Sie sind auf Seite 1von 6

Letter https://doi.org/10.

1038/s41586-019-1539-y

Hindered dialkyl ether synthesis with


electrogenerated carbocations
Jinbao Xiang1,2,9, Ming Shang1,9, Yu Kawamata1, Helena Lundberg1,3, Solomon H. Reisberg1, Miao Chen1, Pavel Mykhailiuk1,4,5,
Gregory Beutner6, Michael R. Collins7, Alyn Davies8, Matthew Del Bel7, Gary M. Gallego7, Jillian E. Spangler7, Jeremy Starr8,
Shouliang Yang7, Donna G. Blackmond1 & Phil S. Baran1*

Hindered ethers are of high value for various applications; however, presence of the requisite secondary alcohol—provides the ether in only
they remain an underexplored area of chemical space because 11% yield8.
they are difficult to synthesize via conventional reactions1,2. Distinct from sterically sensitive SN2 and strongly acidic carboca-
Such motifs are highly coveted in medicinal chemistry, because tion pathways, there is a third class of ether synthesis that has been
extensive substitution about the ether bond prevents unwanted known for many years but has remained largely underexplored. This
metabolic processes that can lead to rapid degradation in vivo. class (Fig. 1a, yellow inset) stems from the oldest synthetic organic
Here we report a simple route towards the synthesis of hindered electrochemical reaction, the Kolbe dimerization, which was discov-
ethers, in which electrochemical oxidation is used to liberate ered9 in 1847. In the so-called interrupted Kolbe variant, known as
high-energy carbocations from simple carboxylic acids. These the Hofer–Moest reaction10, electrolytic oxidation of a carboxylic acid
reactive carbocation intermediates, which are generated with low under mildly alkaline conditions generates a carbocation that can be
electrochemical potentials, capture an alcohol donor under non- captured by incipient nucleophiles10–18. A distinct advantage of this
acidic conditions; this enables the formation of a range of ethers reaction is the non-acidic generation of high-energy carbocations
(more than 80 have been prepared here) that would otherwise be directly from carboxylic acids.
difficult to access. The carbocations can also be intercepted by We were therefore surprised to find a dearth of applications of the
simple nucleophiles, leading to the formation of hindered alcohols Hofer–Moest reaction in synthetic contexts (Fig. 1b). Indeed, much
and even alkyl fluorides. This method was evaluated for its ability to more complex catalytic systems that take advantage of photolytic con-
circumvent the synthetic bottlenecks encountered in the preparation ditions have been developed to make alkyl–aryl ethers19. There are
of 12 chemical scaffolds, leading to higher yields of the required probably two reasons for the limited exploration of electrolytic decar-
products, in addition to substantial reductions in the number of boxylative ether synthesis: first, the barrier to entry into electrosynthesis
steps and the amount of labour required to prepare them. The use has traditionally been high for a practicing synthetic organic chem-
of molecular probes and the results of kinetic studies support the ist20. Second, all Hofer–Moest systems known so far have required
proposed mechanism and the role of additives under the conditions solvent-quantities of the alcoholic nucleophile, which is untenable for
examined. The reaction manifold that we report here demonstrates complex ethereal substrates. The alcoholic solvent—in addition to func-
the power of electrochemistry to access highly reactive intermediates tioning as a reagent—permits current to pass and also acts as an electron
under mild conditions and, in turn, the substantial improvements sink to balance the electrochemical process, liberating hydrogen gas.
in efficiency that can be achieved with these otherwise-inaccessible Herein, we report how the generation of carbocations from unac-
intermediates. tivated, aliphatic carboxylic acids—and their subsequent capture by
The Williamson ether synthesis3,4 is a long-established method by heteroatom nucleophiles—can be leveraged to provide a wide array of
which to synthesize primary alkyl ethers via SN2 substitution (Fig. 1a). hindered C–X bonds.
However, in contexts involving secondary or tertiary alkyl halides the Several key literature precedents can be highlighted that aided this
reaction often derails, leading to elimination byproducts or to no development. It is known that carbon-based electrodes favour the
reaction at all. Hindered ether 1, which is a key intermediate in the desired carbocation generation, whereas platinum electrodes favour
synthesis of an aurora kinase modulator, exemplifies this commonly unproductive radical (Kolbe-type) dimerization21. It is also known that
faced challenge. Despite the documented utility of hindered ethers1,2, inert, non-oxidizable anions (for example, ClO4−) enhance cation-like
very little progress has been made in facilitating access to them. The reactivity in the Hofer–Moest reaction22. Mildly alkaline conditions
alternative workhorse method, the Mitsunobu reaction, also fails in are known to be beneficial for the desired carbocation formation22,23.
such settings owing to the steric demands of the SN2 process and the Finally, simple undivided cells are generally used in this process, which
pKa requirements of the nucleophile5. To the best of our knowledge, suggests that cathodic reduction would not interfere substantially with
the most frequently used method for the synthesis of hindered dialkyl the reaction.
ether bonds still uses carbocation chemistry accessed from olefins It became immediately apparent that limiting the amount of alco-
(hydroalkoxylation) under strongly acidic conditions6. Although this hol posed several considerable challenges: the decomposition of the
transformation has been known for nearly a century, its use is sub- carbocation due to the low nucleophilicity of alcohols; the competitive
stantially limited in scope owing to sluggish reactivity and a lack of trapping of the carbocation by water; the consumption of alcohols by
chemoselectivity7. For example, in order to prepare hindered ether 1, anodic oxidation; and the necessity of an external electron-acceptor in
a multi-step route via 4-hydroxyproline 2 (R = CO2Me) was used. The order to balance electrons. Figure 1c summarizes the results of around
synthesis required over 6 days of reaction time and gave the required 1,000 experiments (see Supplementary Information for an extensive
product in less than 4% overall yield; the key C–O bond-forming reac- sampling) that were undertaken in order to solve these problems.
tion—the treatment of methylenecyclobutane with BF3·Et2O in the Not surprisingly, initial exploratory experiments using the proposed
1
Department of Chemistry, Scripps Research, La Jolla, CA, USA. 2The Center for Combinatorial Chemistry and Drug Discovery of Jilin University, The School of Pharmaceutical Sciences, Jilin
University, Jilin, People’s Republic of China. 3Department of Chemistry, KTH Royal Institute of Technology, Stockholm, Sweden. 4Enamine Ltd, Kiev, Ukraine. 5Chemistry Department, Taras
Shevchenko National University of Kyiv, Kiev, Ukraine. 6Chemical and Synthetic Development, Bristol-Myers Squibb, New Brunswick, NJ, USA. 7Department of Chemistry, La Jolla Laboratories,
Pfizer Inc, San Diego, CA, USA. 8Pfizer Medicinal Sciences, Groton, CT, USA. 9These authors contributed equally: Jinbao Xiang, Ming Shang. *e-mail: pbaran@scripps.edu

N A t U r e | www.nature.com/nature
RESEARCH Letter

a Pathways to hindered dialkyl ethers b Previous decarboxylative etherifications


Me
Photochemistry
X
R1 [Ir] cat. R1
X = LG or OH R2 R2
+ [Cu] cat.
SN2; problematic on hindered tertiary systems R3 CO2A* HO R3 O
FG hQ FG
Me Alkyl NHPI esters Phenols Alkyl aryl ethers
O BF3-cat. addition HO
NCbz (11% yield) NCbz
Hydroalkoxylation; <4% overall yield (>6 d) Electrochemistry
1, kinase R A
OH 2 R1
modulator intermediate R1 CO2H O
Me + Alkyl–OH Alkyl
CO2H R2 R2
R3 Anodic oxidation R3
This work

Direct decarboxylative etherification Typical acid scope Alcohol scope (solvent)


A R2 HO2C OH
X R 1 R1 MeOH
R1 + R3 R–OH R1 O R6 CO2H
R1 CO2H R2
2 R2 R5 Me OH
R Anodic CO2H
R3 R2 R3 R4 OH
oxidation X = N, O, Si
Alkyl acid Carbocation Hindered ethers Me Me
HO
Alkyl–CO2H (Me, nPr, etc.)
• Ubiquitous • Reactive • Difficult to access
• Inexpensive • Non-acidic generation • Medicinally important

c Reaction optimization
Byproducts (from 3)
Me Base Me Me Me
Me Me Me Ph Me Me
Electrolyte (0.1 M) Ph + Me
+ Ph H
Ph CO2H HO Ph Additive, solvent, RT Ph O Ph Ph OH
Me Me
3 4 Me 5
+C/–C, 10 mA, 3 h 6 7 8
(0.2 mmol) (0.6 mmol)
From radical Elimination Hydration
Entry Conditionsa Yield (%)b 5 6 7 8 9 10
1 Et3N, nBu4NClO4, DMF or acetone <1 – – <1 <1 –
2 K2CO3, nBu4NClO4, DMF <1 15 <1 40 <1 – Byproducts (from 4)
3 2,4,6-collidine, nBu4NPF 6, DMF 6 3 28 14 <1 – Me Me O
via Me
4 2,4,6-collidine, nBu4NPF 6, CH2Cl2 40 4 - 14 6 – Ph
5 2,4,6-collidine, nBu4NPF 6, 3 Å MS, CH2Cl2 43 5 2 <1 3 – O Ph Ph O +

6c AgPF6, 2,4,6-collidine, nBu4NPF6, 3 Å MS, CH2Cl2 79 (77)d – – – 7 – Me Me Ph


9 10
7 As in entry 6, 1 equiv. alcohol 34 <1 <1 <1 6 <1 Anodic oxidation Ionization
8 AgPF6, nBu4NPF6, 3 Å MS, CH2Cl2 – <1 <1 – 22 54

Fig. 1 | Background and reaction development. a, The synthesis of 3 (0.2 mmol), 3.0 equiv. of alcohol 4 (except where designated). bYield
hindered ethers is a long-standing challenge in organic synthesis. based on gas chromatography. All entries were performed in triplicate.
c
A; constant-current electrolysis; cat., catalyst; Cbz, carboxybenzyl; Conditions: acid 3 (0.2 mmol), alcohol 4 (0.6 mmol), AgPF6 (0.3 mmol),
LG, leaving group. b, Historical context and previous strategies for 2,4,6-collidine (0.6 mmol), nBu4NPF6 (0.1 M), 3 Å molecular sieves
decarboxylative etherification. FG, functional group; NHPI, (150 mg), dichloromethane (CH2Cl2; 3 ml), current (I) = 10 mA, 3 h.
d
N-hydroxyphthalimide. c, Development and optimization of hindered Isolated yield. DMF, N,N-dimethylformamide; RT, room temperature;
ether synthesis depicted through electromechanistic analysis. aCompound +C/−C represents the graphite electrodes.

conversion of 3 and 4 to 5 as an example—based on the literature prece- The patent literature is replete with examples of hindered ethers in
dent available11,22—led to only trace amounts of product (entry 1). The various pharmaceutical and materials applications. Although carbo-
conversion of the carboxylate was improved by choosing potassium cation-based routes from olefins predominate in the literature, elec-
carbonate or 2,4,6-collidine as a non-oxidizable base (entries 2, 3), trochemical access to such valuable entities has notable advantages in
although 6 and 7 were identified as major byproducts due to the pres- terms of the time required, the step count and the overall yield. Figure 2
ence of radical intermediate22 and elimination pathways, respectively. shows an abbreviated depiction of six such applications as well as the
These problems were effectively suppressed by changing the solvent more than 80 ethers prepared (see Supplementary Information for a full
to dichloromethane (entry 4), which resulted in a large increase of the listing of substrates as well as a comparison to previous routes). Primary
desired ether product. In dichloromethane, hydration of the carboca- carboxylic acids and certain secondary systems are not compatible with
tion (leading to 8) persisted (entry 4), which was suppressed by the electrochemical etherification, because the resultant carbocations are
addition of 3 Å molecular sieves (entry 5). It was also found that dichlo- not sufficiently stable. However, this limitation is inconsequential from
romethane was apparently reduced at the cathode (Supplementary a synthetic perspective, as those ethers can be easily prepared through
Fig. 2), acting as an electron sink. Past approaches using water or simple standard SN2-type approaches3–5. Acids bearing various functional
alcohols as the solvent did not need to address this issue, as the solvent groups are tolerated, such as Boc-protected amines (17), aryl and alkyl
can serve as both the reagent and an electron sink (via proton reduc- halides (25, 27, 30), olefins (26), esters (29, 39), enones (39), ethers (35,
tion). Accordingly, the addition of a better sacrificial oxidant (silver 36, 62) and even oxidation-prone boronic esters (28). Similarly, the
hexafluorophosphate, AgPF6) considerably improved the yield of 5, scope of alcohol coupling partners is vast and includes acid-labile and
leading to our optimized conditions (entry 6). The addition of AgPF6 oxidation-prone chiral secondary benzylic alcohols (17), deuterated
also completely suppressed the formation of 6 and 7, although this systems (49), azetidinyl alcohols (50), protected sugars (53) and olefin-
effect was found to vary by substrate and—in some cases—silver addi- containing alcohols (48, 51, 59), as well as alcohols containing acetals and
tives are not necessary at all. Negative controls confirmed the necessity esters (44), halides (15), nitriles (45), nitro groups (see Supplementary
of a slight stoichiometric excess of the alcoholic partner (entry 7). In Information) and even Lewis-basic heterocycles (42, 43, 45). The ability
the absence of base, no desired product was observed, with the major to tolerate chiral, ionizable secondary alcohols is worth emphasizing,
products being the ketone 9 and the ester 10 (entry 8). because acidic methods for ether synthesis using such alcohols would

N A t U r e | www.nature.com/nature
Letter RESEARCH

O 2,4,6-collidine (3.0 equiv.)


R4 n
Bu4NPF 6 (0.1 M) R1 R4
R1 R5 R2 R5
OH +
R2 HO R6 AgPF 6 (1.5 equiv.) R3 O R6
R3 3 Å MS, CH2Cl2 (3 ml), RT
(0.2 mmol, 1.0 equiv.) (3.0 equiv.) +C/–C, 10 mA, 3 h
Applicationsa
Me CO2Me CO2Me Me Me
O BocN O O Me
NCbz O Br
O Me O
NHCbz NHCbz
Me O Me
Me Me Me Me O
OH
1 11 12 13 14 15
Current route: 2 steps 1 step 1 step 1 step 1 step 1 step
(51%, 15 h) (32%, 3 h) (40%, 3 h) (21%, 3 h) (42%, 3 h) (81%, 3 h)
Previous route: 3 steps 5 steps 7 steps 6 steps 4 steps 2 steps
(<4% overall, >6 d) (31% overall, 2.5 d) (37% overall, >3 d) (24% overall, 2 d) (47% overall, >2 d) (<2% overall, >5 d)
Tertiary acids Secondary alcohols k CD3
Me Me
Me Me D
Me Me Me Me
O O O O CD3
O O Me Me 49: 82%j
Me
Me BocN Me Me
n Me O Me Me Me NCbz
16, 65% 17, 45% b,c n = 2, 18: 54%b,c / 40%c,d
95% e.e. n = 1, 19: 40% b,c
/ 32% c,d CF 3 N Me O
R Ts
46, 60%b 47, 39%e 48, 32%e 50, 70%f
O Cl
O O Cl
R O
Cl Me Me
Me Me Me Me
Me Me Me Me Me
R = H, 22: 61%e
R = nPr, 20: 52%b,c 25, 24%f,g / 43%f,g,h Me
b
R = Me, 23: 58% / 62%b,j / 50%d Me H Me
R = Bn, 21: 41% O
R = PhCH2CH2, 24: 35% b,c
/ 28% c,d
NBoc Me
R R Me Me Me O Me
O Me
H H
CO2Me O O Me
O O Ph e Ph
O 52, 66%
Ph O 51, 54%e
Me Me O
X 53, 52%f O
Me Me Me Me
26, 43%f,g / 31%d R = R = –(CH2) 5–, X = F, 27, 46%f
5, 77% j / 78%f / 54%i R = R = Me, X = BPin, 28, 54%f Tertiary alcoholsk Me

Me Me
Secondary acids BocN Me Me Me Me
Cl Me Me
CO2Me
O Me O Me O Me O
O Me
O Me Cl
54, 42% 55, 65% 56, 52%
Me
Me Me Me
29, 41%f,g 30, 74%, d.r. = 1:1 Me
31, 68%f
Me Me Me O Me
Me
O Me Me Me Me
O O
O 16
O
Me CF 3 Me
Me 57, 46%f 59, 28%b,c
58, 65% j/54%i
32, 80% / 48%f,j i 33, 57%f 34, 7%

α-heteroatom acids Fluorinated ethersk


O Me CF 3 OMe Me
O F F F F NCbz
BocN
O N O O Ph
O O O
O
35, 58% O 37, 82%f
36, 55%j 60, 46%f,g 61, 23%f 62, 88%f,g, d.r. = 1:1
Drugs/natural products CF 2H
O Me
Me Me O
Me H CF 3 OMe Me Me Me CH2F
Me O Ph
Me O AcO H Me O Me O CH2F
H
Me Me
Me Me Me
63, 42%f,g 64, 51%f,g 65, 42%f,g,j
38, 90% j, d.r. = 1:1 39, 35%b,c, d.r. = 1:1
Primary alcoholsk
PEG ethers
Ar H
O Me N N Ph Me Me
Ph O
Me Me O OH
Me Me Ph O S O O O
Cl OMe
O 41, 59%j O Me
H Ar = 2,4-diF 66, 59%f
Me Me Me Cl Me
40, 48%f 42, 68%e –Ph R O
Me Me O O O Me
O Me NC
Ph O Me
Me O
Me Me
O O O Me Et CN BocN
N Me Me
N
Ph O N Cl
Me Me Ph R =
O O Me
N O
Me Me 44, 52%f Me 67, 48%e,j
Me 45, 42%e e 69, 25%b,c
OMe Et 68, 57%
43, 50%f

Fig. 2 | Applications, and partial scope of hindered ether synthesis molecular sieves (100 mg), CH2Cl2 (2 ml), I = 10 mA, 3 h. fAgClO4
via electrochemical decarboxylation. aSee Supplementary Information (0.6 mmol) instead of AgPF6, nBu4NClO4 (0.1 M) instead of nBu4NPF6. g4.0
for literature routes. bAgSbF6 (0.3 mmol) instead of AgPF6. cDBU or 6.0 equiv. alcohol. h1.5 ml CH2Cl2, I = 7.5 mA, 4 h. i nBu4NClO4 (0.1 M)
(1,8-diazabicyclo[5.4.0]undec-7-ene; 0.6 mmol) instead of 2,4,6-collidine. instead of nBu4NPF6, no AgPF6. jReaction performed in triplicate; yield is
d
KSbF6 (0.3 mmol) instead of AgPF6. eAlcohol as limiting reagent, average of three runs. kSee Supplementary Information for more examples.
conditions: alcohol (0.15 mmol), carboxylic acid (0.45 mmol), AgClO4 Boc, tert-butyloxycarbonyl; d.r., diastereomeric ratio; MS, molecular
(0.6 mmol), 2,4,6-collidine (0.675 mmol), nBu4NClO4 (0.2 M), 3 Å sieves; Ts, tosyl.

N A t U r e | www.nature.com/nature
RESEARCH Letter

O 2,4,6-collidine (1.5 equiv.) CO2Me CO2Me


nBu NPF (0.1 M) R1 Nu Commercially
R1 + Nu
4 6
OH R 2 available
Acetone (3 ml), RT
R2 R3 HO2C 70 HO2C 71
R3 +C/–C (10 mA), 3 h
(0.2 mmol, 1.0 equiv) Me OH
F

Applicationsa F H
O
CO2H CO2Me NH2 CO2Me Me Me Me

H H Me
MeO HO MeO HO HO HO
72 73 74 75 76 77
Me Me
Current route: 2 steps 1 step 2 steps 2 steps 3 steps 1 step
(56%, 9 h) (66%, 3 h) (31%, 24 h) (22%, 27 h) (17%, 21 h) (61%, d.r. = 1.1:1, 3 h)
Previous route: 7 steps 9 steps 7 steps 14 steps 7 steps 5 steps
(21% overall, >4 d) (15% overall, >5 d) (12% overall, 3 d) (5% overall, >9 d) (8% overall, 62 h) (yield not reported)
Me
Tertiary and secondary acidsb
OH Me
Me OH Me
18 Me Me Me O
OH Me OH
Me Me OH
OH Me
R H N
OH Me
N Me
MeO2C R R O
Me Me OH
c c c
R = Ph, 78, 52% 80, 95%c 81, 61% yield R = Boc, 82: 32% R = Br, 84: 67% c
86, 53% from 87, 32%c (d.r. = 3:1) 88, 40% c
R = PhO, 79, 66%c (67.1% R = Ts, 83: 69%c R = Bpin, 85: 55%c (1S)-(–)-camphanic acid from dehydroabietic acid from indoprofen
18
O labelled)c
Esterification Fluorination Amidation
F Me F O Me H
F F O F F O Me Me
N

O F N F
O CO2Me O
N
Cl Cl Boc O

89, 31%d 90, 36%d 91, 35%e 92, 58%e 93, 18%e 94, 62%e 95, 14%d

Scale-up

Me Me Me Me A
CO2H OH
Me Me
+ HO Ph O + H2O
Ph CO2H
MeO2C MeO2C
3 (±)-4 71
1.97 g (12 mmol) 4.40 g (36 mmol) (±)-5, 2.08 g (72%)f 1.53 g (7.2 mmol) 0.6 ml 81, 0.86 g (65%)g

Fig. 3 | Applications and partial scope of trapping electrogenerated conditions: 18-crown-6 (0.72 mmol, 3.6 eq), AgClO4 (0.6 mmol, 3 equiv.),
carbocations with other nucleophiles along with scalability 2,4,6-collidine (0.6 mmol, 3 equiv.), nBu4NPF6 (0.1 M), 3 Å MS (150 mg),
demonstration. aSee Supplementary Information for literature routes. CH2Cl2 (3 ml). fConditions for scale-up to (±)-5 (each reaction): 3
b
See Supplementary Information for more examples. cH2O (0.1 ml) as (2.4 mmol), (±)-4 (7.2 mmol), 2,4,6-collidine (3.6 mmol), nBu4NClO4
nucleophile. dCarboxylic acid or phenylacetonitrile as nucleophile (0.1 M), 3 Å molecular sieves (450 mg), CH2Cl2 (9 ml) +C/−C (10 mA),
(0.6 mmol, 3 equiv.), conditions: AgClO4 (0.6 mmol, 3 equiv.), RT, 15 h. gConditions for scale-up to 81 (each reaction): 71 (1.2 mmol),
2,4,6-collidine (0.6 mmol, 3 equiv.), nBu4NClO4 (0.1 M), 3 Å molecular H2O (0.1 ml), 2,4,6-collidine (1.8 mmol), nBu4NPF6 (0.02 M), acetone
sieves (150 mg), CH2Cl2 (3 ml). eKF (0.72 mmol, 3.6 equiv.) as nucleophile, (9 ml), +C/−C (10 mA), RT, 12 h. Nu, nucleophile; pin, pinacolato.

lead to elimination or racemization (17 is formed in 95% enantiomeric synthetic pathways. Six such examples (72–77) are illustrated in Fig. 3
excess (e.e.)). Electrogenerated cations bearing fluorine atoms can also (see Supplementary Information).
be intercepted, opening up a range of organofluorine-containing ether The addition of water to generate various tertiary and secondary
systems that were previously either difficult to access or unknown alcohols is a broadly applicable process, with selected examples depicted
(see Supplementary Information for full listing of fluorinated ethers). in Fig. 3 (for additional examples, see Supplementary Information).
Finally, the synthesis of polyethyleneglycol (PEG) ethers—which histori- Again, the chemoselectivity is on par with that observed in the syn-
cally has been laborious—can be achieved in a modular fashion using this thesis of hindered ethers: aryl bromides (84), boronic esters (85),
process (no PEG ethers analogous to 66–69 are known). To ensure the electron-rich aromatics (87), lactams (88), ethers (79) and esters (81)
robustness of the process, ten randomly selected examples in Fig. 2 were are tolerated. In preliminary studies, the possibility of adding other
run in triplicate, with yields varying by no more than 5% between runs. nucleophiles to the putative electrogenerated carbocations was also
As mentioned above, the use of simple alcohols such as methanol— studied. Carboxylates that are not capable of decarboxylation under
as well as water—is already known in electrochemical decarboxyla- these conditions could act as nucleophiles, thus providing hindered
tive processes10–18 (Fig. 1b). However, the scope of such processes is esters (89, 90)24. Useful organofluorine building blocks could also be
quite limited. The conditions developed here were therefore adapted accessed (91–94) in a process that might be of use in radiolabelling
for these related reactions (Supplementary Information). In order to studies, because the fluorine source used (the inexpensive salt potas-
render this reaction general, the choice of electrolyte (tetrabutylam- sium fluoride) is preferred in such situations25. Finally, using benzoni-
monium hexafluorophosphate, nBu4NPF6) and base (2,4,6-collidine) trile as a nucleophile led to the expected Ritter-type product (95), albeit
was crucial, whereas silver additives were found to be unnecessary. As in lower yield26. The robust and practical nature of this process was
with the synthesis of hindered ethers, the most convincing case for also demonstrated by the gram-scale preparation of 5 and 81 through
the use of this reaction stems from its ability to substantially truncate etherification and hydroxylation processes, respectively. In the case of

N A t U r e | www.nature.com/nature
Letter RESEARCH

large-scale etherification, the silver salt additive could be left out with 11. Corey, E. J., Bauld, N. L., La Londe, R. T., Casanova, J., Jr & Kaiser, E. T. Generation
of cationic carbon by anodic oxidation of carboxylic acids. J. Am. Chem. Soc. 82,
a only minimal effect on yield (78% in the presence of silver salt, com- 2645–2646 (1960).
pared with 72% when this additive is omitted). 12. Luo, X., Ma, X., Lebreux, F., Markó, I. E. & Lam, K. Electrochemical
Extensive mechanistic studies were also undertaken in order to under- methoxymethylation of alcohols – a new, green and safe approach for the
preparation of MOM ethers and other acetals. Chem. Commun. 54, 9969–9972
stand the role of the additives and the nature of the reactive interme- (2018).
diate. In summary, the mechanism is likely to involve the rate-limiting 13. Bunyan, P. J. & Hey, D. H. The electrolysis of some aryl-substituted, aliphatic
oxidation of a carboxylate on the anode to generate a carbocation, acids. J. Chem. Soc. 1360–1365 (1962).
14. Iwasaki, T., Hrorikawa, H., Matsumoto, K. & Miyoshi, M. Electrochemical
followed by nucleophilic attack by an alcohol to afford the ether product synthesis and reactivity of α-alkoxy α-amino acid derivatives. Bull. Chem. Soc.
(see Supplementary Information for full details). Jpn. 52, 826–830 (1979).
It is anticipated that the mild electrogeneration of carbocations 15. Tajima, T., Kurihara, H. & Fuchigami, T. Development of an electrolytic system for
reported herein will find use in numerous settings in which standard non-Kolbe electrolysis based on the acid−base reaction between carboxylic
acids as a substrate and solid-supported bases. J. Am. Chem. Soc. 129,
SN2 and carbocation-based approaches are unsuccessful in forming 6680–6681 (2007).
hindered functionalized carbogenic frameworks. 16. Shtelman, A. V. & Becker, J. Y. Electrochemical synthesis of 1,2-disilylethanes
from α-silylacetic acids. J. Org. Chem. 76, 4710–4714 (2011).
17. Torii, S., Inokuchi, T., Mizuguchi, K. & Yamazaki, M. Electrolytic decarboxylation
Online content reactions. 4. Electrosyntheses of 3-alkyl-2-cycloalken-1-ol acetates from
Any methods, additional references, Nature Research reporting summaries, 1-alkyl-2-cycloalkene-1-carboxylic acids. Preparation of dl-muscone from
source data, extended data, supplementary information, acknowledgements, peer cyclopentadecanone. J. Org. Chem. 44, 2303–2307 (1979).
review information; details of author contributions and competing interests; and 18. Coleman, J. P., Lines, R., Utley, J. H. P. & Weedon, B. C. L. Electro-organic
statements of data and code availability are available at https://doi.org/10.1038/ reactions. Part II. Mechanism of the kolbe electrolysis of substituted
phenylacetate ions. J. Chem. Soc., Perkin Trans. 2 1064–1069
s41586-019-1539-y. (1974).
Received: 21 March 2019; Accepted: 12 July 2019; 19. Mao, R., Balon, J. & Hu, X. Decarboxylative C(sp3)–O cross-coupling. Angew.
Chem. Int. Ed. 57, 13624–13628 (2018).
Published online xx xx xxxx. 20. Yan, M., Kawamata, Y. & Baran, P. S. Synthetic organic electrochemical methods
since 2000: on the verge of a renaissance. Chem. Rev. 117, 13230–13319
1. Roughley, S. D. & Jordan, A. M. The medicinal chemist’s toolbox: an analysis of (2017).
reactions used in the pursuit of drug candidates. J. Med. Chem. 54, 3451–3479 21. Ross, S. D. & Finkelstein, M. Anodic oxidations. V. The Kolbe oxidation of
(2011). phenylacetic acid and 1-methylcyclohexaneacetic acid at platinum and at
2. Fischer, J. & Ganellin, C. R. (eds) Analogue-based Drug Discovery 206–217 (Wiley, carbon. J. Org. Chem. 34, 2923–2927 (1969).
2006). 22. Schäfer, H. J. Recent contributions of Kolbe electrolysis to organic synthesis.
3. Williamson, W. Ueber die theorie der aetherbildung. Justus Liebigs Ann. Chem. Top. Curr. Chem. 152, 91–151 (1990).
77, 37–49 (1851). 23. Koehl, W. J. Anodic oxidation of aliphatic acids at carbon anodes. J. Am. Chem.
4. Kürti, L. & Czakó, B. Strategic Applications of Named Reactions in Organic Soc. 86, 4686–4690 (1964).
Synthesis 484–485 (Elsevier, 2005). 24. Iwasaki, T., Horikawa, H., Matsumoto, K. & Miyoshi, M. An electrochemical
5. Swamy, K. C. K., Kumar, N. N. B., Balaraman, E. & Kumar, K. V. P. P. Mitsunobu and synthesis of 2-acetoxy-2-amino acid and 3-acetoxy-3-amino acid derivatives.
related reactions: advances and applications. Chem. Rev. 109, 2551–2651 (2009). J. Org. Chem. 42, 2419–2423 (1977).
6. Beyerman, H. C. & Heiszwolf, G. J. Reaction of steroidal alcohols with isobutene. 25. Huang, X., Liu, W., Hooker, J. M. & Groves, J. T. Targeted fluorination with the
Usefulness of t-butyl as a hydroxyl-protecting group in a synthesis of fluoride ion by manganese-catalyzed decarboxylation. Angew. Chem. Int. Ed. 54,
testosterone. Recl. Trav. Chim. Pays-Bas 84, 203–212 (1965). 5241–5245 (2015).
7. Smith, M. B. & March, J. March’s Advanced Organic Chemistry 1037–1041 26. Eberson, L. & Nyberg, K. Studies on the Kolbe electrolytic synthesis. V. An
(Wiley, 2007). electrochemical analogue of the Ritter reaction. Acta Chem. Scand. 18,
8. Abraham S. et al. Aurora kinase compounds and methods of their use. 1567–1568 (1964).
International patent no. WO2011088045A1 (2011).
9. Kolbe, H. Beobachtungen über die oxydirende wirkung des sauerstoffs, wenn Publisher’s note: Springer Nature remains neutral with regard to jurisdictional
derselbe mit hülfe einer elektrischen säule entwickelt wird. J. Prakt. Chem. 41, claims in published maps and institutional affiliations.
137–139 (1847).
10. Hofer, H. & Moest, M. Mittheilung aus dem elektrochemischen Laboratorium
der Königl, Technischen Hochschule zu München. Justus Liebigs Ann. Chem.
323, 284–323 (1902). © The Author(s), under exclusive licence to Springer Nature Limited 2019

N A t U r e | www.nature.com/nature
RESEARCH Letter

Methods Acknowledgements Financial support for this work was provided by Pfizer,
Here we describe a typical procedure for the decarboxylative etherification. Further Inc., the National Science Foundation (CCI Phase 1 grant 1740656), and the
National Institutes of Health (grant number GM-118176). China Scholarship
experimental details are provided in the Supplementary Information. Council and Jilin University supported a fellowship to J.X., Zhejiang Yuanhong
General procedure for decarboxylative etherification. With no precautions to Medicine Technology Co. Ltd supported a fellowship to M.S., The Hewitt
exclude air or moisture, the ElectraSyn vial (5 ml) with a stir bar was charged Foundation supported a fellowship to Y.K., The Swedish Research Council
with carboxylic acid (0.2 mmol, 1.0 equiv.), alcohol (0.6 mmol, 3.0 equiv.), 2,4,6- supported a fellowship to H.L., Fulbright Fellowship supported a fellowship to
collidine (0.6 mmol, 3.0 equiv.), nBu4NPF6 (0.3 mmol, 1.5 equiv.), 3 Å molecular P.M., and S.H.R. acknowledges an NSF Graduate Research Fellowship Program
(no. 2017237151) and a Donald and Delia Baxter Fellowship. We thank
sieves (150 mg), AgPF6 (0.3 mmol, 1.5 equiv.) and CH2Cl2 (3.0 ml). The ElectraSyn
D.-H. Huang and L. Pasternack for assistance with NMR spectroscopy; J. Chen
vial cap equipped with anode (graphite) and cathode (graphite) were inserted into for measuring the high-resolution mass spectroscopy data, and A. Rheingold,
the mixture. After pre-stirring for 15 min, the reaction mixture was electrolysed C. E. Moore and M. A. Galella for X-ray crystallographic analysis.
at a constant current of 10 mA for 3 h. The ElectraSyn vial cap was removed, and
electrodes were rinsed with Et2O (2 ml), which was combined with the crude Author contributions J.X., M.S. and P.S.B. conceived the project. J.X., M.S., Y.K.,
mixture. Then, the crude mixture was further diluted with Et2O (30 ml). The H.L., D.G.B. and P.S.B. designed the experiments and analysed the data. J.X. and
M.S. developed the electrochemical decarboxylative methods and performed
resulting mixture was washed with 2 M HCl (20 ml) and NaHCO3 (aq.) (20 ml), their applications. H.L. and Y.K. carried out the mechanistic study. J.X., M.S.,
dried over Na2SO4 and concentrated in vacuo. The crude material was purified S.H.R., M.C., P.M., G.B., M.R.C., A.D., M.D.B., G.M.G., J.E.S., J.S. and S.Y. conducted
by preparative thin-layer chromatography to furnish the desired product. Full experiments to demonstrate the substrate scope. P.S.B. wrote the manuscript.
experimental details and characterization of new compounds can be found in J.X., M.S., Y.K., S.H.R., P.M., H.L. and D.G.B. assisted in writing and editing the
the Supplementary Information. manuscript.

Competing interests The authors declare no competing interests.


Data availability
The data that support the findings of this study are available within the paper and Additional information
its Supplementary Information. Metrical parameters for the structures of (2R)-77 Supplementary information is available for this paper at https://doi.org/
10.1038/s41586-019-1539-y.
and (11R)-138 are available free of charge from the Cambridge Crystallographic Correspondence and requests for materials should be addressed to P.S.B.
Data Centre (https://www.ccdc.cam.ac.uk/) under reference numbers 1918528 Reprints and permissions information is available at http://www.nature.com/
and 1903823, respectively. reprints.

Das könnte Ihnen auch gefallen