Science - USA (2019-08-30)

(Antfer) #1
a silane to regenerate the phos-
phane. The authors combined
this with the above-mentioned
catalytic system of Taniguchi
to realize a fully catalytic Mit-
sunobu reaction ( 9 ). However,
these findings were reevaluated
and critically discussed in the
community ( 10 , 11 ). Taniguchi
and co-workers performed a de-
tailed study on the fully catalytic
system and showed that the re-
action proceeds under retention
of configuration and that the
azo reagent does not participate
( 10 ). They concluded that the
fully catalytic Mitsunobu reac-
tion was not realized.
These previous findings un-
derline the tremendous chal-
lenges associated with the
development of a fully catalytic
system based on the original
Mitsunobu protocol. It has
proved difficult to realize a cat-
alytic system by implementing
a selective reducing agent to re-
generate the phosphane in the
presence of an oxidizing agent
to regain the azodicarboxylate.
Instead of trying to find a
perhaps unachievable balance
in reactivity and selectivity, Bed-
doe et al.’s catalyst circumvents
the above-mentioned challenges.
Their system provides an exciting
approach to the long-standing
challenge of the direct bimolecu-
lar substitution of nonactivated
chiral alcohols, delivers a new
platform in organocatalysis, and
offers new possibilities for sus-
tainable organic synthesis. j

SCIENCE sciencemag.org

GRAPHIC: C. BICKEL/


SC


IENCE


Beddoe et al. now achieve
this goal, by means of a re-
dox-neutral organocatalytic
Mitsunobu reaction. Most
notably, their protocol is
based on a single, easily ac-
cessible catalyst; no further
stoichiometric reductants
or oxidants are required.
Furthermore, the only by-
product is water, greatly
increasing the overall atom
economy of the reaction.
The structure of the cata-
lyst seems simple: a tertiary
phosphane oxide bearing
a phenolic group (see the
figure). On second glance,
however, it becomes clear
that the design is brilliant,
because it enables the dehy-
drative coupling of primary
and secondary alcohols with
various pronucleophiles. Ini-
tially, the catalyst reacts with
the pronucleophile to form a
cyclic oxyphosphonium salt
while releasing water. The
alcohol then reacts with the
phosphonium salt, form-
ing a new P–O bond and si-
multaneously liberating the
phenolic side chain of the
catalyst. This represents the
activated substrate, which
eventually reacts with the
nucleophile, forming the
substitution product under
liberation of the catalyst.
The authors show that this
reaction works for a broad
range of primary (R= H)
and secondary (R, R 9 Þ H)
alcohols. They converted
chiral alcohols under ste-
reochemical inversion to
the respective substituted
products with excellent
selectivity. As they dem-
onstrate, the methodology
allows the use of various
pronucleophiles, leading to
the formation of C–O, C–N, and C–S bonds.
Beddoe et al.’s approach differs conceptu-
ally from earlier attempts to realize catalytic
Mitsunobu reactions. Previous strategies
stayed close to the original procedure, in-
troducing additional reagents and catalysts
to regenerate the phosphane or azodicar-
boxylate and thereby realize protocols that
were catalytic in these components.
In 2006, But and Toy reported a variant
of the Mitsunobu reaction that was catalytic
in the azo reagent, using an iodine-based

oxidant to regenerate the azo reagent ( 7 ). In
2013, Taniguchi et al. ( 8 ) described further
developments in this direction, showing
that molecular oxygen from air could serve
as the terminal oxidant in the presence of
an iron catalyst. However, besides the in-
creased complexity of the overall systems,
both procedures still required overstoichio-
metric amounts of the phosphane.
Two years later, Buonomo and Aldrich re-
alized a protocol that was catalytic in phos-
phane by using stoichiometric amounts of

A dream reaction: Direct nucleophilic substitution of alcohols

A classic solution: The Mitsunobu reaction

Catalytic variant of Beddoe et al.

Mechanistic proposal

Reaction pathway

OH
+ Nu–H +

Nu–H
Nu–

R R 9

OH
Nu–H

Ph 3 P (≥1 equiv.)
CO 2 Et

EtO 2 C
(≥1 equiv.)

NN

+ + +
R R 9

OH
R R 9

OH
R R 9
O
R R 9

Nu
H 20

By-product

Reagents By-products

Alcohol Pronucleophile Product

R R 9

OH
+ Nu–H +
R R 9

Nu
H 20

Organocatalyst
R R 9

Ph 3 P+ P Ph 3

Ph 3 P=O

CO 2 Et

EtO 2 C

NN–

CO 2 Et
Ph 3 P=O
EtO 2 C

NN
H

Nu H
R R 9

Nu
R R 9
CO 2 Et
EtO 2 C

NN
H

H

+
+

Nu

H 20

Nu–

R R 9

OHO Nu–H
P
Ph

Ph

O
R R 9

Ph Ph

OH

P+
Ph O
Ph
P+
Nu–

Nu, nucleophile; R and/or R 9 , H, alkyl, aryl, etc.; Et, ethyl; Ph, phenyl.

REFERENCES AND NOTES


  1. R. H. Beddoe et al., Science 365 , 910
    (2019).

  2. K. C. K. Swamy, N. N. B. Kumar,
    E. Balaraman, K. V. P. P. Kumar, Chem.
    Rev. 109 , 2551 (2009).

  3. O. Mitsunobu, Synthesis 1981 , 1 (1981).

  4. M. Oyo, Y. Masaaki, M. Teruaki, Bull.
    Chem. Soc. Jpn. 40 , 935 (1967).

  5. B. M. Trost, Science 254 , 1471 (1991).

  6. S. Fletcher, Org. Chem. Front. 2 , 739 (2015).

  7. T. Y. S. But, P. H. Toy, J. Am. Chem. Soc. 128 , 9636 (2006).

  8. D. Hirose, T. Taniguchi, H. Ishibashi, Angew. Chem. Int. Ed.
    52 , 4613 (2013).

  9. J. A. Buonomo, C. C. Aldrich, Angew. Chem. Int. Ed. 54 ,
    13041 (2015).

  10. D. Hirose, M. Gazvoda, J. Košmrlj, T. Taniguchi, Org. Lett. 18 ,
    4036 (2016).

  11. R. H. Beddoe, H. F. Sneddon, R. M. Denton, Org. Biomol.
    Chem. 16 , 7774 (2018).


ACKNOWLEDGMENTS
We are grateful for support from the Leibniz ScienceCampus
Phosphorus Research Rostock (www.sciencecampus-rostock.de).
10.1126/science.aay6635

30 AUGUST 2019 • VOL 365 ISSUE 6456 867

Catalytic substitution of alcohols
Beddoe et al. report a redox-neutral organocatalytic analog of the stoichiometric
Mitsunobu reaction for substituting otherwise inert alcohols.

Published by AAAS
Free download pdf