Di R2CN-OC (O) Glorius Paper

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

nature catalysis

Article https://doi.org/10.1038/s41929-022-00883-3

Energy transfer-enabled unsymmetrical


diamination using bifunctional
nitrogen-radical precursors

Received: 8 April 2022 Guangying Tan, Mowpriya Das, Roman Kleinmans, Felix Katzenburg ,
Constantin Daniliuc & Frank Glorius
Accepted: 20 October 2022

Published online: 8 December 2022


Vicinal diamines, especially unsymmetrical ones, are among the most
Check for updates common structural motifs in biologically active molecules, natural products
and pharmaceuticals. While the catalytic diamination of carbon–carbon
double bonds provides rapid access to diamines, these reactions are often
limited to installation of undifferentiated amino functionalities through
transition metals or hypervalent iodine reagent catalysis. Herein we disclose
a metal-free, photosensitized dearomative unsymmetrical diamination
of various electron-rich (hetero)arenes with bifunctional diamination
reagents, producing a series of previously inaccessible vicinal diamines
with excellent regio- and diastereoselectivity. A class of bifunctional
nitrogen-radical precursors was developed to simultaneously generate
two N-centred radicals with different reactivity via an energy transfer
process. In addition, the protocol was also found suitable for a wide range
of alkenes. Notably, the formed vicinal diamines bear two differentiated
amino functionalities, and either imine or amide units could readily and
orthogonally be converted to unprotected amines, thereby facilitating
selective downstream transformations.

Vicinal diamines are the central structural motifs of a wide variety to symmetrical diamination7–23,26–28, namely installation of two undif-
of biologically active molecules, natural products and pharma- ferentiated amino functionalities across alkenes. From the synthetic
ceuticals (Fig. 1a)1–3. Moreover, these diamine frameworks are also perspective, unsymmetrical diamination is a more attractive but
frequently employed as organocatalysts and privileged ligands for challenging task29 due to the lack of suitable amination reagents and
transition-metal catalysis4. Among various synthetic approaches to potential regio- and diastereoselectivity issues. Although some elegant
vicinal diamines, the catalytic diamination/diazidation of carbon–car- unsymmetrical alkene diaminations have been realized30–41, methods
bon double bonds is generally regarded as one of the most efficient to access a diamine precursor that could readily and orthogonally be
and straightforward methods and has been of great interest to those converted into synthetically relevant unsymmetrical diamine products
working in synthetic chemistry5,6. Despite major achievements made in remain highly desirable.
the past few decades7–41, some fundamental issues remain to be solved In recent years, particularly with the advent of visible light photo-
in this field. First, the majority of these methods require transition catalysis42–46, N-centred radicals have received considerable attention
metals or hypervalent iodine reagents as catalysts9–19,23–25,28,33–41. In addi- from the synthetic community because these powerful species offer
tion, the utilization of stoichiometric amounts of metal reagents (for great opportunities for the construction of C–N bonds with many
example, osmium or cobalt)7,8, chemical oxidants11,15,22,23,27,28,33,34,37,38 complementary aspects of both ionic and transition-metal-based
or azide reagents14–16,18–20,26,28,40,41 leads to issues on cost, the environ- methods in terms of retrosynthetic bond disconnection and reaction
ment and safety. Second, the established protocols are often limited selectivity47–52. To date various nitrogen-radical precursors have been

Westfälische Wilhelms-Universität Münster, Organisch-Chemisches Institut, Münster, Germany. e-mail: glorius@uni-muenster.de

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1120


Article https://doi.org/10.1038/s41929-022-00883-3

a H2 N H O
NHMe Me
O CO2Et N
N H O
Me
Ph2HC N Et HN
N H
N OH N H
Et O NH2 N N
N H H
NHAc N N O O
Me
Me
NHMe Ph
O

Asimadoline Tamiflu Psychotrimine Kapakahine F

The majority of vicinal diamine-containing, biologically active molecules bear unsymmetrical amino functionalities

b
Precursors for single N-centred radical Precursors for two N-centred radicals

Established Unknown
N LG N N LG N N N

This study Two differentiated


Single N-centered radical
N-centred radicals

Installation of only one single amino functionality Simultaneous installation of two differentiated amino functionalities

c O

Ph S NCPh2
Me Me
H Thioxanthone (5.0 mol%)
R N O
O N Ph + NHCOR
X Blue LEDs (λmax = 405 nm) X
O O
X = NR, O, S Vicinal diamines
Diamination reagents (Hetero)arenes

EnT Ph

N Ph
Ph

RCONH + N Ph NHCOR
X
Transient Persistent
electrophilic ambiphilic

Metal and additive free Exclusive diastereo- and regioselectivity Differentiated amino groups Orthogonal deprotection

Fig. 1 | Development of nitrogen-radical precursors. a, Selected examples of vicinal diamine-containing, biologically active molecules, drugs and natural products.
b, The development of nitrogen-radical precursors. LG, leaving group. c, This study: photochemical dearomative unsymmetrical diamination.

developed, and the radicals thus formed can be broadly divided into and alkenes (Fig. 1c). Noteworthy features of this method include the
four classes on the basis of N-hybridization and substituents, including following. (1) This mild diamination reaction was performed with
ambiphilic iminyl radicals, electrophilic amidyl radicals, nucleophilic an inexpensive and commercially available thioxanthone as organic
aminyl radicals and their protonated analogues, and electrophilic photosensitizer, without the requirement for any transition metals and
aminium radicals52. Despite continuous progress in research, to the additives, and is thus of high practicability. (2) This protocol exhibits
best of our knowledge all reported nitrogen-radical precursors are remarkable compatibility in regard to dearomative diamination of a
designed for the production of single N-centred radicals, and so they variety of (hetero)arenes, and is of exclusive diastereo- and regiose-
have been developed only for the installation of single or undifferenti- lectivity on account of the different reactivities of iminyl and amidyl
ated amino functionalities (Fig. 1b, left). To realize the more challenging radicals. (3) Most notably, the formed vicinal diamines bear two dif-
unsymmetrical diamination, we reasoned whether we could design a ferentiated amino functionalities and either imine or amide groups
bifunctional reagent that is, in principle, amenable to the simultaneous could readily and orthogonally be converted into unprotected amines,
generation of two N-centred radicals at an equal rate but with differ- thereby facilitating downstream transformation. Taken as a whole, the
ent reactivity (Fig. 1b, right). If successful, the regioselective stepwise current protocol offers a promising solution towards the key limita-
addition of two differentiated N-centred radicals across carbon–car- tions of catalytic diamination.
bon double bonds might provide direct access to the unsymmetrical
vicinal diamines. Results
Herein we report a class of oxime ester-based, bifunctional Reaction development
nitrogen-radical precursors utilized for the simultaneous generation of Considering that iminyl and amidyl radicals are both types of widely
ambiphilic iminyl and electrophilic amidyl radicals via an energy trans- explored N-centred radicals showing ambiphilic and electrophilic
fer (EnT) strategy53–55, thereby providing a rapid and versatile protocol properties52, respectively, our investigation commenced with
for unsymmetrical diamination of various electron-rich (hetero)arenes the development of iminyl- and amidyl-containing bifunctional

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1121


Article https://doi.org/10.1038/s41929-022-00883-3

a
N O

Thioxanthone (5.0 mol%)


+ N N N
N Acetone (0.1 M), RT, 12 h N
S
Boc Blue LEDs (λmax = 405 nm) Boc
H1 S1–S12 1–8 Thioxanthone

Me Me Ph Me Me Ph Me Me Ph Me Me Ph Me Me Ph
H H H H
PhthN O Ph N O Ph N O PhO N O BnO N O
O N Ph S O N Ph O N Ph O N Ph O N Ph
O O
O O O O O O O O
S1 S2 S3 S4 S5

ET = 64.7 kcal mol–1 ET = 49.5 kcal mol–1 ET = 47.5 kcal mol–1 ET = 46.9 kcal mol–1 ET = 49.4 kcal mol–1

NCPh2 NCPh2 NCPh2 NCPh2 NCPh2

NPhth NHSO2Ph NHCOPh NHCO2Ph NHCO2Bn


N N N N N
Boc Boc Boc Boc Boc

1, 25%, >95:5 d.r. 2, 29%, >95:5 d.r. 3, 53%, 77:23 r.r., >95:5 d.r. 4, 47%, 87:13 r.r.; >95:5 d.r. 5, 25%, >95:5 d.r.

NCPh2 NCPh2
Me Me Ph Me Me Ph Me Me
H H H
N O C6F5 O N O N O
Teoc O N Ph NHTeoc O N Ph NHCO2CH2C6F5 Troc O NPhth
N N
O O O O
Boc Boc

S6 6, 57%, >95:5 d.r. S7 7, 43%, >95:5 d.r. S11, N.R.

ET = 46.8 kcal mol–1 ET = 49.8 kcal mol–1 ET = 64.4 kcal mol–1

NCPh2
Me Me Ph Me Ph Ph Me Me
H H H H
N O N O N O MeO N O
Troc O N Ph Troc O N Ph Troc O N Ph NHTroc O NPhth
N
O O O O O
Boc
S8, 72%, >95:5 d.r. S9, 39%, >95:5 d.r. S10, 33%, >95:5 d.r. 8 S12, N.R.
ET = 49.4 kcal mol–1 ET = 46.8 kcal mol–1 ET = 47.3 kcal mol–1 ET = 64.5 kcal mol–1

b c
High c
Entry Variation from Yield of 8 Entry Variation from Yield of 8 –100%
standard conditions standard conditions Low c
Big scale –75%
–50%
1 Without thioxanthone N.R. 6 THF instead of acetone 38% –25%
High I 0% Low H2O
2 Without light N.R. 7 4CzlPN instead of thioxanthone N.R. +25%
+50%

3 Thioxanthone (1.0 mol %) 46% 8 Mes-Acr-Me instead of thioxanthone N.R.


Low I Low O2

4 EtOAc instead of acetone 62% 9 [Ru(bpy)3](PF6)2 instead of thioxanthone N.R.

High T High O2
5 CH2Cl2 instead of acetone 48% 10 [Ir(dF(CF3)ppy)2(dtbbpy)](PF6) 67%
instead of thioxanthone Low T

Fig. 2 | Reaction development. a, Screening of bifunctional nitrogen-radical conditions is illustrated as a black outline in a colour-coded radar diagram. The
precursors. Reaction conditions: H1 (0.4 mmol), S1–S12 (0.2 mmol) and diagram provides an intuitive readout of the general sensitivity of the reaction
thioxanthone (5.0 mol%) in acetone (0.1 M), irradiation with 18 W blue LEDs based on the shape of the black line—namely, an almost round shape around the
(λmax = 405 nm) under Ar at room temperature (RT) for 12 h. Isolated yields are 0% deviation line indicating low sensitivity; any line deflecting from that to the
shown. The d.r. values were determined by 1H NMR analysis. Triplet energies red zone indicates high sensitivity. CPCM, conductor-like polarizable continuum
were calculated at the CAM-B3LYP/def2-TZVPP level of theory after geometry model; N.R., no reaction; ET, triplet energy; T, temperature; I, light intensity;
optimization at the CAM-B3LYP/def2-SVP level of theory, applying a CPCM c, concentration; Teoc, 2-(trimethylsilyl)ethoxycarbonyl. S1–S12: numbering of
continuum solvation model for acetone. b, Impact of other reaction parameters. bifunctional nitrogen-radical precursors.
c, Assessment of condition-based sensitivity. Deviation from standard reaction

nitrogen-radical precursors. As shown in Fig. 2a, various oxime tert-butyl 1H-indole-1-carboxylate H1 as a N-centred radical accep-
ester-based nitrogen-radical precursors were synthesized and their tor. We hypothesized that these precursors could undergo N–O bond
reactivities examined under visible-light-sensitized conditions using fragmentation through photosensitization to their triplet excited

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1122


Article https://doi.org/10.1038/s41929-022-00883-3

state via EnT catalysis, resulting in iminyl and amidyl radicals along S8 was completely inhibited by the addition of 2.0 equivalent of
with the release of CO2 and aldehyde/ketone. Indeed, the desired 2,2,6,6-tetramethylpiperidinooxy (TEMPO) as a radical scavenger;
dearomative diamination reactions of H1 with a series of bifunctional TEMPO-adduct 9 and radical–radical cross-coupled product 10 were
nitrogen-radical precursors S1–S10 were observed in the presence of detected by high-resolution mass spectroscopy, and products 11–13
the inexpensive and commercially available thioxanthone as photo- were observed as the major byproducts by gas chromatography–mass
sensitizer (5.0 mol%), after irradiation for 12 h with blue light-emitting spectrometry (Supplementary Figs. 9 and 10). These results clearly
diodes (LEDs; 18 W, λmax = 405 nm) in acetone. To our pleasant surprise, hinted at the radical nature of the reaction and also demonstrated the
the desired vicinal diamine, product 8, was obtained at 72% isolated generation of iminyl and amidyl radicals (Fig. 3a). Second, the reaction
yield with excellent regio- and diastereoselectivity when employing of S8 with A38 under standard conditions formed cyclic product 14 at
2,2,2-trichloroethoxycarbonyl (Troc)-protected bifunctional reagent 22% yield, further suggesting the involvement of two N-centred radicals
S8. Reasonably, the monomethyl or α-unsubstituted Troc-protected (Fig. 3b). Subsequently, product 8 was obtained at 36% yield when the
bifunctional reagents S9 and S10 gave a reduced yield of product 8, reaction mixture of S8 and H1 was irradiated with a higher-energy light
probably because of the fact that α-substituents may have increased source (λmax = 365 nm) in the absence of any photocatalyst, indicating
the rate of the radical decarboxylation step56,57. After replacement of that single-electron transfer events are unlikely in this transformation
the benzophenone iminyl with a phthalimidyl unit (S11 and S12) the (Fig. 3c). Then, ultraviolet-visible (UV-vis) absorption spectroscopy of
transformation was shut down completely, suggesting that the benzo- all reactants and their mixtures was measured, with thioxanthone found
phenone iminyl radical with a relatively long lifespan had played a criti- to be the only absorbing species in the reaction near the irradiation
cal role in dearomative diamination58. To understand the reactivities wavelength (λmax = 405 nm) (Fig. 3d). Moreover, Stern–Volmer analysis
of these bifunctional reagents, their triplet energies were determined demonstrated that the luminescence emission of thioxanthone was
by density functional theory (DFT) calculations, ranging from 46.8 to quenched efficiently by S8 rather than by H1, implying that an interac-
64.7 kcal mol−1 (Fig. 2 and Supplementary Table 4). Meanwhile, the tion between thioxanthone and S8 might exist in the reaction medium
triplet energy of thioxanthone was calculated as 61.8 kcal mol−1 using (Fig. 3e). To understand the nature of this interaction, comparison of
the same computational method, with a solvation model for acetone as certain photocatalysts with varying properties was conducted (Fig. 3f).
applied for the bifunctional reagents. Accordingly, EnT from thioxan- As a result, yields of product 8 correlate to the triplet state energy of
thone to S2–S10 is thermodynamically feasible. In contrast, EnT to S1, photocatalysts while they are unrelated to redox properties44,60. These
S11 and S12 is endergonic, thus leading to either reduced efficiency or results implied that an EnT process is likely to have been operational
missing reactivity. To further explore the lack of reactivity of S11 and in the reaction. Furthermore, cyclic voltammetry measurement was
S12, the spin-density distributions of triplet states were determined conducted. As shown in Fig. 3g, no obvious oxidation peak of S8 and H1
(Supplementary Fig. 14). For both species, spin density is predomi- was observed before +1.4 V versus saturated calomel electrode (SCE),
nantly localized on the phthalimide functionalities and it is expected which suggests that these two compounds could not be oxidized by the
that a high energy barrier must overcome to enable homolytic N–O excited state *thioxanthone (half-wave potential, E1/2[PC]*/[PC]− = +1.18 V
bond breakage. In addition, the reaction of H1 with S8 was analysed versus SCE, where PC represents photocatalyst). Similarly, no obvi-
by gas chromatography–mass spectrometry to elucidate the composi- ous reduction peak of S8 and H1 was observed before −1.5 V versus
tion of reaction mixtures (Supplementary Fig. 3). S8 was completely SCE, which means that these two compounds could not be reduced
consumed, given its high reactivity in generating a pair of N-centred by *thioxanthone (E1/2[PC]+/[PC]* = −1.11 V versus SCE). Therefore, the ther-
iminyl and amidyl radicals under standard reaction conditions. The modynamic feasibility of a single-electron transfer (SET) reduction of
protonation of both N-centred iminyl and amidyl radicals, as well as S8 by the excited thioxanthone was also excluded by means of cyclic
the homocoupling reaction of persistent N-centred iminyl radical, voltammetry (Fig. 3g).
were the main side reactions observed in the reaction. Meanwhile, a Taking the above results together, a plausible mechanism is pro-
small amount of 1H-indole was also observed in the reaction mixtures, posed as outlined in Fig. 3h. The reaction starts with an interaction
resulting from the deterioration of H1. Subsequently, the impact of between S8 and the excited thioxanthone through a photo-induced
other reaction parameters was also investigated (Fig. 2b). As expected, EnT process to generate the excited S8*. The triplet energy of S8 is
this transformation did not proceed in the absence of either thioxan- computed to be 49.4 kcal mol−1, which is sufficient to undergo tri-
thone or blue light (Fig. 2b, entries 1 and 2). Satisfactorily, product 8 plet–triplet EnT with an excited thioxanthone (ET = 65.5 kcal mol−1).
could be obtained at 48% yield following reduction of the loading of Then, S8* undergoes N–O bond fragmentation to form a transient
thioxanthone to 1.0 mol% (Fig. 2b, entry 3). The reaction appeared amidyl radical and a persistent iminyl radical (species 15 and 16) along
to be less sensitive to the solvents, because replacement of acetone with the release of CO2 and acetone. Based on the persistent radical
with EtOAc, CH2Cl2 or THF as solvent still furnished 8 at moderate to effect61,62, the electrophilic amidyl radical 15 is captured by H1 at the
good yields (Fig. 2b, entries 4–6). Moreover, different photocatalysts C2 position to produce transient C-centred radical 17, followed by a
were screened and the [Ir(dF(CF3)ppy)2(dtbbpy)](PF6) complex also radical–radical coupling process with the ambiphilic iminyl radical
proved to be a suitable catalyst for this dearomative diamination with- 16 from the opposite, sterically accessible side to produce the desired
out diminution of the outcome. Furthermore, screening of the ratio of trans-diamine 8. We speculated that the observed regioselectivity
H1 and S8 was conducted under standard conditions (Supplementary of this reaction might have been driven mainly by the stability and
Table 2). Either reducing the dosage of H1 or switching the stoichiom- electronic property—as well as the steric hindrance—of the generated
etry of H1 and S8 led to diminished yields of 8 (Supplementary Table 2, C-centred radical after the addition of the N-centred amidyl radical
entries 1–5), probably because of the deterioration of H1 in the reaction. to H1. The addition to H1 at the C2 position generated a stabilized
Finally, a conditions-based sensitivity screening suggested that this benzyl radical (17), while the addition at the C3 position would yield a
protocol was less sensitive towards the concentration of substrates, highly nucleophilic α-aminoalkyl radical (18). Although C3 is relatively
moisture and scale-up, suffering only from slightly diminished yields more nucleophilic than C2, the polar match of ambiphilic N-centred
at high oxygen concentration or low light intensity (Supplementary iminyl radical with stabilized benzyl radical (17) might be higher than
Table 3 and Supplementary Fig. 5)59. that with highly nucleophilic α-aminoalkyl radical (18). Moreover,
the steric hindrance of the generated C-centred radical undoubtedly
Mechanistic investigations played a key role in regiocontrol. Clearly, radical 17 was less sterically
To gain some insights into this dearomative diamination, a series of hindered than radical 18, facilitating the radical–radical cross-coupling
mechanistic studies were performed. First, the reaction of H1 and of C-centred radical with N-centred iminyl radical to furnish the desired

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1123


Article https://doi.org/10.1038/s41929-022-00883-3

a
Me
Me Ph Ph
NCPh2 Troc
Me Me Ph Standard conditions N
H O NH NH N
+ N O Me Me N N
N NHTroc TrocNH2
Troc O N Ph N
TEMPO (2.0 equiv.) NHTroc Ph Ph
Boc O Ph Ph Ph Ph
Boc N
Boc 11 12 13
H1 (2.0 equiv.) S8 (1.0 equiv.) 8, Trace 9 10
Detected by HRMS Detected by HRMS Major byproducts

b c
NCPh2
Me Me Ph Me Me Ph N H1
H A38 Ph2CN NHTroc
H
N O N O Boc NHTroc
Troc O N Ph EtO2C CO2Et Troc O N Ph N
O EtO2C CO2Et O Acetone (0.1 M), RT, 24 h
Boc
Standard conditions UV lamp (λ max = 365 nm)
S8 14, 22%, >95:5 d.r. S8 8, 36%, >95:5 d.r.

d e g
1.20 250
[PC]
MeCN
H1 1.16 200
1.0 S8 in MeCN
S8
H1 in MeCN
[PC] + H1

Current (µA)
1.12 150
[PC] + S8
Intensity

S8
[PC] + H1 + S8
I0/I

1.08 H1 100
0.5

1.04
50
E1/2 = 1.47 V
1.00
0 0

300 350 400 450 0 2 4 6 8 10 12 3.0 2.5 2.0 1.5 1.0 0.5 0
Wavelength (nm) Concentration (mM) Potential (V versus SCE)

f
NCPh2
Me Me Ph 0
H [PC]
N O
+ Troc O N Ph NHTroc –10
N Acetone, hν N
O
Current (µA)

Boc Boc –20


MeCN
H1 S8 8
S8 in MeCN
–30
+ – H1 in MeCN
Entry Photocatalyst ET (kcal mol–1) E1/2 [PC] /[PC]* (V) E1/2[PC]*/[PC] (V) Yield of 8 (%)
–40
1 Thioxanthone 65.5 –1.11 +1.18 72
–50
2 [Ir(dF(CF3)ppy)2(dtbbpy)](PF6) 61.8 –0.89 +1.21 67

3 fac-Ir(ppy)3 58.1 –1.73 +0.31 ND 0 –0.5 –1.0 –1.5 –2.0 –2.5


Potential (V versus SCE)
4 [Ir(ppy)2(dtbbpy)](PF6) 49.2 –0.96 +0.66 43

5 [Ru(bpy)3](PF6)2 46.5 –0.81 +0.77 ND

h
O
* CO2 + N
Me Me Ph [PC] [PC]* Me Me Ph Me Me Ph
H H Boc
N O N O H1
Troc O N Ph Troc O N Ph TrocNH + N Ph
O O Fragmentation Radical addition
EnT catalysis
S8 S8* 15 16
ET = 49.4 kcal mol–1
Ph
N NHTroc
Ph NCPh2

N Radical–radical cross-coupling
versus Ph NHTroc
NHTroc N
Boc N
N
Ph Boc
17 Boc 18 8

Stabilized benzyl radical and less sterically hindered Highly nucleophilic a-aminoalkyl radical and more sterically hindered

Fig. 3 | Mechanistic investigations. a, TEMPO trapping experiment. dF(CF3)ppy, 2-(2,4-difluorophenyl)−5-(trifluoromethyl)pyridine; dtbbpy,


b, Radical probe experiment. c, Direct excitation. d, UV-vis absorption 4,4′-di-tert-butyl-2,2′-bipyridine; ppy, 2-phenylpyridine; bpy, 2,2’-bipyridine.
spectrum. e, Stern–Volmer quenching studies. f, Comparison of different HRMS, high-resolution mass spectrometry; I0, luminescence intensity without
photocatalysts for 8. g, Cyclic voltammetry. h, Proposed mechanism. the quencher; I, intensity in the presence of the quencher; ND, not detected.

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1124


Article https://doi.org/10.1038/s41929-022-00883-3

NCPh2
Me Me Ph Thioxanthone (5.0 mol%) O
H Acetone (0.1 M), RT, 12 h
R + N O R NHTroc + + CO2
Troc O N Ph Blue LEDs (λ max = 405 nm) Me Me
X X
X = NR, O or S O
(Hetero)arenes S8
a
NCPh2 NCPh2 NCPh2 NCPh2

NHTroc NHTroc NHTroc NHTroc


N N N N N
Boc OMe t
Bu O S Ph
O O O
8, 72%, >95:5 d.r. 19, 51%, >95:5 d.r. 20, 68%, >95:5 d.r. 21, 52%, >95:5 d.r.

NCPh2 NCPh2 NCPh2 NCPh2 NCPh2


Me MeO BnO
NHTroc NHTroc NHTroc NHTroc NHTroc
N Me N N N N

Ph Boc Boc Boc Boc


O
22, 43%, >95:5 d.r. 23, 70%, >95:5 d.r. 24, 74%, >95:5 d.r. 25, 68%, >95:5 d.r. Me 26, 59%, >95:5 d.r.
Me
Boc Me O
NCPh2 NCPh2 NCPh2 NCPh2 NCPh2
O F Br Me B
O
NHTroc NHTroc NHTroc NHTroc NHTroc
N N Cl N N N
Boc Boc Boc Boc Boc
27, 52%, >95:5 d.r. 28, 66%, >95:5 d.r. 29, 70%, >95:5 d.r. 30, 63%, >95:5 d.r. 31, 62%, >95:5 d.r.
O O
NCPh2 NCPh2 NCPh2 NCPh2 NCPh2
NC Bn
H N N
NHTroc NHTroc H NHTroc NHTroc NHTroc
N N N N N N
Boc Boc Boc Boc Boc
32, 50%, >95:5 d.r. 33, 59%, >95:5 d.r. 34, 42%, >95:5 d.r. 35, 70%, >95:5 d.r. 36, 65%, >95:5 d.r.
MeO2C
NHCO2Bn
Ph2CN Ph2CN Ph2CN Ph2CN Ph2CN
CO2Me NHBoc
Me CO2Et
NHTroc NHTroc NHTroc NHTroc NHTroc
N N N N N
Boc Boc Boc Boc Boc
37, 44%, >95:5 d.r. 38, 56%, >95:5 d.r. 39, 65%, >95:5 d.r. 40, 56%, >95:5 d.r. 41, 30%, 50:50 d.r.
NCPh2 NCPh2 H from Z-Trp-OMe
Me Br O O
Limitations
NHTroc NHTroc H H NHTroc Me
N N H
H Me NHTroc
O N N
O O N
Me Me Me Me Me H
O H27 Me Ph
42, 48%, 77:23 d.r. 43, 40%, 79:21 d.r.
No reaction 44, 16% 45, 15%
from gemfibrozil derivative from dehydrocholic acid derivative
b
NCPh2 Ph2CN Ph2CN Ph2CN Ph2CN
Me Me Me Cl Me
O
NHTroc NHTroc NHTroc NHTroc NHTroc
O O O O Me O
Cl Me N
46, 61%, >95:5 d.r. 47, 55%, >95:5 d.r. 48, 58%, >95:5 d.r. 49, 65%, >95:5 d.r. 50, 62%, >95:5 d.r.
CN Me
O

NCPh2 Ph2CN Ph2CN N NCPh2


OAc CO2Me Ph2CN N Me
NHTroc NHTroc NHTroc NHTroc NHTroc
O O O O O
51, 49%, >95:5 d.r. 52, 51%, >95:5 d.r. 53, 47%, >95:5 d.r. 54, 53%, >95:5 d.r. 55, 66%, >95:5 d.r.
c d
CO2Et NCPh2 NCPh2
NCPh2 Br NCPh2 NHTroc
Ph2CN
NHTroc NHTroc
S S
NHTroc
S NHTroc
56, 32%, >95:5 d.r. 57, 30%, >95:5 d.r. 58, 45%, >95:5 d.r. 59, 40%, >95:5 d.r. 60, 43%, >95:5 d.r.

Fig. 4 | Dearomative unsymmetrical diamination of (hetero)arenes. Reaction conditions: (hetero)arenes H1–H44 (0.4 mmol), S8 (0.2 mmol) and thioxanthone
(5.0 mol%) in acetone (0.1 M), irradiation with 18 W blue LEDs (λmax = 405 nm) under Ar at room temperature for 12 h. Isolated yields are shown. The d.r. values were
determined by 1H NMR analysis. H1–H44: numbering of (hetero)arenes.

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1125


Article https://doi.org/10.1038/s41929-022-00883-3

R1 Ph NCPh2 O
Me Me
H Thioxanthone (5.0 mol%) R1
+ N O NHTroc + + CO2
R2 Troc O N Ph Acetone (0.1 M), RT, 12 h R2 Me Me
R3 O R3
Blue LEDs (λ max = 405 nm)
Alkenes S8

NCPh2 NCPh2 62, R = tBu, 73% NCPh2


NHTroc NHTroc 63, R = NHBoc, 64% NHTroc

64, R = Cl, 76% N


R
61, 75% 65, R = Br, 72% 66, 57%

NCPh2 NCPh2 NCPh2 NCPh2 NCPh2


S NHTroc NHTroc NHTroc NHTroc NC NHTroc
Me Ph
N Me
Me I O
Me
67, 57% 68, 62% 69, 78% 70, 63%, >95:5 d.r. 71, 46%, >95:5 d.r.

Me
NCPh2 NCPh2 NCPh2 NCPh2 NCPh2 NCPh2
Ph N
NHTroc NHTroc NHTroc NHTroc NHTroc NHTroc
OH B
Ph O N O
Me 4 O
O
O O
72, 32%, >95:5 d.r. 73, 58%, 57:43 d.r. 74, 75% 75, 71% 76, 63% 77, 35%

NCPh2 NCPh2 O NCPh2 NCPh2 NCPh2


R NHTroc Boc NHTroc NHTroc EtO NHTroc Me NHTroc
N N N P
H H EtO
78, R = CHO, 54% Me O O

79, R = Boc, 48% 80, 43%, 50:50 d.r. 81, 55% 82, 38% 83, 61%

O NCPh2 O NHTroc NCPh2 NCPh2


MeO Me NHTroc NCPh2 NC NHTroc NHTroc
MeO + MeO
O
84, 50:50 d.r. Me 85, 50:50 d.r. Me OH
86, 48% 87, 42%

84 + 85, 36%, 53:47 r.r.


O
O NCPh2 NCPh2 O NCPh2 O NCPh2
NHTroc N NHTroc NHTroc PhthN NHTroc
N S O O
O 7
O
NPhth
O
88, 44% 89, 40% 90, 36%, 50:50 d.r. 91, 39%
Complex alkenes
NCPh2
Me
Me NHTroc
NCPh2 Me NCPh2 Me NCPh2
Me O
NHTroc NHTroc NHTroc
Me O
Me Me O
Me
Me O O O
Me Me
O
92, 47%, 55:45 d.r. 93, 42%, 66:34 d.r. 94, 68%, 50:50 d.r. Me Me
from carvone from nootkatone from camphanic acid derivative 95, 66%, 50:50 d.r.

Me from D-gamma-tocopherol
Me
O
O NCPh2 Me O NCPh2 Me NCPh2
H
NHTroc NHTroc NHTroc
O O O Me H O O
Me Me Me Me Me
Me N
Me S H H
O O O O
96, 48% 97, 39% 98, 40%, 50:50 d.r.
H
from gemfibrozil derivative from probenecid derivative from dehydrocholic acid derivative

Fig. 5 | Unsymmetrical diamination of alkenes. Reaction conditions: alkenes A1–A38 (0.4 mmol), S8 (0.2 mmol) and thioxanthone (5.0 mol%) in acetone (0.1 M),
irradiation with 18 W blue LEDs (λmax = 405 nm) under Ar at room temperature for 12 h. Isolated yields are shown. The d.r. values were determined by 1H NMR analysis.
A1–A38: numbering of alkenes.

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1126


Article https://doi.org/10.1038/s41929-022-00883-3

a
NH2 NCPh2 NCPh2
HCl (1 N), MeOH, r.t., 2 h TBAF, THF, r.t., 3 h
NHTroc NHTroc NH2
N N N
t
Bu tBu tBu
O O O

99, 92% 20 (10 mmol scale, 56%) 100, 78%

b c

NH2 NCPh2 NCPh2 NCPh2 NH3 Cl


HCl (1 N), MeOH TBAF, THF HCl (6 N)
NHTroc NHTroc NH2 NHTroc HO NHTroc
r.t., 2 h r.t., 3 h 100 oC, 12 h
N
O

101, 95% 61 102, 83% 76 103, 90%

d
NCPh2
Ph R Ph
Me Me DIAD, PPh3, THF Me Me R
H Standard conditions
N O N O + N
Troc O N Ph Troc O N Ph N N
EtOH or CD3OD Troc
O O Boc Boc
0 oC to r.t., 12 h
S8 S13, R = Et, 98% H1 104, R = Et, 57%, 56:44 d.r.
S14, R = CD3, 96% 105, R = CD3, 50%, 52:48 d.r.

Fig. 6 | Synthetic applications. a, Orthogonal deprotection of 20. b, Orthogonal deprotection of 61. c, Transformation of 76. d, Derivativation of S8. TBAF,
tetra-n-butylammonium fluoride; DIAD, diisopropyl azodicarboxylate.

diamine product. Notably, while 17 is a nucleophilic and relatively radical. This ultimately facilitated the smooth passage of the reaction
stabilized benzyl radical, the radical–radical cross-coupling of 17 with toward the desired dearomative diamination rather than the rearo-
the electrophilic amidyl radical 15 was not observed in the reaction mative C–H amidation. In addition, the protocol could be extended
(Supplementary Fig. 4). smoothly to various benzofurans (Fig. 4b, 46–55), as well as to benzo-
thiophenes (Fig. 4c, 56–58). Surprisingly, conjugated arenes—such as
Reaction scope phenanthrene and anthracene, despite having lower reactivity—were
After investigation of the reaction mechanism, the scope of this dearo- also found suitable for this transformation, forming the products at
mative diamination was systematically investigated using bifunctional acceptable yields (Fig. 4d, 59, 60). Most notably, high diastereoselec-
reagent S8 as N-centred radical source under optimized reaction condi- tivity was obtained within these diamination reactions because the
tions. As shown in Fig. 4, a range of representative electron-rich (hetero) stepwise addition of amidyl and iminyl radicals across carbon–carbon
arenes, such as indoles, benzofurans, benzothiophenes, phenanthrene bonds proceeded from the opposite, sterically accessible side, produc-
and anthracene, smoothly underwent this protocol to form a variety of ing a series of cyclic trans-diamines.
unsymmetrical diamines at moderate to good yields. Indoles with differ- The applicability of this method towards alkene substrates was
ent N-protection groups, such as tert-butoxycarbonyl (Boc), methoxy- also examined. As summarized in Fig. 5, a wide range of alkenes, rang-
carbonyl, pivaloyl, sulfonyl or benzoyl, reacted with S8 at satisfactory ing from styrenes, cyclic alkenes, enynes, electron-poor or -rich alk-
yields (Fig. 4a, 8, 19–22), and the structure of product 21 was further enes and even to unactivated alkenes, was included in the reaction
confirmed by X-ray crystallography. Various N-Boc-protected indoles (61–98), delivering the corresponding unsymmetrical diamines at
bearing either electron-donating or -withdrawing substituents on the moderate to good yields and with excellent regioselectivity. It is of
aromatic ring or C3 position were highly compatible with this reac- note that acyclic 1,2-disubstituted alkenes were also suitable for this
tion (Fig. 4a, 23–40). Especially sensitive functional groups, such as reaction. For example, (E)-prop-1-en-1-ylbenzene and (E)-tert-butyl
chloro, bromo, acyloxy, ester, amide, cyano, formyl—and even boronic prop-1-en-1-ylcarbamate smoothly underwent the reaction, yielding
ester—were well tolerated under the optimized conditions. To further the corresponding products 73 (57:43 diastereomeric ratio (d.r.))
test the compatibility of this method in more structurally complex and 80 (50:50 d.r.) at yields of 58 and 43%, respectively. Moreover,
contexts, several complicated indole-containing, biologically active (E)-methyl hex-3-enoate reacted with S8 to give an inseparable dias-
compounds—such as Boc-protected Z-Trp-OMe and indoles embedded tereoisomeric (50:50 d.r.) and regioisomeric (53:47 regioisomeric
in gemfibrozil or dehydrocholic acid—were tested, affording the corre- ratio (r.r.)) mixture (84 and 85) at 36% total yield. Finally, the synthetic
sponding diamines at moderate yield (41–43). Unprotected indoles and application of this protocol was further highlighted by the reaction of
N-alkyl or N-aryl indoles were, however, unsuitable for this protocol. For S8 with complex alkenes (92–98), including natural alkene-containing
instance, the reaction of 1H-indole H27 with S8 did not yield any desired products such as carvone (92) and nootkatone (93), as well as those
diamine product; neither did the reactions of S8 with 1-methyl-1H-indole derived from camphanic acid (94), d-gamma-tocopherol (95), gemfi-
H28 or 3-methyl-1-phenyl-1H-indole H29. Instead, the C–H amidation brozil (96), probenecid acid (97) and dehydrocholic acid (98).
products 44 and 45, resulting from rearomatization after the addition Subsequently, to highlight the synthetic utility of this method, a
of N-centred amidyl radical to indoles, were obtained at yields of 16 and 10 mmol-scale reaction of H3 with S8 was performed under standard
15%, respectively. These results suggested that the electron-withdrawing reaction conditions yielding product 20 at 56% yield (Fig. 6a). Moreover,
group at the indole N atom is indispensable for dearomative diamina- both imine and amide units on compound 20 were orthogonally depro-
tion. We speculated that the electron-withdrawing group at the indole tected under mild conditions to yield the corresponding unprotected
N atom might decrease the nucleophilicity of C-centred benzyl radical, amines (99 and 100)63, thereby facilitating downstream transformation.
thereby improving the polar match with ambiphilic N-centred iminyl Similarly, orthogonal deprotections of product 61 were conducted

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1127


Article https://doi.org/10.1038/s41929-022-00883-3

(Fig. 6b, 101 and 102). This protocol therefore offers general access to 2. Lucet, D., Le Gall, T. & Mioskowski, C. The chemistry of
diamine precursors that could be selectively converted to syntheti- vicinal diamines. Angew. Chem. Int. Ed. Engl. 37,
cally relevant unsymmetrical diamine products. Additionally, treat- 2580–2627 (1998).
ment of product 76 with HCl (6 N) at 100 °C for 12 h generated a valuable 3. Newhouse, T., Lewis, C. A., Eastman, K. J. & Baran, P. S. Scalable
1,2-diamino acid 103 (Fig. 6c). Finally, N-alkyl-substituted bifunctional total syntheses of N-linked tryptamine dimers by direct
nitrogen-radical precursors S13 and S14 were readily prepared through indole-aniline coupling: psychotrimine and kapakahines B and F.
the Mitsunobu reaction of S8 with EtOH or CD3OD at excellent yields J. Am. Chem. Soc. 132, 7119–7137 (2010).
(Fig. 6d)64. To our pleasant surprise, S13 and S14 smoothly underwent 4. Kizirian, J.-C. Chiral tertiary diamines in asymmetric synthesis.
this reaction, yielding the desired diamines (104 and 105) at moder- Chem. Rev. 108, 140–205 (2008).
ate yields. These findings further extend the scope of bifunctional 5. Cardona, F. & Goti, A. Metal-catalysed 1,2-diamination reactions.
nitrogen-radical precursors and highlight the practicality of this method. Nat. Chem. 1, 269–275 (2009).
6. Jong, S., de, Nosal, D. G. & Wardrop, D. J. Methods for
Conclusions direct alkene diamination, new & old. Tetrahedron 68,
In summary, we have demonstrated photosensitized dearomative 4067–4105 (2012).
unsymmetrical diamination of various electron-rich (hetero)arenes 7. Chong, A. O., Oshima, K. & Sharpless, K. B. Synthesis of
that is highly regio- and diastereoselective and runs under mild and dioxobis(tert-alkylimido)osmium(VIII) and oxotris(tert-alkylimido)
transition-metal- and additive-free conditions. The developed bifunc- osmium(VIII) complexes. Stereospecific vicinal diamination of
tional nitrogen-radical precursors were elaborately designed for this olefins. J. Am. Chem. Soc. 99, 3420–3426 (1977).
transformation to simultaneously generate two N-centred radicals with 8. Becker, P. N., White, M. A. & Bergman, R. G. A new method for
different reactivities via an EnT process. In addition, the protocol is 1,2-diamination of alkenes using cyclopentadienylnitrosylcobalt
also applicable to a wide range of alkenes. Notably, the formed vicinal dimer/NO/LiAlH4. J. Am. Chem. Soc. 102, 5676–5677 (1980).
diamines bear two differentiated amino functionalities, and either 9. Bar, G. L. J., Lloyd-Jones, G. C. & Booker-Milburn, K. I.
imine or amide units could readily and orthogonally be converted Pd(II)-catalyzed intermolecular 1,2-diamination of conjugated
to unprotected amines thereby facilitating downstream transforma- dienes. J. Am. Chem. Soc. 127, 7308–7309 (2005).
tion. We believe that this method represents an operationally simple, 10. Zhao, B., Du, H., Cui, S. & Shi, Y. Synthetic and mechanistic studies
effective and practical route to unsymmetrical vicinal diamines in both on Pd(0)-catalyzed diamination of conjugated dienes. J. Am.
academic research and industry. Chem. Soc. 132, 3523–3532 (2010).
11. Röben, C., Souto, J. A., González, Y., Lishchynskyi, A. &
Methods Muñiz, K. Enantioselective metal-free diamination of styrenes.
Representative procedure for diamination of (hetero)arenes Angew. Chem. Int. Ed. Engl. 50, 9478–9482 (2011).
or alkenes 12. Zhao, B. et al. Cu(I)-catalyzed diamination of conjugated dienes.
In an oven-dried 10 ml Schlenk tube equipped with a Complementary regioselectivity from two distinct mechanistic
poly­tet­ra­fluoroethylene-coated, rare-earth ‘extra power’ oval stir- pathways involving Cu(II) and Cu(III) species. J. Am. Chem. Soc.
ring bar, bifunctional diamination reagent (0.2 mmol, 1.0 equiv. unless 133, 20890–20900 (2011).
otherwise stated), (hetero)arene or alkene (0.4 mmol, 2.0 equiv. if 13. Zhu, Y., Cornwall, R. G., Du, H., Zhao, B. & Shi, Y. Catalytic
solid) and thioxanthone photosensitizer (2.1 mg, 5 mol%) were charged diamination of olefins via N–N bond activation. Acc. Chem. Res.
under air, then the vessel was evacuated and refilled with argon four 47, 3665–3678 (2014).
times. Dry acetone (2.0 ml, 0.1 M) and appropriate (hetero)arene or 14. Fumagalli, G., Rabet, P. T. G., Boyd, S. & Greaney, M. F.
alkene (0.4 mmol, 2.0 equiv. if liquid) were added under argon coun- Three-component azidation of styrene-type double bonds:
terflow. The vessel was tightly sealed and stirred under irradiation with light-switchable behavior of a copper photoredox catalyst.
18 W blue LEDs (λmax = 405 nm) for 12 h, unless otherwise stated. After Angew. Chem. Int. Ed. Engl. 54, 11481–11484 (2015).
irradiation, the resulting homogenous solution was transferred to a 15. Yuan, Y.-A., Lu, D.-F., Chen, Y.-R. & Xu, H. Iron-catalyzed direct
25 ml round-bottom flask with the inclusion of CH2Cl2 (2 × 3 ml). NEt3 diazidation for a broad range of olefins. Angew. Chem. Int. Ed.
(approximately 0.5 ml) and SiO2 were added to this solution and vola- Engl. 55, 534–538 (2016).
tiles were removed under reduced pressure, yielding a powder that was 16. Fu, N., Sauer, G. S., Saha, A., Loo, A. & Lin, S. Metal-catalyzed
loaded onto a column. Purification by flash-column chromatography electrochemical diazidation of alkenes. Science 357,
on SiO2, prebasified with NEt3 using pentane/EtOAc mixtures, yielded 575–579 (2017).
the corresponding diamine products. 17. Muñiz, K., Barreiro, L., Romero, R. M. & Martínez, C. Catalytic
asymmetric diamination of styrenes. J. Am. Chem. Soc. 139,
Data availability 4354–4357 (2017).
Materials and methods, experimental procedures, mechanistic stud- 18. Shen, S.-J., Zhu, C.-L., Lu, D.-F. & Xu, H. Iron-catalyzed direct olefin
ies, computational studies, sensitivity assessment and nuclear mag- diazidation via peroxyester activation promoted by nitrogenbased
netic resonance (NMR) spectra are available in the Supplementary ligands. ACS Catal. 8, 4473–4482 (2018).
Information. Crystallographic information data files and xyz coordi- 19. Fu, N., Sauer, G. S. & Lin, S. A general, electrocatalytic
nates of the optimized structures are available as Supplementary files. approach to the synthesis of vicinal diamines. Nat. Protoc. 13,
Crystallographic data for the structures reported in this article have 1725–1743 (2018).
been deposited at the Cambridge Crystallographic Data Centre under 20. Siu, J. C., Parry, J. B. & Lin, S. Aminoxyl-catalyzed electrochemical
deposition nos. CCDC 2145161 (21) and 2145162 (66). Copies of crystal- diazidation of alkenes mediated by a metastable charge-transfer
lographic data can be obtained free of charge via https://www.ccdc. complex. J. Am. Chem. Soc. 141, 2825–2831 (2019).
cam.ac.uk/structures/. All other data are available from the authors 21. Cai, C.-Y., Shu, X.-M. & Xu, H.-C. Practical and stereoselective
upon reasonable request. electrocatalytic 1,2-diamination of alkenes. Nat. Commun. 10,
4953–4959 (2019).
References 22. Tao, Z., Gilbert, B. B. & Denmark, S. E. Catalytic, enantioselective
1. Ricci, A. Amino Group Chemistry: From Synthesis to the Life syn-diamination of alkenes. J. Am. Chem. Soc. 141,
Sciences (Wiley, 2008). 19161–19170 (2019).

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1128


Article https://doi.org/10.1038/s41929-022-00883-3

23. Minakata, S., Miwa, H., Yamamoto, K., Hirayama, A. & Okumura, S. 43. Narayanam, J. M. R. & Stephenson, C. R. J. Visible light photoredox
Diastereodivergent intermolecular 1,2-diamination of unactivated catalysis: applications in organic synthesis. Chem. Soc. Rev. 40,
alkenes enabled by iodine catalysis. J. Am. Chem. Soc. 143, 102–113 (2011).
4112–4118 (2021). 44. Prier, C. K., Rankic, D. A. & MacMillan, D. W. C. Visible
24. Southgate, E. H., Pospech, J., Fu, J., Holycross, D. R. & Sarlah, D. light photoredox catalysis with transition metal complexes:
Dearomative dihydroxylation with arenophiles. Nat. Chem. 8, applications in organic synthesis. Chem. Rev. 113,
922–928 (2016). 5322–5363 (2013).
25. Wertjes, W. C., Okumura, M. & Sarlah, D. Palladium-catalyzed 45. Skubi, K. L., Blum, T. R. & Yoon, T. P. Dual catalysis strategies
dearomative syn-1,4-diamination. J. Am. Chem. Soc. 141, in photochemical synthesis. Chem. Rev. 116, 10035–10074
163–167 (2019). (2016).
26. Wu, J., Dou, Y., Guillot, R., Kouklovsky, C. & Vincent, G. 46. Stephenson, C. R. J., Yoon, T. & MacMillan, D. W. C. Visible Light
Electrochemical dearomative 2,3-difunctionalization of indoles. Photocatalysis in Organic Chemistry (Wiley, 2018).
J. Am. Chem. Soc. 141, 2832–2837 (2019). 47. Zard, S. Z. Recent progress in the generation and use
27. Wang, D., Yu, H., Sun, S. & Zhong, F. Intermolecular vicinal of nitrogen-centred radicals. Chem. Soc. Rev. 37,
diaminative assembly of tetrahydroquinoxalines via metal-free 1603–1618 (2008).
oxidative [4+2] cycloaddition strategy. Org. Lett. 22, 48. Chen, J.-R., Hu, X.-Q., Lu, L.-Q. & Xiao, W.-J. Visible light
2425–2430 (2020). photoredox-controlled reactions of N-radicals and radical ions.
28. Liu, J. et al. Diastereoselective 2, 3-diazidation of indoles via Chem. Soc. Rev. 45, 2044–2056 (2016).
copper(II)-catalyzed dearomatization. Chin. Chem. Lett. 31, 49. Kärkäs, M. D. Photochemical generation of nitrogen-centered
1332–1336 (2020). amidyl, hydrazonyl, and imidyl radicals: methodology
29. Wang, M.-M., Nguyen, T. V. T. & Waser, J. Diamine synthesis via the developments and catalytic applications. ACS Catal. 7,
nitrogen-directed azidation of σ- andπ-C-C Bonds. J. Am. Chem. 4999–5022 (2017).
Soc. 143, 11969–11975 (2021). 50. Jackman, M. M., Cai, Y. & Castle, S. L. Recent advances in iminyl
30. Li, G., Wei, H.-X., Kim, S. H. & Carducci, M. D. A novel electrophilic radical cyclizations. Synthesis 49, 1785–1795 (2017).
diamination reaction of alkenes. Angew. Chem. Int. Ed. Engl. 40, 51. Zhao, Y. & Xia, W. Recent advances in radical-based C–N bond
4277–4280 (2001). formation via photo-/electrochemistry. Chem. Soc. Rev. 47,
31. Jiang, H., Nielsen, J. B., Nielsen, M. & Jørgensen, K. A. 2591–2608 (2018).
Organocatalysed asymmetric β-amination and multicomponent 52. Davies, J., Morcillo, S. P., Douglas, J. J. & Leonori, D. Hydroxylamine
syn-selective diamination of α,β-unsaturated aldehydes. Chem. derivatives as nitrogen-radical precursors in visible-light
Eur. J. 13, 9068–9075 (2007). photochemistry. Chem. Eur. J. 24, 12154–12163 (2018).
32. Simmons, B., Walji, A. M. & MacMillan, D. W. C. Cycle- 53. Strieth-Kalthoff, F., James, M. J., Teders, M., Pitzer, L. &
specific organocascade catalysis: application to olefin Glorius, F. Energy transfer catalysis mediated by visible
hydroamination, hydro-oxidation, and amino-oxidation, and light: principles, applications, directions. Chem. Soc. Rev. 47,
to natural product synthesis. Angew. Chem. Int. Ed. Engl. 48, 7190–7202 (2018).
4349–4353 (2009). 54. Zhou, Q.-Q., Zou, Y.-Q., Lu, L.-Q. & Xiao, W.-J. Visible-light-
33. Iglesias, A. ,́ Pérez, E. G. & Muñiz, K. An intermolecular induced organic photochemical reactions through
palladium-catalyzed diamination of unactivated alkenes. Angew. energy-transfer pathways. Angew. Chem. Int. Ed. Engl. 58,
Chem. Int. Ed. Engl. 49, 8109–8111 (2010). 1586–1604 (2019).
34. Martínez, C. & Muñiz, K. Palladium-catalyzed vicinal 55. Strieth-Kalthoff, F. & Glorius, F. Triplet energy transfer
difunctionalization of internal alkenes: diastereoselective photocatalysis: unlocking the next level. Chem 6,
synthesis of diamines. Angew. Chem. Int. Ed. Engl. 51, 1888–1903 (2020).
7031–7034 (2012). 56. Jiang, H. & Studer, A. Amidyl radicals by oxidation of
35. Zhang, H. et al. Copper-catalyzed intermolecular aminocyanation α-amido-oxy acids: transition-metal-free amidofluorination
and diamination of alkenes. Angew. Chem. Int. Ed. Engl. 52, of unactivated alkenes. Angew. Chem. Int. Ed. Engl. 57,
2529–2533 (2013). 10707–10711 (2018).
36. Zhang, B. & Studer, A. Copper-catalyzed intermolecular 57. Jiang, H., Seidler, G. & Studer, A. Carboamination of
aminoazidation of alkenes. Org. Lett. 16, 1790–1793 (2014). unactivated alkenes through three-component radical
37. Olson, D. E., Su, J. Y., Roberts, D. A. & Du Bois, J. Vicinal conjugate addition. Angew. Chem. Int. Ed. Engl. 58,
diamination of alkenes under Rh-catalysis. J. Am. Chem. Soc. 136, 16528–16532 (2019).
13506–13509 (2014). 58. Patra, T., Das, M., Daniliuc, C. G. & Glorius, F. Metal-free,
38. Ciesielski, J., Dequirez, G., Retailleau, P., Gandon, V. & Dauban, photosensitized oxyimination of unactivated alkenes with
P. Rhodium-catalyzed alkene difunctionalization with nitrenes. bifunctional oxime carbonates. Nat. Catal. 4, 54–61 (2021).
Chem. Eur. J. 22, 9338–9347 (2016). 59. Pitzer, L., Schäfers, F. & Glorius, F. Rapid assessment
39. Govaerts, S. et al. Photoinduced olefin diamination of the reaction-condition-based sensitivity of chemical
with alkylamines. Angew. Chem. Int. Ed. Engl. 59, transformations. Angew. Chem. Int. Ed. Engl. 58,
15021–15028 (2020). 8572–8576 (2019).
40. Makai, S., Falk, E. & Morandi, B. Direct synthesis of 60. Nikitas, N. F., Gkizis, P. L. & Kokotos, C. G. Thioxanthone: a
unprotected 2-azidoamines from alkenes via an iron- powerful photocatalyst for organic reactions. Org. Biomol. Chem.
catalyzed difunctionalization reaction. J. Am. Chem. Soc. 142, 19, 5237–5253 (2021).
21548–21555 (2020). 61. Fische, H. The persistent radical effect: a principle for selective
41. Fan, Z., Wang, Z., Shi, R. & Wang, Y. Dirhodium(II)-catalyzed radical reactions and living radical polymerizations. Chem. Rev.
diamination reaction via a free radical pathway. Org. Chem. Front. 101, 3581–3610 (2001).
8, 5098–5104 (2021). 62. Leifert, D. & Studer, A. The persistent radical effect
42. Xuan, J. & Xiao, W.-J. Visible-light photoredox catalysis. Angew. in organic synthesis. Angew. Chem. Int. Ed. Engl. 59,
Chem. Int. Ed. Engl. 51, 6828–6838 (2012). 74–108 (2020).

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1129


Article https://doi.org/10.1038/s41929-022-00883-3

63. Huang, C.-Y. et al. Widely applicable deprotection method Additional information
of 2,2,2-trichloroethoxycarbonyl (Troc) group using Supplementary information The online version
tetrabutylammonium fluoride. J. Carbohydr. Chem. 29, contains supplementary material available at
289–298 (2010). https://doi.org/10.1038/s41929-022-00883-3.
64. Jiang, H., Yu, X., Daniliuc, C. G. & Studer, A. Three-component
aminoarylation of electron-rich alkenes by merging photoredox Correspondence and requests for materials should be addressed to
with nickel catalysis. Angew. Chem. Int. Ed. Engl. 60, Frank Glorius.
14399–14404 (2021).
Peer review information Nature Catalysis thanks Wujiong Xia,
Acknowledgements Fangrui Zhong and the other, anonymous, reviewer(s) for their
We thank P. Bellotti, X. Yu, X. Zhang (all WWU) and H. Keum (KAIST) contribution to the peer review of this work.
for helpful assistance and discussions. Generous financial support
provided by the Alexander von Humboldt Foundation (G.T.), Fonds Reprints and permissions information is available at
der Chemischen Industrie (R.K., Kekulé Scholarship no. 106151) www.nature.com/reprints.
and Deutsche Forschungsgemeinschaft (no. SFB 858) is gratefully
acknowledged. Publisher’s note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.
Author contributions
F.G. and G.T. conceived the project. G.T. and R.K. performed the Springer Nature or its licensor (e.g. a society or other partner) holds
initial screening experiments. G.T. and M.D. performed synthetic exclusive rights to this article under a publishing agreement with
experiments. F.K. conducted computations. C.D. analysed X-ray the author(s) or other rightsholder(s); author self-archiving of the
structures. G.T. and F.G. supervised research and wrote the manuscript accepted manuscript version of this article is solely governed by the
with contributions from all authors. terms of such publishing agreement and applicable law.

Competing interests © The Author(s), under exclusive licence to Springer Nature Limited
The authors declare no competing interests. 2022

Nature Catalysis | Volume 5 | December 2022 | 1120–1130 1130

You might also like