14B). electron-rich para-alkoxyphenyl groups lower the oxidation

Scheme 13. [4 + 2] Annulation Scope of Bicyclic Cyclobutylanilines

G DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

Scheme 14. [4 + 2] Annulation in Flow

Scheme 15. Preliminary Data for Indole Synthesis Using Visible Light

peak potential of anilines, the phenyl group was exchanged for a


para-methoxyphenyl (PMP) group (compound 21). When this
■ VISIBLE LIGHT PHOTOCATALYZED SYNTHESIS OF
SUBSTITUTED INDOLES
compound was subjected to the photochemistry conditions, the Initial screening revealed Ru(bpz)3(PF6)2 as the optimal
desired indole 22 was formed. This result marked the first- photocatalyst in acetonitrile (Table 3, entries 1 and 2). It was
found that silica gel significantly increased reaction efficiency,
reported example of electrophilic addition to an olefin using
probably due to its mild acidity and ability to adsorb oxygen
photogenerated amine radical cations. With this new reaction in (entry 2). This reaction only proceeded under aerobic
hand, we set out to find the optimal reaction conditions and a conditions, suggesting a critical role of atmospheric oxygen
substrate scope tolerant of this process.34 for catalyst turnover (entry 3). Ru(bpy)3(PF6)2, possessing a
H DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

Table 3. Reaction Screen ■ FUSED INDOLINE SYNTHESIS USING TETHERED


STYRENYL ANILINES
We later expanded the scope of this chemistry by exploiting the
presumable benzylic cation to generate fused indolines.38 We
wondered whether a tethered styrenyl aniline bearing a
nucleophile could efficiently trap the cation in an intra-
time molecular SN1 fashion to yield the C2,C3-fused indoline. To
entry additive catalyst (h) yield of 23 (%) test this hypothesis, we synthesized styrenyl aniline 26 and
1 none Ru(bpz)3(PF6)2 24 31 subjected it to the photochemistry conditions (Scheme 18).
2 silica gel Ru(bpz)3(PF6)2 5 88 Fused indoline 27 was obtained as a single diastereomer.
3 silica gel (degassed) Ru(bpz)3(PF6)2 24 0 Further reaction screening identified the optimal conditions to
4 silica gel Ru(bpy)3(PF6)2 24 19 include Ru(bpy)3(PF6)2 and acetic acid with exposure to air.
5 TFA Ru(bpz)3(PF6)2 24 21 A substrate scope was also developed for this chemistry,
which included a variety of oxygen- and nitrogen-based
nucleophiles with a range of structural modifications (Scheme
19A). As expected, the olefin’s geometry had no effect on the
lower reduction potential for its photoexcited state (0.78 V),
diastereoselectivity of the reaction, since >99:1 dr was observed
was less effective in the transformation (entry 4). In order to
in all cases. Additionally, the 3,4,5-trimethoxyphenyl group was
test the acidity parameters, trifluoroacetic acid was examined,
identified as the removable N-aryl group, which was cleaved
and a lower yield of the indole product 23 was obtained due to
using CAN in sulfuric acid, thus improving the substrate scope
immediate decomposition of the product by the strong acid
(Scheme 19B). These findings revealed a photocatalyzed
(entry 5).
indoline formation by way of diamination39 or amino-
A substrate scope was then developed to demonstrate
hydroxylation40 of alkenes to form two consecutive chiral
applicability of this new reaction. Virtually every position of the
centers, including a tetrasubstituted carbon center.


indole ring could be varied when generating monosubstituted
indoles. In fact, the only structural requirement for this
MECHANISM
chemistry was the para-alkoxyphenyl group on the nitrogen
atom. A condensed substrate scope is shown in Scheme 16. A mechanism for these transformations is detailed in Scheme
An additional observation was made when using internal 20. Upon visible light irradiation, the photogenerated amine
olefins such as 24 in Scheme 17. In this case, a 1,2-alkyl shift/ radical cation, which remains in the protonated form
aromatization generated the 2,3-disubstituted indole product. presumably due to the presence of the weak acid, adds
This phenomenon has been documented by the Driver group electrophilically to the styrene, generating the quaternary
when a benzylic cation intermediate is formed.36 Therefore, we ammonium salt. Although we present this as an electrophilic
concluded that our chemistry would likely go through a similar addition, we cannot rule out a 5π pericyclic Nazarov-like
intermediate. In all our 1,2-shift cases, aromatic groups were cyclization,41 which could give the same intermediate.
found to migrate much faster than alkyl groups, resulting in a Subsequent deprotonations and redox manipulations provide
single disubstituted indole product of type 25. the key benzylic cation, which sets up the indole formation by
This chemistry marked a new way to use photogenerated way of deprotonation or 1,2-shift/deprotonation. Alternatively,
amine radical cations, effectively achieving amination of vinylic the key benzylic cation is trapped intramolecularly by a tethered
nucleophile to afford a fused indoline.


C−H bonds using mild, aerobic conditions. The only other
known example to date was reported by Knowles, in which his
group generated substituted pyrrolidines and piperidines CONCLUSION
through olefin hydroamination using photogenerated amine In summary, photogenerated amine radical cations were
radical cations.37 exploited to generate a variety of carbocycles and heterocycles.

Scheme 16. Summarized Substrate Scope

I DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

Scheme 17. 1,2-Shift Substrate Scope

Scheme 18. Examination of Indoline Formation Scheme 20. Proposed Mechanism

Scheme 19. Substrate Scope and Removal of para-


Alkoxyphenyl Component

synthetically useful benzylic cation intermediate, and a series of


redox manipulations. All of our chemistries display the power of
photogenerated amine radical cations to participate in cascade
reactions to provide structurally diverse and interesting
products.

■ AUTHOR INFORMATION
Corresponding Author
*E-mail: nzheng@uark.edu.
Funding
Our initial reports centered on the C−C bond cleavage of
photogenerated amine radical cations derived from cyclo- This work was funded by the University of Arkansas, the
propyl- and cyclobutyl-anilines to generate distonic radical Arkansas Biosciences Institute, NIGMS (Grant P30
cations. These odd-electron species displayed some unique GM103450), and NSF (Career CHE-1255539).
reactivity, permitting the annulation reaction with various Notes
alkenes and alkynes in an overall redox neutral process. We also The authors declare no competing financial interest.
developed a new application of photogenerated amine radical
Biographies
cations in electrophilic addition to tethered alkenes. Substituted
indoles and indolines were synthesized from styryl anilines. The Scott Morris received a B.S. in Biochemistry from Stephen F. Austin
amine radical cations set up the cascade reaction including the State University in 2010. He is now completing his Ph.D. in organic
electrophilic addition to the styrene, the formation of a chemistry at the University of Arkansas under the guidance of

J DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

Professor Nan Zheng, focusing on the formation of C−N bonds using (12) For reviews, see: (a) Maity, S.; Zheng, N. A photo touch on
visible light photoredox catalysis. amines: new synthetic adventures of nitrogen radical cations. Synlett
2012, 23, 1851−1856. (b) Chen, J.-R.; Hu, X.-Q.; Lu, L.-Q.; Xiao, W.-
Jiang Wang received his B.S. in Chemistry from Lanzhou University, J. Visible light photoredox-controlled reactions of N-radicals and
China. He joined the University of Arkansas in 2011 and is currently a radical ions. Chem. Soc. Rev. 2016, 45, 2044−2056.
Ph.D. candidate under the direction of Professor Nan Zheng, studying (13) Zhu, M.; Zheng, N. Photoinduced cleavage of N-N bonds of
photoredox catalysis. aromatic hydrazines and hydrazides by visible light. Synthesis 2011,
Nan Zheng did his undergraduate studies at the University of Science 2011, 2223−2236.
(14) Stirk, K. M.; Kiminkinen, L. K. M.; Kenttamaa, H. I. Ion−
and Technology of China and obtained his Ph.D. from the University
molecule reactions of distonic radical cations. Chem. Rev. 1992, 92,
of Michigan with Professor William R. Roush. He then performed
1649−1665.
postdoctoral studies at MIT with Stephen L. Buchwald before starting (15) (a) Giese, B. Formation of carbon-carbon bonds by addition of
his independent career at the University of Arkansas in 2008. His radicals to alkenes. Angew. Chem., Int. Ed. Engl. 1983, 22, 753−754.
research interest is centered on photoredox catalysis. (b) Cekovic, Z.; Saicic, R. Free radical annulation of cyclopentane ring.


Tetrahedron Lett. 1986, 27, 5893−5896. (c) Saicic, R. N.; Cekovic, Z.
ACKNOWLEDGMENTS Cyclopentane ring formation in the cycloaddition reaction of 3-alkenyl
radicals to radicophilic olefins. Tetrahedron 1990, 46, 3627−3640.
We are grateful to all our co-workers for their outstanding (16) For reviews and examples of these processes, see: (a) Reissig,
intellectual and experimental contribution. H.-U.; Zimmer, R. Donor-Acceptor-Substituted Cyclopropane De-


rivatives and Their Application in Organic Synthesis. Chem. Rev. 2003,
REFERENCES 103, 1151−1196. (b) Yu, M.; Pagenkopf, B. L. Recent advances in
donor-acceptor (DA) cyclopropanes. Tetrahedron 2005, 61, 321−347.
(1) For reviews, see: (a) Stella, L. Radicals in Organic Synthesis, 1st (c) Lebold, T. P.; Kerr, M. A. Intramolecular annulations of donor-
ed.; Wiley-VCH: Weinheim, Germany, 2001; Vol. 2, pp 407−424. acceptor cyclopropanes. Pure Appl. Chem. 2010, 82, 1797−1812.
(b) Zard, S. Z. Recent progress in the generation and use of nitrogen- (d) Mel’nikov, M. Y.; Budynina, E. M.; Ivanova, O. A.; Trushkov, I. V.
centered radicals. Chem. Soc. Rev. 2008, 37, 1603−1618. Recent advances in ring-forming reactions of donor-acceptor cyclo-
(2) Hioe, J.; Sakic, D.; Vrcek, V.; Zipse, H. The stability of nitrogen- propanes. Mendeleev Commun. 2011, 21, 293−301. (e) Cavitt, M. A.;
centered radicals. Org. Biomol. Chem. 2015, 13, 157−169. Phun, L. H.; France, S. Intramolecular donor-acceptor cyclopropane
(3) Newcomb, M.; Marquardt, D. J.; Deeb, T. M. N-Hydroxypyr- ring-opening cyclizations. Chem. Soc. Rev. 2014, 43, 804−818.
idine-2-thione carbamates. IV. A comparison of 5-exo cyclizations of (f) Schneider, T. F.; Kaschel, J.; Werz, D. B. A new golden age for
an aminyl radical and an aminium cation radical. Tetrahedron 1990, 46, donor-acceptor cyclopropanes. Angew. Chem., Int. Ed. 2014, 53, 5504−
2317−2328. 5523. (g) Grover, H. K.; Emmett, M. R.; Kerr, M. A. Carbocycles from
(4) Hu, J.; Wang, J.; Nguyen, T. H.; Zheng, N. The chemistry of donor-acceptor cyclopropanes. Org. Biomol. Chem. 2015, 13, 655−671.
amine radical cations produced by visible light photoredox catalysis. (h) Novikov, R. A.; Tomilov, Y. V. Dimerization of donor-acceptor
Beilstein J. Org. Chem. 2013, 9, 1977−2001.
cyclopropanes. Mendeleev Commun. 2015, 25, 1−10.
(5) Lee, J.; U, J. S.; Blackstock, S. C.; Cha, J. K. Facile Ring Opening
(17) Maity, S.; Zhu, M.; Shinabery, R. S.; Zheng, N. Intermolecular
of Tertiary Aminocyclopropanes by Photooxidation. J. Am. Chem. Soc.
[3 + 2] Cycloaddition of Cyclopropylamines with Olefins by Visible-
1997, 119, 10241−10242.
Light Photocatalysis. Angew. Chem., Int. Ed. 2012, 51, 222−226.
(6) (a) Beeson, T. D.; Mastracchio, A.; Hong, J.-B.; Ashton, K.;
(18) Madelaine, C.; Buzas, A. K.; Kowalska-Six, J. A.; Six, Y.; Crousse,
MacMillan, D. W. C. Enantioselective Organocatalysis Using SOMO
B. Diastereoselective Ti-mediated preparation of bicyclic amino-
Activation. Science 2007, 316, 582−585. (b) Lee, H. B.; Sung, M. J.;
cyclopropanes from N-alkenyl amides. Tetrahedron Lett. 2009, 50,
Blackstock, S. C.; Cha, J. K. Radical Cation-Mediated Annulation.
Stereoselective Construction of Bicyclo[5.3.0]decan-3-ones by Aerobic 5367−5371.
Oxidation of Cyclopropylamines. J. Am. Chem. Soc. 2001, 123, 11322− (19) Spellmeyer, D. C.; Houk, K. N. Force-field model for
11324. intramolecular radical additions. J. Org. Chem. 1987, 52, 959−974.
(7) (a) Zhong, B.; Silverman, R. B. Identification of the Active Site (20) (a) Nguyen, T. H.; Maity, S.; Zheng, N. Visible light mediated
Cysteine in Bovine Liver Monoamine Oxidase B. J. Am. Chem. Soc. intermolecular [3 + 2] annulation of cyclopropylanilines with alkynes.
1997, 119, 6690−6691. (b) Wessjohann, L. A.; Brandt, W.; Thiemann, Beilstein J. Org. Chem. 2014, 10, 975−980. (b) Nguyen, T. H.; Morris,
T. Biosynthesis and Metabolism of Cyclopropane Rings in Natural S. A.; Zheng, N. Intermolecular [3 + 2] Annulation of Cyclopropylani-
Compounds. Chem. Rev. 2003, 103, 1625−1647. lines with Alkynes, Enynes, and Diynes via Visible Light Photo-
(8) Shono, T.; Matsumura, Y.; Inoue, K.; Ohmizu, H.; Kashimura, S. catalysis. Adv. Synth. Catal. 2014, 356, 2831−2837.
Electroorganic chemistry. 62. Reaction of iminium ion with (21) Wille, U. Radical Cascades Initiated by Intermolecular Radical
nucleophile: a versatile synthesis of tetrahydroquinolines and Addition to Alkynes and Related Triple Bond Systems. Chem. Rev.
julolidines. J. Am. Chem. Soc. 1982, 104, 5753−5757. 2013, 113, 813−853.
(9) Madelaine, C.; Six, Y.; Buriez, O. Electrochemical aerobic (22) Li, X.; Grimm, M. L.; Igarashi, K.; Castagnoli, N., Jr.; Tanko, J.
oxidation of aminocyclopropanes to endoperoxides. Angew. Chem., Int. M. The first calibration of an aminiumyl radical ion clock: Why N-
Ed. 2007, 46, 8046−8049. cyclopropylanilines may be poor mechanistic probes for single electron
(10) Condie, A. G.; Gonzalez-Gomez, J. C.; Stephenson, C. R. J. transfer. Chem. Commun. 2007, 25, 2648−2650.
Visible-Light Photoredox Catalysis: Aza-Henry Reactions via C-H (23) Kavarnos, G. J.; Turro, N. J. Photosensitization by reversible
Functionalization. J. Am. Chem. Soc. 2010, 132, 1464−1465. electron transfer: theories, experimental evidence, and examples. Chem.
(11) For reviews, see: (a) Teply, F. Photoredox catalysis by Rev. 1986, 86, 401−449.
[Ru(bpy)3]2+ to trigger transformations of organic molecules. (24) Verkade, J. M. M.; van Hemert, L. J. C.; Quaedflieg, P. J. L. M.;
Organic synthesis using visible-light photocatalysis and its 20th Alsters, P. L.; van Delft, F. L.; Rutjes, F. P. J. T. Mild and efficient
century roots. Collect. Czech. Chem. Commun. 2011, 76, 859−917. deprotection of the amine protecting p-methoxyphenyl (PMP) group.
(b) Xuan, J.; Xiao, W.-J. Visible-Light Photoredox Catalysis. Angew. Tetrahedron Lett. 2006, 47, 8109−8113.
Chem., Int. Ed. 2012, 51, 6828−6838. (c) Prier, C. K.; Rankic, D. A.; (25) Wiberg, K. B. The Chemistry of Cyclobutanes; Wiley: Chichester,
MacMillan, D. W. C. Visible Light Photoredox Catalysis with U.K., 2005. Khoury, P. R.; Goddard, J. D.; Tam, W. Ring strain
Transition Metal Complexes: Applications in Organic Synthesis. energies: substituted rings, norbornanes, norbornenes and norborna-
Chem. Rev. 2013, 113, 5322−5363. dienes. Tetrahedron 2004, 60, 8103−8112.

K DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX
Accounts of Chemical Research Article

(26) Jin, J.; Newcomb, M. Rate Constants for Reactions of Alkyl


Radicals with Water and Methanol Complexes of Triethylborane. J.
Org. Chem. 2007, 72, 5098−5103.
(27) Maeda, Y.; Ingold, K. U. Kinetic applications of electron
paramagnetic resonance spectroscopy. 35. The search for a
dialkylaminyl rearrangement. Ring opening of N-cyclobutyl-N-n-
propylaminyl. J. Am. Chem. Soc. 1980, 102, 328−331.
(28) Horner, J. H.; Martinez, F. N.; Musa, O. M.; Newcomb, M.;
Shahin, H. E. Kinetics of Dialkylaminium Cation Radical Reactions:
Radical Clocks, Solvent Effects, Acidity Constants, and Rate Constants
for Reactions with Hydrogen Atom Donors. J. Am. Chem. Soc. 1995,
117, 11124−11133.
(29) Wang, J.; Zheng, N. The cleavage of a C-C Bond in
cyclobutylanilines by visible-light photoredox catalysis: Development
of a [4 + 2] annulation method. Angew. Chem., Int. Ed. 2015, 54,
11424−11427.
(30) Minisci, F.; Zammori, P.; Bernardi, R.; Cecere, M.; Galli, R.
Nucleophilic character of alkyl radicals generated in redox processes.
III. Reactivity of alkyl radicals towards conjugated olefins. Tetrahedron
1970, 26, 4153−4166.
(31) Wang, J.; Nguyen, T. H.; Zheng, N. Photoredox-catalyzed [4 +
2] annulation of cyclobutylanilines with alkenes, alkynes, and diynes in
continuous flow. Sci. China: Chem. 2016, 59, 180−183.
(32) Ross, S. D.; Finkelstein, M.; Rudd, E. J. Anodic oxidations. VIII.
Anodic oxidation of N,N-dimethylmethanesulfonamide in alcohols and
in acetic acid. J. Org. Chem. 1972, 37, 2387−2891.
(33) (a) Karady, S.; Corley, E. G.; Abramson, N. L.; Weinstock, L. M.
Cyclization of electrochemically generated nitrogen radicals. A novel
synthesis of 11-substituted dibenzo[a,d]cycloheptenimines. Tetrahe-
dron Lett. 1989, 30, 2191−2194. (b) Newcomb, M.; Deeb, T. M. N-
Hydroxypyridine-2-thione carbamates as aminyl and aminium radical
precursors. Cyclizations for synthesis of the pyrrolidine nucleus. J. Am.
Chem. Soc. 1987, 109, 3163−3165.
(34) Maity, S.; Zheng, N. A Visible-Light-Mediated Oxidative C-N
Bond Formation/Aromatization Cascade: Photocatalytic Preparation
of N-Arylindoles. Angew. Chem., Int. Ed. 2012, 51, 9562−9566.
(35) Crutchley, R. J.; Lever, A. B. P. Ruthenium(II) tris(bipyrazyl)
dication - a new photocatalyst. J. Am. Chem. Soc. 1980, 102, 7128−
7129.
(36) Stokes, B. J.; Liu, S.; Driver, T. G. Rh2(II)-Catalyzed Nitro-
Group Migration Reactions: Selective Synthesis of 3-Nitroindoles
from β-Nitro Styryl Azides. J. Am. Chem. Soc. 2011, 133, 4702−4705.
(37) Musacchio, A. J.; Nguyen, L. Q.; Beard, G. H.; Knowles, R. R.
Catalytic Olefin Hydroamination with Aminium Radical Cations: A
Photoredox Method for Direct C-N Bond Formation. J. Am. Chem.
Soc. 2014, 136, 12217−12220.
(38) Morris, S. A.; Nguyen, T. H.; Zheng, N. Diastereoselective
Oxidative C-N/C-O and C-N/C-N Bond Formation Tandems
Initiated by Visible Light: Synthesis of Fused N-Arylindolines. Adv.
Synth. Catal. 2015, 357, 2311−2316.
(39) De Jong, S.; Nosal, D. G.; Wardrop, D. J. Methods for direct
alkene diamination, new & old. Tetrahedron 2012, 68, 4067−4105.
(40) Nilov, D.; Reiser, O. The Sharpless asymmetric amino-
hydroxylation - scope and limitation. Adv. Synth. Catal. 2002, 344,
1169−1173.
(41) Radosevich, A. T.; Wiest, O. Quantum Mechanical Study of the
Ring-Closing Reaction of the Hexatriene Radical Cation. J. Org. Chem.
2001, 66, 5808−5813.

L DOI: 10.1021/acs.accounts.6b00263
Acc. Chem. Res. XXXX, XXX, XXX−XXX

You might also like