You are on page 1of 7

Article

Metal-free photoinduced C(sp3)–H borylation


of alkanes

https://doi.org/10.1038/s41586-020-2831-6 Chao Shu1, Adam Noble1 ✉ & Varinder K. Aggarwal1 ✉

Received: 23 May 2020

Accepted: 27 August 2020 Boronic acids and their derivatives are some of the most useful reagents in the
Published online: 28 October 2020 chemical sciences1, with applications spanning pharmaceuticals, agrochemicals and
functional materials. Catalytic C–H borylation is a powerful method for introducing
Check for updates
these and other boron groups into organic molecules because it can be used to
directly functionalize C–H bonds of feedstock chemicals without the need for
substrate pre-activation1–3. These reactions have traditionally relied on
precious-metal catalysts for C–H bond cleavage and, as a result, display high
selectivity for borylation of aromatic C(sp2)–H bonds over aliphatic C(sp3)–H bonds4.
Here we report a mechanistically distinct, metal-free borylation using hydrogen atom
transfer catalysis5, in which homolytic cleavage of C(sp3)–H bonds produces alkyl
radicals that are borylated by direct reaction with a diboron reagent. The
reaction proceeds by violet-light photoinduced electron transfer between an
N-alkoxyphthalimide-based oxidant and a chloride hydrogen atom transfer catalyst.
Unusually, stronger methyl C–H bonds are borylated preferentially over weaker
secondary, tertiary and even benzylic C–H bonds. Mechanistic studies indicate that
the high methyl selectivity is a result of the formation of a chlorine radical–boron ‘ate’
complex that selectively cleaves sterically unhindered C–H bonds. By using a
photoinduced hydrogen atom transfer strategy, this metal-free C(sp3)–H borylation
enables unreactive alkanes to be transformed into valuable organoboron reagents
under mild conditions and with selectivities that contrast with those of established
metal-catalysed protocols.

The importance of organoboron compounds has triggered intensive for borylation of alkyl groups (BDE = 98 kcal mol−1 for cyclohexyl–H)24
research into methods for introducing boron into organic molecules, over aromatic rings (BDE = 113 kcal mol−1 for phenyl–H)25.
the most direct of which is C–H borylation (Fig. 1a)2,3. This approach Herein we report a photoinduced borylation of non-activated C(sp3)–
benefits from introducing a synthetically valuable boron moiety into H bonds using chloride as a HAT catalyst. The use of a radical-mediated
inherently unreactive feedstock chemicals by substitution of ubiqui- strategy enables a wide range of functionalized alkanes to be boryl-
tous C–H bonds. The C(sp2)–H borylation of aromatic compounds is ated under mild conditions (ambient temperature, violet-light irra-
well established (for example, using iridium catalysts)6,7, but C(sp3)– diation) and circumvents the high C(sp2)–H selectivity inherent in
H borylation of alkanes is less developed4. The majority of reported transition-metal-catalysed borylations. Furthermore, direct photo-
C(sp3)–H borylations use activated substrates8–13 or those possessing excitation of an N-alkoxyphthalimide oxidant allows these reactions
heteroatom directing groups14–18, whereas borylations of non-activated to occur in the absence of a metal catalyst26, therefore providing
C(sp3)–H bonds require the use of precious-metal catalysts and harsh room-temperature and metal-free C(sp3)–H bond borylations.
reaction conditions19–23. Furthermore, the higher reactivity of aromatic Our initial design plan for C–H borylation was inspired by our
C–H bonds prohibits selective borylation of non-activated C(sp3)–H decarboxylative borylation of N-hydroxyphthalimide esters 1 with
bonds in the presence of sterically accessible C(sp2)–H bonds23. bis(catecholato)diboron (2, B2(cat)2; Fig. 1c)27. In that process, photoin-
We envisioned an alternative C(sp3)–H borylation using a hydro- duced electron transfer (PET) under blue-light irradiation provided
gen atom transfer (HAT) strategy5. Here, the C–H bond is cleaved by phthalimide radical anion intermediates 3. In the case of aliphatic esters
intermolecular reaction with a heteroatom-centred radical (X•), rather (R1 = alkyl), decarboxylative fragmentation furnished phthalimide anion
than with a metal catalyst, and subsequent homolytic substitution of a 4 and an alkyl radical, 5, which was borylated by reaction with B2(cat)2
diboron reagent by the resulting alkyl radical intermediate forms the to form boronic ester 6. We reasoned that if N-alkoxyphthalimides
C–B bond (Fig. 1b). An important feature of this proposed mechanism 7 were used in place of esters 1, fragmentation of the radical anion
is that regioselectivity should be determined by the bond dissociation intermediate 8 by N–O bond cleavage would produce a highly reactive
energy (BDE) of the C–H bond, and thus there will be high selectivity oxygen-centred radical (9) that is capable of cleaving strong C(sp3)–H

School of Chemistry, University of Bristol, Bristol, UK. ✉e-mail: a.noble@bristol.ac.uk; v.aggarwal@bristol.ac.uk


1

714 | Nature | Vol 586 | 29 October 2020


a b OR
RO OR Metal catalyst OR H RO OR HAT catalyst
B
H + B B B + B B OR
RO OR OR RO OR

HAT Borylation
H H
[M]
X H X B2(OR) 4 B(OR) 2
Alkyl radical
Via alkyl metal C(sp2)–H C(sp3)–H intermediate
intermediate High reactivity Low reactivity
Mild conditions, transition metal-free, C(sp3)–H selective

c O O
X O O Visible light X
N + B B N
O R1 O O PET O R1
O O

1 (X = O), 7 (X = H 2) B2(cat)2 (2) 3 (X = O), 8 (X = H 2 )

For X = O: decarboxylative borylation (ref.27)

O O
O –CO 2 B2(cat)2
N N + R1 R1 B(cat)
O R1 β-scission Borylation
3 O 4 O 5 6

For X = H2: HAT borylation (this work)

R2 H HO R1
O
10 11
B2(cat)2
N 4 + O R1 R2 R2 B(cat)
1
O R β-scission HAT Borylation
8 O 9 12 13

d Catalyst
14 (1 equiv.) Pinacol, Et3N
+ B 2(cat)2
H MeCN (0.025 M), 25 °C, 24 h B(cat) B(pin)
390 nm LEDs
15 2 16 17
(10 equiv.) (1 equiv.)

Entry Catalyst (mol%) Yield of 17 by GC


O 1 None 0%
2 CeCl6(NEt 4) 2 (10) 36%
N
O CF 3 3 NEt4Cl (20) 33%
O 4 ClB(cat) (20) 61%
14

Fig. 1 | Catalytic C–H borylation reactions. a, Transition-metal-catalysed C–H borylation of N-hydroxyphthalimide esters 1 (ref. 27) and proposed C(sp3)–H
borylations. These reactions proceed via cleavage of the C–H bond by the borylation using N-alkoxyphthalimides 7. d, Reaction development. Yields are
metal catalyst to form alkyl metal intermediates. b, Proposed radical-mediated based on molar equivalents of 2. LED, light-emitting diode; cat, catecholato;
C(sp3)–H borylation using HAT catalysis. c, Photoinduced decarboxylative pin, pinacolato; GC, gas chromatography.

bonds via HAT28–30. Intermolecular HAT between 9 and an alkane 10 alkoxyphthalimide 14 in acetonitrile (Fig. 1d), but unfortunately no
(R2 = alkyl) would generate alcohol 11 and alkyl radical 12, which would C–H borylation product was observed (Fig. 1d, entry 1). On the basis of
be rapidly borylated by reaction with B2(cat)2 to give C–H borylation a recent report on alkoxy-radical-mediated C(sp3)–H functionalization
product 1331. using cerium catalysis32, we investigated the effect of cerium salts on
We postulated that the nature of the oxygen-centred radical 9 would the borylation reaction (Supplementary Table 1)33. Pleasingly, in the
be crucial for successfully implementing our reaction design. For pro- presence of cerium(iv) chloride tetraethylammonium chloride com-
ductive C–H borylation, 9 must (1) be stable to β-fragmentation to alkyl plex (CeCl6(NEt4)2), irradiation with violet light gave boronic ester 17
radicals, (2) be electrophilic, therefore favouring reaction with hydridic in 36% yield (Fig. 1d, entry 2). Unexpectedly, a control experiment with
aliphatic hydrogen atoms over direct reaction at the electrophilic boron just tetraethylammonium chloride (NEt4Cl) as catalyst resulted in the
atoms of 2, and (3) form strong O–H bonds (BDE > 100 kcal mol−1) upon formation of 17 in a similar yield, suggesting that chloride is catalys-
HAT, thus allowing borylation of strong C(sp3)–H bonds. Considering ing the reaction (Fig. 1d, entry 3). Further improvements were made
these requirements, we selected N-(2,2,2-trifluoroethoxy)phthalimide through analysis of a range of other chloride sources (Supplementary
(14) (Fig. 1d) as this provides the electrophilic 2,2,2-trifluoroethoxy Table 2), with 20 mol% B-chlorocatecholborane (ClB(cat)) proving
radical, which has previously been reported to cleave strong C(sp3)–H optimal, enabling formation of 17 in 61% yield (Fig. 1d, entry 4). Owing
bonds32. to the instability of the initially formed catechol boronic ester 16 to
Our initial investigations focused on the borylation of norbor- hydrolysis, in all cases in situ transesterification to the stable pinacol
nane (15), which was irradiated in the presence of B 2(cat)2 and boronic ester 17 was performed before isolation.

Nature | Vol 586 | 29 October 2020 | 715


Article
O
H ClB(cat) (20 mol%), 14 (1 equiv.) B(pin) O
+ B2 (cat) 2 Cl B N
MeCN (0.025 M), 390 nm LEDs, O O CF3
2 then pinacol, Et3N O
ClB(cat) 14
(10 equiv.) (1 equiv.)

H
Br β
Cl β γ
18 (n = 0): 40% β α B(pin) NC α B(pin)
B(pin) B(pin)
B(pin) α B(pin)
19 (n = 1): 55% α
H β

17: 58% ( )n 20 (n = 2): 50% 22: 48%d 23: 37%d 24: 59%d 25: 50%
(66%a, 47%c ) α:β = 85:15 α:β = 84:16 α:β = 57:43 α:β:γ = <1:25:75
21 (n = 3): 52%
>97:3 r.r. α: 96:4 d.r. α: 81:19 (exo:endo) α: 83:17 (exo:endo) γ: 50:50 (trans:cis)
86:14 d.r. β: >97:3 d.r. β: 72:28 (exo:endo) β: 7:93 (exo:endo) β: 96:4 (trans:cis)

γ α ε γ α Cl
β B(pin) Cl B(pin)
B(pin) B(pin) B(pin)
B(pin) ( )n
β

26: 35%e 27 (n = 0): 39%, >97:3 r.r.f 30: 46% (44%b) 31: 40% 32: 50% (46%b) 33: 41%
α:β:γ = 54:36:10 28 (n = 1): 43%, >97:3 r.r. α:β:γ:ε = 54:<1:<1:46 f >97:3 r.r.f >97:3 r.r. >97:3 r.r.
29 (n = 2): 33% (28%b), >97:3 r.r.
O
H α γ
Br B(pin) Br B(pin) N B(pin) CF 3 N
B(pin) NC B(pin)
R β
O
O
34 (R = Me): 35%, >97:3 r.r.f 36: 30% (29%b) 38: 21% 39: 40%
35 (R = H): 42%, >97:3 r.r. >97:3 r.r. 37: 26%, >97:3 r.r.f >97:3 r.r.f α:β:γ = <1:28:72

R tBu

B(pin) β O B(pin)
γ B(pin)
B(pin) B(pin) α
B(pin) Ar O
δ

40 (R = H): 19%, >97:3 r.r. 43: 51% 44: 25% 45: 28% 46: 51%, α:β:γ:δ = <1:14:57:29 47 (Ar = 2-furyl): 30%
41 (R = Me): 14%, >97:3 r.r. >97:3 r.r.f >97:3 r.r. β: 89:11 d.r. (trans:cis) >97:3 r.r.
42 (R = CN): 24%, >97:3 r.r. γ: 95:5 d.r. (trans:cis) 48 (Ar = 3-thiophenyl): 26%
δ: >97:3 d.r. (trans:cis) >97:3 r.r.
O

O O O O
O B(pin) O B(pin)
B(pin) B(pin)
O MeO B(pin) MeO
MeO O
(Phth)N N(Phth)
O
49: 15% 50: 25% 52: 20% 53: 25% (19%b, 18%g)
>97:3 r.r. >97:3 r.r.f 51: 19% >97:3 r.r., 55:45 d.r.f >97:3 r.r.f
>97:3 r.r., 59:41 d.r.f

a
Fig. 2 | Photoinduced C–H borylations of alkanes. Reactions were performed 20 equiv., b5 equiv., c3 equiv. dr.r. and d.r. were determined by 1H NMR analysis
with 0.3 mmol of B2(cat)2. Yellow and purple spheres represent protons or alkyl after oxidation to the corresponding alcohol. eThe yield was determined by GC
groups. Yields are of isolated products. Regioisomeric ratios (r.r.) and analysis. fr.r. and d.r. were determined by 1H NMR analysis. gUsing 1.0 equiv.
diastereomeric ratios (d.r.) were determined by GC analysis. Numbers in alkane and 1.2 equiv. B2(cat)2. N(Phth), N-phthalimide; t Bu, tert-butyl.
parentheses show yields obtained using different alkane stoichiometry:

With these metal-free C(sp3)–H borylation conditions, we proceeded strength of the C–H bond, thus favouring reaction at methylene (BDE =
to explore the scope of the reaction (Fig. 2). Unfunctionalized cyclic 99 kcal mol−1 for iPr–H) over methyl groups (BDE = 101 kcal mol−1 for
alkanes, which are commonly used as solvents in metal-catalysed Et–H)25. Pentane was borylated with a methyl/methylene selectivity
C–H borylations owing to their low reactivity23, were successfully of approximately 1:1 (26), whereas for substrates containing sterically
borylated to provide alkylboronic esters 17–22. In addition, boryla- hindered methylene groups, complete methyl selectivity was observed
tion of cycloalkanes functionalized with halides or nitrile groups was (28–31). Notably, 2,2,4,4-tetramethylpentane (31) was borylated effi-
also possible (23–25). High regio- and diastereoselectivities were ciently; to our knowledge this is the first borylation of a tert-butyl group
obtained in the reactions of norbornane (17), trans-decalin (23) and in the absence of directing groups11,22,26. Selective reaction at sterically
exo-2-chloronorbornane (24), with no borylation of tertiary C–H bonds hindered methyl groups was also observed for various functionalized
observed. Cyclopentane carbonitrile (26) reacted selectively at the acyclic alkanes, including those containing halides (32–36), protected
methylene distal to the electron-withdrawing nitrile group. amines (37, 38) and nitriles (39). The regioselectivity appeared to be
We next investigated the effectiveness of our protocol for the boryla- sensitive to electronic effects, with no borylation observed at steri-
tion of acyclic alkanes. Although various unfunctionalized substrates cally unhindered methylene groups proximal to electron-withdrawing
were successfully borylated (26–31), we were surprised to observe groups (33, 36–39). For substrates possessing even weaker methine
selective borylation of methyl groups over methylene groups. This C–H bonds, in no cases were tertiary boronic ester products observed.
selectivity is the opposite of that expected in homolytic cleavage of We also tested substrates possessing both C(sp3)–H bonds and steri-
C–H bonds via HAT, where selectivity is typically determined by the cally unhindered aromatic C(sp2)–H bonds, which give high selectively

716 | Nature | Vol 586 | 29 October 2020


for aromatic C–H borylation with iridium catalysis23. Isopropylbenzene H ClB(cat) (20 mol%), 14 (1 equiv.) B(pin)

(40) was borylated with complete selectivity for the methyl groups + B2(cat)2
Si MeCN (0.025 M), 390 nm LEDs Si
over both the aromatic and the much weaker benzylic C–H bonds.
2 then pinacol, Et 3N
Similar selectivity was observed for several other isopropylbenzene (10 equiv.) (1 equiv.)
derivatives (41–42); notably 4-isopropyltoluene (41), which possesses
an unhindered benzylic methyl group8. The reaction was extended
to sterically hindered methyl groups of tert-butylbenzenes (43–44),
which are unreactive under metal-catalysed conditions. Borylation Si B(pin) Si B(pin) Si B(pin)
Cl
of C(sp3)–H bonds distal to aromatic rings was also possible, includ-
54: 47% (40%a, 35%d) 55: 35%, >97:3 r.r.e 56: 30%, >97:3 r.r.
ing neopentyl (45) and cyclohexyl benzene (46), and various heter-
oaromatic carboxylate esters (47, 48). In addition, terminal alkenes EtO
Cl Si B(pin) Br Si B(pin) Si B(pin)
were tolerated, with 4,4-dimethyl-1-pentene reacting at the tert-butyl
O
group to provide boronic ester 49. More complex substrates were 57: 40%, >97:3 r.r.e 58: 39%, >97:3 r.r. 59: 35%, >97:3 r.r.
also successfully borylated, including ibuprofen methyl ester (50), a
Cl
galactose derivative (51), and phthalimide-protected derivatives of
O BzO
the amino acids leucine (52) and tert-leucine (53). Although reactions AcO Si B(pin) Cl Si B(pin) Si B(pin)
were routinely run using 10 equivalents of alkane, the excess alkane O
could be recovered in >80% yield, and in several cases we showed that 60: 30%, >97:3 r.r. 61: 33%, >97:3 r.r. 62: 46%, >97:3 r.r.
5 or 3 equivalents could also be used without a substantial reduction
in yield (17, 29, 30, 32, 36 and 53). Alternatively, improved yields could
O O
be obtained with 20 equivalents of alkane (17). O Si B(pin) Si B(pin)
Usually, C(sp3)–H borylations require substrates to be used in excess, O O
and it was only very recently this limitation was overcome using a modi- 63: 41%, >97:3 r.r. 64: 31%, >97:3 r.r.
fied iridium/phenanthroline ligand system23. This is especially valuable
O
for complex substrates. Therefore, we tested our methodology using
MeO
the alkane as the limiting reagent and found that, under slightly modi-
fied conditions, the complex amino acid-derived product 53 could be O O
Ph O
isolated in comparable yield with 85% recovery of unreacted alkane (for
O O
further alkane stoichiometry studies, see Supplementary Information, 65: 33% (25%a) Si B(pin) 66: 25% Si B(pin)
section 2.7). >97:3 r.r. >97:3 r.r.
Organosilanes are an important compound class that are used exten-
sively in organic synthesis and materials science. It has been previously β
O ( )n α
demonstrated that iridium-catalysed C(sp3)–H borylation of methyl B Si Si B(pin) Et
Si B(pin)
silanes provides a convenient and direct method to access syntheti- O Si B(pin)
Et Et
cally useful (borylmethyl)silanes11. However, that method uses 5–10
67: 33% 68 (n = 0): 50% (45%b, 61%c ) 70: 49%
mol% iridium, and methylsilanes cannot be selectively borylated in the >97:3 r.r. 69 (n = 1): 51%, 96:4 r.r.e α:β = 2:1
presence of sterically unhindered aromatic C–H bonds. By contrast,
our metal-free, HAT-mediated borylation provided complete selec- Fig. 3 | Photoinduced C–H borylations of silanes. Reactions were performed
tivity for reaction of α-silyl C(sp3)–H bonds (54–56, 62–65; Fig. 3). In with 0.3 mmol of B2(cat)2. Yields are of isolated products. r.r. was determined by
1
addition, good functional-group tolerance was demonstrated by the H NMR analysis. Numbers in parentheses show yields obtained using different
silane stoichiometry: a3 equiv., b5 equiv., c20 equiv. dUsing 1.0 equiv. silane and
successful borylation of substrates containing halides (56–58, 61),
1.2 equiv. B2(cat)2. er.r. was determined by GC analysis.
carboxylate esters (59–66), enoates (64), ketones (65), and pinacol
boronic esters (67). Permethyloligosilanes were also borylated in good
yield, including hexamethyldisilane (68) and bis(trimethylsilyl)meth-
ane (69). Interestingly, tetraethylsilane (70) reacted selectively at the cyclisation–borylation sequence (Fig. 4a). Through fluorescence
methylene (α) position, again contrasting with the iridium-catalysed quenching experiments (Supplementary Fig. 9), we determined that
selectivity, which gives only the product of methyl (β) borylation11. As chlorine radicals could be formed via PET between 14 and chloride
before, we explored the effect of stoichiometry of organosilane and anions, where photoexcitation of 14 produces a strongly oxidising
found that although the yields declined with reduced loading (54, 65 excited state (Ep/2(14*/14•−) = 1.46 V versus saturated calomel electrode
and 68), a 35% yield of boronic ester 54 could still be achieved using (SCE) in MeCN, where Ep/2 is the half-peak potential) capable of under-
the organosilane as the limiting reagent. going exergonic single-electron transfer (SET) with chloride anions
To gain insight into the mechanism of this metal-free C–H boryla- (Ep/2(Cl•/Cl−) = 1.00 V versus SCE in MeCN for NEt4Cl). The formation
tion, we performed a series of experiments to confirm the identity of chloride anions from ClB(cat) is supported by NMR experiments
of the HAT species and determine how it is generated. Testing other showing hydrolysis of ClB(cat) by trace water in acetonitrile (Supple-
N-alkoxyphthalimides showed that replacing the trifluoroethoxy group mentary Figs. 18–20). Interestingly, fluorescence quenching experi-
of 14 with a phenoxy group caused only a small reduction in reaction ments showed that the excited state of 14 was quenched with similar
efficiency (Supplementary Table 4). This indicated that alkoxy radi- efficiency by ClB(cat), NEt4Cl and B2(cat)2 (Supplementary Figs. 9–11).
cals are unlikely to be responsible for HAT because the relatively weak The initial concentration B2(cat)2 is high with respect to ClB(cat) or chlo-
O–H bond of phenol (BDE = 90 kcal mol−1), compared to non-activated ride anions, and so the predominant PET process is probably between
C(sp3)–H bonds (BDE = 97–101 kcal mol–1) makes HAT thermodynami- 14 and B2(cat)2.
cally disfavoured25. Instead, the vital role of chloride led us to suspect On the basis of these observations, we propose the mechanism out-
the involvement of chlorine radicals. This was confirmed upon sub- lined in Fig. 4b. After photoexcitation of 14 to 73, reductive quench-
jecting 1,6-heptadiene (71) to modified reaction conditions using ing by B2(cat)2 and subsequent β-scission of radical anion 74 gives the
1 equivalent of ClB(cat), which gave chlorinated (cyclopentylmethyl) trifluoroethoxy radical 75. Oxygen-centred radicals are known to react
boronic ester 72 derived from a chlorine radical addition–5-exo-trig with catechol boronic esters to form radical ‘ate’ complexes, where

Nature | Vol 586 | 29 October 2020 | 717


Article
a B2(cat) 2 (1 equiv.)
Cl B(pin) Via:
ClB(cat) (1 equiv.), 14 (1 equiv)
71 Cl Cl
MeCN (0.025 M), 390 nm LEDs, Cl

then pinacol, Et3N Addition Cyclization


71 72: 46%, 70:30 d.r.

b Initiation

O O * O O
Violet light B2(cat) 2 β-scission
N N N N + CF 3 O
O CF 3 O CF 3 SET O CF 3
O O –B2(cat)2 O O
14 73 74 4 75

B2(cat) 2
Radical equilibrium 78
2

CF 3 O 77
HAT
75 H

15 O
+

O
HCl NH
Cl B
O O 81
Cl O
B B(cat)
CF3 O O
Cl B(cat)
76 79
O

Cl N
+

O 4
O O
O B
O
CF 3 O N Cl B
O CF 3
78 O B(cat)
O
14 80 16

c B2 (cat)2 stoichiometry Borate 78 additive


26α: B(pin) 2.5 2.5
Primary/secondary

Primary/secondary

B2(cat) 2 2 2
ClB(cat) (20 mol%)
selectivity

selectivity

B(pin) 1.5 1.5


+ 14 (1 equiv.) 26β:
1 1
γ α MeCN (0.025 M)
0.5 0.5 CF 3 OB(cat)
β 390 nm LEDs, 78
then pinacol, Et3N B(pin) 0 0
26 0.5 1 1.5 2 0 0.5 1 1.5 2 2.5
B2(cat)2 equiv. Borate 78 equiv.

d O
γ
O β
ClB(cat) (20 mol%), 14 (1 equiv.) γ
β NMe
+ B2(cat)2 + α B(pin) +
N α
Me MeCN (0.025 M), 390 nm LEDs,
then pinacol, Et3N
82 2 83 29: 21% 84: 3%
(10 equiv.) (1 equiv.) (1 equiv.) >97:3 r.r. α:β:γ = 2:76:22

Fig. 4 | Mechanistic studies. a, Evidence for the formation of chlorine radicals. b, Proposed mechanism. c, Effect of B2(cat)2 and borate 78 concentration on
regioselectivity in the borylation of pentane. d, Trapping of tertiary and hindered secondary alkyl radicals.

the unpaired electron is delocalized onto the catecholate ligand34. radicals have been reported to be strong single-electron reductants36,
Therefore, we propose that chlorine radicals are formed through the regeneration of 76 can occur through SET between chloride-stabilized
reaction of 75 with ClB(cat) via radical ‘ate’ complex 76. HAT from 15 boryl radical 80 and 14 (Supplementary Figs. 32–34).
(BDE = 99 kcal mol−1)35 to either a chlorine radical or complex 76 gener- Given that regioselectivities in HAT processes from simple alkanes to
ates hydrochloric acid (BDE = 103 kcal mol−1)25, alkyl radical 77 and trif- chlorine radicals follow the reactivity trend of methine > methylene >
luoroethyl borate 78. Borylation of 77 with B2(cat)2 proceeds via radical methyl37, we proceeded to investigate the origin of the unexpected regi-
complex 79 (as previously described)31, with cleavage of the B–B bond oselectivity in our borylation reaction. Under our standard conditions,
facilitated by reaction with chloride. As Lewis base adducts of boryl pentane was borylated in a 54:46 primary:secondary ratio. However,

718 | Nature | Vol 586 | 29 October 2020


the regioselectivity was found to be dependent on the stoichiometry 8. Shimada, S., Batsanov, A. S., Howard, J. A. K. & Marder, T. B. Formation of aryl- and
benzylboronate esters by rhodium-catalyzed C–H bond functionalization with
of B2(cat)2, with higher equivalents leading to increased secondary pinacolborane. Angew. Chem. Int. Ed. 40, 2168–2171 (2001).
selectivity (Fig. 4c). Conversely, adding trifluoroethyl borate 78 to the 9. Ishiyama, T., Ishida, K., Takagi, J. & Miyaura, N. Palladium-catalyzed benzylic C–H
reaction resulted in an increase in primary selectivity. These reactions borylation of alkylbenzenes with bis(pinacolato)diboron or pinacolborane. Chem. Lett.
30, 1082–1083 (2001).
provide evidence that the regioselectivity is determined during both 10. Liskey, C. W. & Hartwig, J. F. Iridium-catalyzed C−H borylation of cyclopropanes. J. Am.
the HAT and C–B bond-forming steps. Higher primary selectivity with Chem. Soc. 135, 3375–3378 (2013).
lower B2(cat)2 concentration is indicative of competing reaction path- 11. Ohmura, T., Torigoe, T. & Suginome, M. Functionalization of tetraorganosilanes and
permethyloligosilanes at a methyl group on silicon via iridium-catalyzed C(sp3)−H
ways for secondary alkyl radical intermediates, whereby a slower rate borylation. Organometallics 32, 6170–6173 (2013).
of borylation for these more hindered radicals results in deleterious 12. Larsen, M. A., Wilson, C. V. & Hartwig, J. F. Iridium-catalyzed borylation of primary
pathways (for example, single-electron oxidation). Higher primary benzylic C−H bonds without a directing group: scope, mechanism, and origins of
selectivity. J. Am. Chem. Soc. 137, 8633–8643 (2015).
selectivity with higher borate 78 concentration provides indirect 13. Palmer, W. N., Obligacion, J. V., Pappas, I. & Chirik, P. J. Cobalt-catalyzed benzylic
evidence for the formation of radical ‘ate’ complex 76, which undergoes borylation: enabling polyborylation and functionalization of remote, unactivated
primary-selective HAT reactions. To probe the high methyl selectivity C(sp3)−H bonds. J. Am. Chem. Soc. 138, 766–769 (2016).
14. Ros, A., Fernández, R. & Lassaletta, J. M. Functional group directed C–H borylation.
for substrates possessing both methylene and methine groups, we Chem. Soc. Rev. 43, 3229–3243 (2014).
performed the borylation of 2,5-dimethylhexane (82) in the presence 15. Li, Q., Liskey, C. W. & Hartwig, J. F. Regioselective borylation of the C−H bonds in
of an alkene radical trap (N-methyl-N-phenyl-methacrylamide (83); alkylamines and alkyl ethers. Observation and origin of high reactivity of primary C−H
bonds beta to nitrogen and oxygen. J. Am. Chem. Soc. 136, 8755–8765 (2014).
Fig. 4d). This provided boronic ester 29 in 21% yield with >97:3 methyl 16. Larsen, M. A., Cho, S. H. & Hartwig, J. Iridium-catalyzed, hydrosilyl-directed borylation of
(α) selectivity, as well as 3% of oxindole 84, which was formed with high unactivated alkyl C−H bonds. J. Am. Chem. Soc. 138, 762–765 (2016).
selectivity for methine (β) C–H functionalization. This confirms the 17. He, J., Shao, Q., Wu, Q. & Yu, J.-Q. Pd(ii)-catalyzed enantioselective C(sp3)−H borylation.
J. Am. Chem. Soc. 139, 3344–3347 (2017).
non-productive formation of sterically hindered secondary and tertiary 18. Reyes, R. L., Iwai, T., Maeda, S. & Sawamura, M. Iridium-catalyzed asymmetric borylation
alkyl radicals. However, the low yield of 84 and the lack of other tertiary of unactivated methylene C(sp3)−H bonds. J. Am. Chem. Soc. 141, 6817–6821 (2019).
radical-derived side products suggest that the regioselectivity is largely 19. Chen, H. & Hartwig, J. F. Catalytic, regiospecific end-functionalization of alkanes:
rhenium-catalyzed borylation under photochemical conditions. Angew. Chem. Int. Edn
determined during the HAT step. We propose that this is a result of HAT Engl. 38, 3391–3393 (1999).
occurring directly to complex 76, rather than a ‘free’ chlorine radical, 20. Chen, H., Schlecht, S., Semple, T. C. & Hartwig, J. F. Thermal, catalytic, regiospecific
functionalization of alkanes. Science 287, 1995–1997 (2000).
where complexation of the chlorine radical with borate 78 makes a more
21. Murphy, J. M., Lawrence, J. D., Kawamura, K., Incarvito, C. & Hartwig, J. F.
sterically demanding HAT species that is able to selectively functional- Ruthenium-catalyzed regiospecific borylation of methyl C–H bonds. J. Am. Chem. Soc.
ize less hindered but stronger C–H bonds38, thus providing unusually 128, 13684–13685 (2006).
22. Ohmura, T., Torigoe, T. & Suginome, M. Iridium-catalysed borylation of sterically hindered
high primary C–H selectivity.
C(sp3)–H bonds: remarkable rate acceleration by a catalytic amount of potassium
We have introduced an approach to borylations of non-activated tert-butoxide. Chem. Commun. 50, 6333–6336 (2014).
C(sp3)–H bonds for the synthesis of alkylboronic esters from simple 23. Oeschger, R. et al. Diverse functionalization of strong alkyl C–H bonds by undirected
borylation. Science 368, 736–741 (2020).
alkanes. Using a photoinduced HAT strategy, a broad range of alkanes
24. Ciriano, M. V., Korth, H.-G., van Scheppingen, W. B. & Mulder, P. Thermal stability of
were borylated under mild conditions and with regioselectivities 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO) and related N-alkoxyamines. J. Am. Chem.
distinct from those of established metal-catalysed protocols, namely Soc. 121, 6375–6381 (1999).
25. Blanksby, S. J. & Ellison, G. B. Bond dissociation energies of organic molecules.
high selectivity for C(sp3)–H over C(sp2)–H bonds. The reaction also
Acc. Chem. Res. 36, 255–263 (2003).
provides a rare example of a radical-mediated C–H functionalization 26. Prokofjevs, A. & Vedejs, E. N-directed aliphatic C–H borylation using borenium cation
proceeding with high selectivity for substitution of methyl C–H bonds equivalents. J. Am. Chem. Soc. 133, 20056–20059 (2011).
27. Fawcett, A. et al. Photoinduced decarboxylative borylation of carboxylic acids. Science
over weaker secondary, tertiary and benzylic C–H bonds. Although
357, 283–286 (2017).
further developments are needed to improve the modest yields and 28. Guo, J.-J., Hu, A. & Zuo, Z. Photocatalytic alkoxy radical-mediated transformations.
high substrate loading, this radical-mediated C–H borylation enables Tetrahedr. Lett. 59, 2103–2111 (2018).
29. Kim, S., Lee, T. A. & Song, Y. Facile generation of alkoxy radicals from
the transformation of feedstock chemicals into valuable organoboron N-alkoxyphthalimides. Synlett 1998, 471–472 (1998).
products with selectivities that complement those of existing methods. 30. Zhang, J., Li, Y., Zhang, F., Hu, C. & Chen, Y. Generation of alkoxyl radicals by photoredox
catalysis enables selective C(sp3)–H functionalization under mild reaction conditions.
Angew. Chem. Int. Ed. 55, 1872–1875 (2016).
31. Cheng, Y., Mück-Lichtenfeld, C. & Studer, A. Transition metal-free 1,2-carboboration of
Online content unactivated alkenes. J. Am. Chem. Soc. 140, 6221–6225 (2018).
Any methods, additional references, Nature Research reporting sum- 32. Hu, A., Guo, J.-J., Pan, H. & Zuo, Z. Selective functionalization of methane, ethane, and
higher alkanes by cerium photocatalysis. Science 361, 668–672 (2018).
maries, source data, extended data, supplementary information, 33. Qiao, Y., Yang, Q. & Schelter, E. J. Photoinduced Miyaura borylation by a rare-earth-metal
acknowledgements, peer review information; details of author con- photoreductant: the hexachlorocerate(iii) anion. Angew. Chem. Int. Ed. 57, 10999–11003
tributions and competing interests; and statements of data and code (2018).
34. Baban, J. A., Goodchild, N. J. & Roberts, B. P. Electron spin resonance studies of
availability are available at https://doi.org/10.1038/s41586-020-2831-6. radicals derived from 1,3,2-benzodioxaboroles. J. Chem. Soc. Perkin Trans. 2 1986,
157–161 (1986).
35. Nunes, P. M. et al. C–H bond dissociation enthalpies in norbornane. An experimental and
1. Hall, D. G. (ed.) Boronic Acids: Preparation and Applications in Organic Synthesis Medicine
computational study. Org. Lett. 10, 1613–1616 (2008).
and Materials (Wiley, 2011).
36. Sandfort, F., Strieth-Kalthoff, F., Klauck, F. J. R., James, M. J. & Glorius, F. Deaminative
2. Mkhalid, I. A. I., Barnard, J. H., Marder, T. B., Murphy, J. M. & Hartwig, J. F. C–H activation for
borylation of aliphatic amines enabled by visible light excitation of an electron
the construction of C–B bonds. Chem. Rev. 110, 890–931 (2010).
donor–acceptor complex. Chem. Eur. J. 24, 17210–17214 (2018).
3. Xu, L. et al. Recent advances in catalytic C–H borylation reactions. Tetrahedron 73,
37. Tedder, J. M. Which factors determine the reactivity and regioselectivity of free radical
7123–7157 (2017).
substitution and addition reactions? Angew. Chem. Int. Edn Engl. 21, 401–410 (1982).
4. Hartwig, J. F. Regioselectivity of the borylation of alkanes and arenes. Chem. Soc. Rev.
38. Carestia, A. M., Ravelli, D. & Alexanian, E. J. Reagent-dictated site selectivity in
40, 1992–2002 (2011).
intermolecular aliphatic C–H functionalizations using nitrogen-centered radicals.
5. Capaldo, L. & Ravelli, D. Hydrogen atom transfer (HAT): a versatile strategy for substrate
Chem. Sci. 9, 5360–5365 (2018).
activation in photocatalyzed organic synthesis. Eur. J. Org. Chem. 2017, 2056–2071 (2017).
6. Cho, J.-Y., Tse, M. K., Holmes, D., Maleczka, R. E., Jr & Smith, M. R., III. Remarkably selective
iridium catalysts for the elaboration of aromatic C–H bonds. Science 295, 305–308 (2002). Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
7. Ishiyama, T. et al. Mild iridium-catalyzed borylation of arenes. High turnover numbers, published maps and institutional affiliations.
room temperature reactions, and isolation of a potential intermediate. J. Am. Chem. Soc.
124, 390–391 (2002). © The Author(s), under exclusive licence to Springer Nature Limited 2020

Nature | Vol 586 | 29 October 2020 | 719


Article
Author contributions A.N. and V.K.A. conceived the project, directed the research and prepared
Data availability the manuscript; C.S. performed the experimental work; all authors analysed the results.

Materials and methods, experimental procedures, characterization Competing interests The authors declare no competing interests.
data, spectra and additional mechanistic discussions are available in
Additional information
the Supplementary Information.
Supplementary information is available for this paper at https://doi.org/10.1038/s41586-020-
2831-6.
Correspondence and requests for materials should be addressed to A.N. or V.K.A.
Acknowledgements We thank the EPSRC (EP/R004978/1) for funding. We gratefully Peer review information Nature thanks the anonymous reviewers for their contribution to the
acknowledge A. Sedikides and A. Lennox (University of Bristol) for performing cyclic peer review of this work. Peer reviewer reports are available.
voltammetry experiments. Reprints and permissions information is available at http://www.nature.com/reprints.

You might also like