You are on page 1of 12

Review Article

https://doi.org/10.1038/s44160-021-00002-3

Stereoselective synthesis through remote


functionalization
Itai Massad   , Rahul Suresh   , Lucas Segura    and Ilan Marek    ✉

Transition-metal-catalysed alkene isomerization is instrumental to remote functionalization processes, in which a chemical


transformation is induced at a position remote from the initial reactive site. The dynamic nature of alkene isomerization, which
is crucial for such transformations, often leads to substantial difficulties in controlling the stereochemistry of C(sp3) centres
along the carbon skeleton. This review features synthetic methods that tackle this issue and strategically leverage alkene isom-
erization to control C(sp3) stereocentres in complex organic molecules. Stereocentres can be created at either the initiation or
termination site of an isomerization process, retained during chain-walking across the molecular backbone or revealed through
restructuring of the carbon skeleton by selective ring-opening of a strained ring. As each of these options imposes different
mechanistic requirements, the different examples are divided according to the stage in the chain-walking process at which ste-
reochemistry is established and the type of intermediates that lead to stereodefined products.

T
ransition metal complexes have been known to promote the 2. Retention of stereochemistry during the chain-walking process.
positional isomerization of alkenes for over half of a century, 3. Formation of stereocentres at the termination site of alkene
a process also termed ‘chain-walking’. Although this simple isomerization (through the capture of alkyl- or allylmetal inter-
process was historically deemed a bothersome side reaction, con- mediates or by subsequent reactions of the isomerized alkene).
temporary synthetic methods that capitalize on controlled alkene
isomerization have provoked researchers to explore its synthetic Stereocentre formation at the initiation site
potential1–7. A unique feature of methods that employ alkene isom- C–H bond formation. Arguably, the simplest way to establish ste-
erization is the possibility to induce a chemical transformation at reochemistry at the initiation site of a remote functionalization pro-
a position isolated from the initial reactive site, a concept termed cess is by face-selective insertion of a hydride onto a trisubstituted
remote functionalization (Fig. 1a). Every remote functionalization alkene or an allylmetal intermediate.
process requires an ‘initiation’ moiety that steers and chemically The first notable feat in this field was the enantioselective isom-
interacts with the reagent or catalyst. As the ‘termination’ step occurs erization of allylamines to give enamines. After pioneering work
at a structurally distant location from the initiation site, a commu- on cobalt-catalysed isomerization9, a significant step forward was
nicative process must take place between these sites. Alternatively, achieved using rhodium catalysts. In 1984, Noyori reported the first
relocation of a π bond to a neighbouring position can also be use of the then newly developed BINAP ligand (BINAP, (S)- or (R)-
achieved in a controlled manner via monoisomerization, with com- 2,2′-bis(diphenylphosphino)-1,1′-binaphthalene), which enabled
plete control over the alkene stereochemistry in certain cases. the Rh-catalysed isomerization of allylamines to give their isomeric
The synthetic power of remote functionalization through alkene enamines with extremely high enantioselectivity and close to quan-
isomerization is convincingly demonstrated by the Ni-catalysed titative yields (Fig. 2a,b)10,11. The power of this catalytic system is
carboxylation of alkyl bromides reported by Martin and co-work- underlined by its use in the Takasago ton-scale route towards men-
ers, in which alkane-derived mixtures of regioisomeric alkyl bro- thol (10, Fig. 2a), in which enantioselective isomerization serves as
mides 2 are funnelled into valuable linear carboxylic acids, such the key step12.
as 3 (Fig. 1a)8. The isomerization was identified to proceed through a
Although the dynamic nature of alkene isomerization is at the 1,3-hydride shift at a relatively early stage11, with the enantioselec-
heart of the above transformation and its many congeners, it often tivity being dependent on the ligand and the alkene stereochemistry
leads to substantial difficulties in other contexts. Specifically, the of the starting material 5 (Fig. 2b). Therefore, a close correlation
configuration of neighbouring stereocentres is at constant peril between the initial E/Z ratio of the starting material and the enantio-
in the presence of an alkene isomerization catalyst capable of their meric excess of the resulting product 7 is to be expected. Extensive
epimerization. There is considerable incentive to overcome this chal- computational and experimental work helped to shed light on the
lenge and develop stereoselective remote functionalization processes, complete mechanism and specifically showed that the enantioselec-
which would form stereocentres at otherwise desolate positions in tivity is determined prior to the cleavage of the allylic C–H bond,
the molecule or even establish several stereocentres in a single step. during which the allylic hydrogen atom migrates to the Rh(I) as a
This Review showcases synthetic methods that control the ste- hydride with the assistance of the nitrogen lone pair13.
reochemistry of C(sp3) stereocentres through alkene isomerization. The enantioselective isomerization of allylic alcohols to give
These methods can be divided into three categories (Fig. 1b): aldehydes represents a simple and atom-economic route towards
β-stereogenic aldehydes, but has unfortunately proved more chal-
1. Formation of stereocentres at the initiation site of a chain-walk- lenging than the isomerization of the corresponding allylamines.
ing event (through migratory insertion onto alkenes). The first noteworthy example was reported by Tani in 1985 and

Schulich Faculty of Chemistry, Technion – Israel Institute of Technology, Haifa, Israel. ✉e-mail: chilanm@technion.ac.il

Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth 37


Review Article NATure SynThesis
a Alkene isomerization
(or chain-walking)
Reactive initiation
site M
M

Unreactive remote Functionalized


position product

Representative example

O
1 equiv. Br2 Br [Ni] catalyst
.
OH
MnO2 0
CO2 (1 atm), Mn
DMF, r.t.
1 2 3
Linear alkane Mixture of regioisomeric Valuable carboxylic acid
alkyl bromides 65% (from heptane)
No isolation required

b
At the At the
initiation site? termination site?

M
Formation of stereocentres

During the
chain-walking process?

Fig. 1 | Remote functionalization and general considerations. a, Remote functionalization processes that rely on alkene isomerization to access a
remote position through repeated isomerization (chain-walking). A representative example is shown, in which a readily available regioisomeric mixture
of alkyl bromides is funnelled into a single carboxylic acid. b, This review highlights remote functionalization processes that selectively establish C(sp3)
stereocentres in the product. Stereochemical information may be established at any position from the initiation to the termination site, each imposing
different requirements on the catalytic system. M, transition-metal-based reagent or catalyst; r.t., room temperature.

achieved a 53% enantiomeric excess at best11. Later, extensive opti- This approach is unfortunately restricted to specific substrates to a
mization of a planar chiral phosphaferrocene ligand by Fu and co- achieve high enantioselectivity27.
workers for Rh-based catalysts led to an 86% enantiomeric excess14.
The limitations of this system remain substantial, because efficient C–C and C–B bond formation. The enantioselective formation of
enantioinduction is limited to substrates that bear bulky substitu- a C–C or C–B bond as an initiation event for a chain-walking pro-
ents, and long reaction times at high temperatures are required15,16. cess leads to further increase in molecular complexity compared
In 2009, Mazet and collaborators reported an Ir-based cata- with that of the formation of a C–H bond. The Sigman group has
lyst for the highly enantioselective isomerization of trisubstituted developed an enantioselective redox-relay Heck reaction, in which
allylic alcohols, such as 11 (Fig. 2c)17,18. Several generations of a quaternary carbon stereocentre is formed at the initiation site by
ligand optimization culminated in a system that operates at room face-selective migratory insertion (Fig. 3a)28. The resulting alkyl-
temperature with excellent levels of enantioselectivity and a rela- palladium intermediate 22 then undergoes a series of β-hydride
tively general substrate scope, which represents significant progress elimination and reinsertion steps to reach the thermodynamic
in the field19. Further efforts were dedicated to the development sink of carbonyl formation. This transformation, which allows the
of Ru- (refs. 20–22), Pd- (ref. 23) and Rh- (refs. 24,25) based catalysts, highly enantioselective formation of remote quaternary carbon
which achieved high levels of enantioinduction for otherwise chal- stereocentres, is joined by analogous C–C, C–N and C–O bond-
lenging substrate classes, such as tetrasubstituted secondary allylic forming reactions which together wield considerable synthetic
alcohols (albeit with imperfect diastereoselectivity), and higher power29–43 and has even been applied in the total synthesis of com-
homologues, such as homoallylic alcohols. Overall, however, a gen- plex natural products44.
eral catalytic system for the enantioselective isomerization of allylic Another notable example that establishes a stereocentre at the
alcohols remains elusive. initiation site while allowing further increase of molecular com-
An alternative approach towards the enantioselective isom- plexity was reported by Hall and co-workers, who converted vinyl
erization of allylic alcohols was reported by Martín-Matute and nonaflates into enantioenriched allylboronate esters, such as 25, in
co-workers26, in which the base TBD (1,5,7-triazabicyclo-[4.4.0] the presence of a palladium catalyst and pinacolborane (Fig. 3b)45–47.
dec-5-ene) was used to enantiospecifically isomerize secondary Recent studies support a mechanism in which the initial genera-
allylic alcohols and ethers. The enantiospecificity achieved is close tion of a vinylboronate directs the [Pd]-H species to one enantio-
to perfect (Fig. 2d) and is proposed to originate from the short-lived topic face of the alkene, followed by alkene isomerization to set the
nature of the intimate ion pair 16 resulting from deprotonation, stereochemistry of the borylated centre48. The allylboronate reacts
which ensures protonation on the same enantiotopic face. The sub- with aldehydes in the same pot, affording diastereo- and enantioen-
stantial decrease in enantiospecificity in more polar solvents lends riched allylation product 26.
credence to this hypothesis.
An elegant strategy to control stereochemistry by isomerization Stereoretentive remote functionalization processes
relates to desymmetrization reactions (Fig. 2e), in which stereocen- At first glance, any chain-walking process that crosses a tertiary ste-
tres can be established at a remote position from the initiation site. reocentre should proceed with complete racemization, as an achiral

38 Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth


NATure SynThesis Review Article
a
NEt2 [Rh] H

LiNEt2 [Rh((S)-BINAP)] NEt 2

NEt2

4 E- 5 6 (R,E)-7
Myrcene

H3O

H2, Raney Ni ZnBr2

OH OH O

10 9 8
(–)-Menthol Isopulegol (R)-Citronellal

b c
Cy 12 (7.5 mol%) Cy

[Rh((R)-BINAP)] NEt2
Ph OH H 2 activation, Ph O
11 then degassed 14
NEt2 THF, 35 °C, 22 h
85%, 99/1 e.r.

Z-5 [Rh((S )-BINAP)] (R,E)-7 OMe

H
P N
NEt 2 NEt2 [Ir] Cy Ph
O H
N
[Rh((R)-BINAP)] Ir O
P H
t t
Bu Bu –
BArF
E -5 (S,E )-7
12 13

d e
NiBr2
N (–)-diop
Ph OH Ph O LiBHEt3
TBD [Ni]
N N O O O O O
Ph Et 2O O
F3 C Ph F3C H t
Bu
Toluene H H H
F3C Ph –70 °C
15 80 °C 17
18 h 144 h
Ph OH 82%, 95% e.s. 19
18 20
86%, 96/4 e.r.
16

Fig. 2 | Asymmetric C-H bond formation at the initiation site. a, Noyori’s application of asymmetric allylamine isomerization to the total synthesis of
menthol, still used today for ton-scale production. b, Rh/BINAP-catalysed isomerization of allylamines to give enantioenriched E-enamines. c, Asymmetric
Ir-catalysed mono-isomerization of primary allylic alcohols through a 1,2-hydride shift mechanism. d, Base-catalysed enantiospecific isomerization of
electron-poor allylic alcohols. e, Desymmetrization via isomerization opens the possibility to reveal an C(sp3) stereocentre at a distal position from the
isomerized double bond. coe, cyclooctene; DCE, 1,2-dichloroethane; BArF, tetrakis[3,5-bis(trifluoromethyl)phenyl]borate; diop, 2,3-O-isopropylidene-2,3-
dihydroxy-1,4-bis(diphenylphosphino)butane; d.r., diastereomeric ratio; e.r., enantiomeric ratio; e.s, enantiospecificity; TBD, 1,5,7-triazabicyclodec-5-ene.

trisubstituted alkene is intermediately formed (Fig. 4a). It was, how- species remains coordinated to one enantiotopic face of the alkene
ever, observed that the redox-relay Heck arylation of (R)-citronellol 28 (Fig. 4a), which re-establishes the stereocentre with the same
(27), shown in Fig. 4a, displays complete stereoretention of the configuration on migratory insertion. Conversely, if the catalyst dis-
remote tertiary stereocentre28. Importantly, the configuration of the engages from the substrate during chain-walking, the intermediate
tertiary centre in the product was independent of the enantiomer alkene leads to two enantiomers by coordination from either enan-
of the chiral ligand employed. A contrasting result was reported by tiotopic face of 30 (Fig. 4b), as was shown to be the case in Mazet and
Mazet and colleagues23,49, in which the Pd-catalysed isomerization co-workers’ study23. The factors at play in determining whether the
of the same substrate led to complete racemization of the tertiary process is dissociative or associative were explored by both groups.
centre (Fig. 4b). First and foremost, it appears that the more strongly donating
The difference between these two contrasting outcomes lies in ligands labilize the alkene ligand due to the trans-effect, in line with
the strength of the association between the catalyst and the alkene the experimental observations in which a bisphosphine ligand led to
formed during chain-walking. If the catalyst remains strongly bound racemization and a pyridine–oxazoline ligand did not. Additionally,
to the newly formed alkene, as in Prater and Sigman’s case42, the ste- the reaction solvent and its interaction with the metal counter-
reochemical information is retained owing to the fact that the Pd–H ion also plays a substantial role, with more strongly coordinating

Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth 39


Review Article NATure SynThesis

a
4-OMe-C 6H4B(OH) 2 (3 equiv.)
TBSO
TBSO Pd(CH 3CN)2 (OTs) 2 (6 mol%)
TBSO OH
(S)-5-CF 3-t Bu-PyrOx (9 mol%)
2
2 3 Å MS, Cu(OTf)2 (3 mol%) 2
OH O 2 balloon, DMF, r.t., 24 h Ar [Pd] O

F3C
MeO
21 22 23
O
N 61%, 93/7 e.r.
N
t Bu

(S)-5-CF3-tBu-PyrOx

b
HBpin (1.1 equiv.)
ONf Bpin
Pd(OAc)2 (3 mol%) p-MeC6H4CHO
(+)-Taniaphos (3.3 mol%) (1.1 equiv.)

PhNMe 2 (1.1 equiv.) Toluene (1 M), r.t., 16 h


N N N
Boc CPME (0.33 M), r.t., 16 h Boc Boc
OH
PPh2
24 25 26
PPh2 80%, 95.5/4.5 e.r.
Fe
Me2N

(+)-Taniaphos

Fig. 3 | Asymmetric C–C and C–B bond formation at the initiation site. a, The enantioselective redox-relay Heck reaction establishes remote quaternary
stereocentres with excellent stereocontrol. b, Asymmetric borylation–isomerization of alkenyl nonaflates leads to enantioenriched allylboronate esters,
which can be used in stereospecific carbonyl allylation. CPME, cyclopentyl methyl ether; DMF, dimethylformamide; MS, molecular sieves; OTs, tosylate;
OTf, triflate; ONf, nonaflate (perfluorobutanesulfonate); Taniaphos, 1-[(R)-(dimethylamino)[2-(diphenylphosphino)phenyl]methyl]-2-(diphenylphosphino)-
(2R)-ferrocene); TBS, tert-butyldimethylsilyl; (S)-5-CF3-tBu-PyrOx, (S)-4-(tert-butyl)-2-(5-(trifluoromethyl)pyridin-2-yl)-4,5-dihydrooxazole.

solvents leading to more associative chain-walking, but this effect Stereocentre formation at the termination site
has not been conclusively explained yet49. Other reports of remote Allyl- and benzylmetal intermediates. Chain-walking is often ter-
functionalization with varying degrees of enantioretention have minated by the formation of a stabilized organometallic intermedi-
accumulated over the years, and also correlate the degree of racemi- ate, which is subsequently trapped to regenerate the catalyst.
zation with the extent of dissociation during isomerization50–53. In an instructive illustration of this concept, Lu and co-workers
In the case of a quaternary carbon stereocentre along the molec- reported the Co-catalysed asymmetric remote hydroboration of
ular backbone, a scenario in which chain-walking directly crosses the unactivated internal alkene 40 using a newly developed chiral
the stereocentre is, of course, not viable. An alternative way to com- NNN-pincer ligand 41 (Fig. 5a)51. In the postulated mechanism,
plete a chain-walking process on a molecule that contains a qua- the chain-walking is thermodynamically driven by the forma-
ternary carbon would be to ‘skip’ this interfering motif by the ring tion of a benzylcobalt intermediate 42, which is trapped by HBpin
opening of a strained ring54. (4,4,5,5-tetramethyl-1,3,2-dioxaborolane) to afford the enantioen-
The first illustration of this concept was reported by our group riched product and regenerate the Co–H catalyst.
using stoichiometric amounts of the Zr-based Negishi reagent (Fig. Zhu and colleagues reported an asymmetric Ni-catalysed
4c)55,56. When cyclopropane 32, which features a quaternary carbon remote hydroalkynylation of allylbenzene derivative 44, which is
stereocentre and an ω-alkenyl chain, is exposed to the above zircon- proposed to proceed through a similar benzylnickel intermediate
ocene reagent, reversible alkene isomerization is initiated. As soon 45 (Fig. 5b)61. The alkyne functionality featured in the product
as the Zr-coordinated alkene reaches the position adjacent to the serves as a versatile synthetic handle for further transformations.
cyclopropane 33, the strained ring is sprung open to give zircona- This example of remote asymmetric C-C bond formation is, how-
cyclobutane 34, which can selectively react with two different elec- ever, still limited in scope.
trophiles. The quaternary carbon stereocentre previously contained Fang and co-workers reported a Ni-catalysed remote hydro-
in the strained ring is now part of the elaborated acyclic product 35. cyanation reaction of the non-conjugated diene 47 based on the
An updated demonstration of this concept was developed TADDOL-based bisphosphite ligand 48 (TADDOL, α,α,α′,α′-
using the Pd-catalysed Heck reaction of cyclopropane carbinol 36 tetraaryl-2,2-disubstituted 1,3-dioxolane-4,5-dimethanol), in
as a trigger for the chain-walking process (Fig. 4d)57. Again, this which the non-conjugated alkene selectively serves as the initiation
dynamic process is shifted by cleavage of the strained ring, after site (Fig. 5c)62. Chain-walking then ensues to result in the allylnickel
which the catalytic cycle is completed with the oxidation of the ter- intermediate 49, which undergoes reductive elimination to afford
minal alcohol group and release of the Pd catalyst. Notably, the ste- an enantioenriched allylic nitrile. The substrate scope is convinc-
reochemistry of the tertiary centre in the product 39 is well-defined ingly broad, and dienes that feature numerous combinations of
thanks to the non-dissociative chain-walking process58. The sub- alkenes reacted with high regio- and enantioselectivity.
strates and catalytic systems used in this reaction can be varied to As early as 1994, Larock et al. reported a beautiful three-com-
create an additional stereocentre at the initiation site and result in ponent process in which the Heck arylation of non-conjugated
stereodefined acyclic molecules that possess multiple quaternary dienes is followed by chain-walking to result in allylpalladium
and tertiary stereocentres59,60. intermediates, which are trapped with nucleophilic amines63.

40 Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth


NATure SynThesis Review Article
a
Retention of a preinstalled stereocentre during chain-walking

Pd(CH3CN2)(OTs)2 (10 mol%)


Ph-B(OH)2 (3 equiv.)
H-[Pd]
OH L (14 mol%) OH O
Ph Ph
3 Å MS, Cu(OTf)2 (4 mol%)
DMF, r.t., 24 h
27 29
(R)-citronellol L: (S)-5-CF3-tBu-PyrOx: 65%
28 >99/1 e.r.
>99/1 e.r.
Strong association between (R)-5-CF3-tBu-PyrOx: 62%
[Pd]-H and alkene leads to retention >99/1 e.r.
b
Racemization of a preinstalled stereocentre during chain-walking

[(dcpe)PdMeCl] (5 mol%) H-[Pd]

OH NaBArF (5.5 mol%) O


OH
Cyclohexene (50 mol%)

27 31
(R)-citronellol 38%
30 1/1 e.r.
>99/1 e.r.
Dissociation from the catalyst
leads to racemization
c
Zr-mediated alkene isomerization-cyclopropane ring-opening as a means to "walk over" quaternary stereocentres

1. Cp2ZrC4H8 (2 equiv.), Et2


Bu
2
3. Acetone OH Bu
32 4. CuI 35
5. Propargyl chloride 53%, 98/2 E/Z, 98/2 d.r.
6. H3O

[Zr]
[Zr] Bu
Bu
2

33 34
d
Pd-catalysed Heck arylation as a trigger for ring-opening

p-MeOC6H4Br (1.2 equiv.)


Bu Pd(OAc)2 (5 mol%), (p-CF3C6H4)3P (15 mol%) Bu
OH
NaHCO3 (2.5 equiv.), TBACl (2 equiv.)

O
MeO 39
36 44%, 98/2 d.r.

Bu Ar
OH [Pd] OH
Ar

[Pd] Bu

37 38

Fig. 4 | Walking ‘over’ tertiary and quaternary stereocentres. a, Chain-walking occurs without dissociation from the substrate, and thus ensures the
retention of stereochemical information. b, Dissociative chain-walking results in the loss of stereochemical information as the intermediate alkene can be
engaged from either enantiotopic face. c, Zr-mediated alkene isomerization–cyclopropane ring-opening. A cyclopropane ring that features a quaternary
stereocentre can allow to ‘walk over’ quaternary centres. The first demonstration of this strategy used Negishi’s zirconocene reagent. Alkene isomerization
followed by cleavage of the cyclopropane results in a bis-nucleophilic organozirconocene and leaves the stereodefined centre untouched. d, Alternatively,
chain-walking can be initiated by a Heck arylation, which culminates in ring opening and alcohol oxidation. This serves as a catalytic counterpart to the
previous transformation. DCE, 1,2-dichlorethane; dcpe, 1,2-bis(dicyclohexylphosphino)ethane; TBA, tetrabutylammonium; THF, tetrahydrofuran; Cp,
cyclopentadienyl.

Even though many examples of both enantioselective Pd-catalysed attack of the amine nucleophile onto the allylpalladium intermedi-
Heck arylation and allylic substitution reactions are known, an ate 54, finally realizing this long-standing goal (Fig. 5d)64.
enantioselective variant of this process remained elusive until
recently. In 2020, Zhou and colleagues reported that the use of a Alkylmetal intermediates. Instead of trapping a thermodynami-
bespoke chiral bisphosphite ligand 53 enables the enantioselective cally stabilized organometallic intermediate, selectivity can also

Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth 41


Review Article NATure SynThesis

a Co(OAc) 2 (2.5 mol% )


41 (3 mol%) [Co] Bpin
HBpin (1.2 equiv.) Ph
Ph Ph
Ph 3 Ph
Et2 O, r.t., 20 h

40 43
4.1/1Z/E 90%, 99/1 e.r.
42

b NiI2(xH 2O) (5 mol%)


(S)-5-CF3-t Bu-PyrOx (6 mol%) Et
(MeO)3SiH (2.5 equiv.) Br
MeO [Ni] TIPS MeO
Bromoalkyne (1.5 equiv.) MeO
K3PO4·4H2O (2.5 equiv) TIPS
AcO AcO
NaI (2.0 equiv) AcO
44 PhCF3, 0 °C, 12 h 46
45 78%, 95/5 e.r.
90:10 r.r.

c Ni(cod)2 (5 mol%)
48 (5 mol%) CN
[Ni]
Me 2C(OH)CN (3.0 equiv.)
Ph Ph Ph
toluene, 50 °C, 12 h
47 50
49 93%, 95/5 e.r.

d PdCl 2 (5 mol%)
[Pd]
H 53 (6 mol%) Ph
N proton sponge (2 equiv.)
+ NHR2
Ph N
PhI 55
ethylene glycol, 60 °C, 72 h 90%
51 52
54 97/3 e.r.

O
= (R )-BINOL, Ar = 3-CF3 -C6H4 t Bu
O O
O
Ar Ar O
N P
H O O P
N O
O O
N NPh
O O O
P t Bu
2
Ar Ar O

41 48 53

Fig. 5 | Formation of stereocentres at the termination site through allyl- and benzylmetal intermediates. a, In the Co-catalysed enantioselective
remote hydroboration of alkenes, the thermodynamically favoured benzylcobalt intermediate undergoes reductive elimination to afford enantioenriched
alkylboronate esters. b, Ni-catalysed isomerization–hydroalkynylation provides valuable enantioenriched propargylbenzene derivatives and proceeds
through a similar benzylnickel intermediate. c, Ni-catalysed asymmetric remote hydrocyanation of non-conjugated dienes. The terminal alkene selectively
reacts with the Ni–H catalyst, followed by chain-walking to result in a stabilized allylnickel species which furnishes an enantioenriched allylic nitrile.
d, A Pd-catalysed asymmetric three-component Heck arylation–isomerization–amination cascade. The enantiodetermining step is the trapping of the
allylpalladium intermediate with a nucleophilic amine. cod, cyclooctadiene; r.r., regioisomeric ratio; proton sponge, 1,8-bis(dimethylamino)naphthalene.

hinge on kinetic differentiation between the regioisomeric alkyl- isomerization-cycloisomerization process, affording synthetically
metal intermediates generated during chain-walking. valuable cyclopentanone derivatives66.
In an elegant example by Zhang and co-workers, an aldehyde In work reported by Nishimura and co-workers in 2017, an
C–H bond serves as a hydride source for a Rh catalyst (through Ir-based catalyst isomerizes ω-alkenyl ethers, for example, 60, into
C–H oxidative addition), which allowed cyclization onto a pendant the corresponding enol ethers, which then undergo an enantiose-
alkene65. The resulting alkylmetal intermediate 57 did not immedi- lective hydroarylation reaction catalysed by the same Ir complex
ately undergo reductive elimination, but first underwent a single- (Fig. 6b)67,68. The authors showed that the enantioselectivity is inde-
position isomerization that led to 58, which underwent reductive pendent of the stereochemical purity of the intermediate enol ether,
elimination to form a five-membered ring (Fig. 6a). Detailed com- and instead proposed that it arises from different rates of reductive
putational investigations by the authors suggest that rotation along elimination by the different diastereomers of alkyliridium interme-
a C–C bond prior to the β-hydride elimination required for isom- diate 61.
erization is the turnover-limiting and enantio-determining step65. A Kochi and collaborators explored an alternative strategy to shift
prior study from Dong and colleagues describes the desymmetri- the chain-walking equilibrium, by selective cyclization of an alkyl-
zation of α,α-diallyl aldehydes by an enantioselective Rh-catalysed metal intermediate 64 onto another alkene present in the molecule

42 Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth


NATure SynThesis Review Article
a (R )-DM-MeO-BIPHEP
(5 mol%)
[Rh(cod)Cl] 2
(2.5 mol%)
O O O
NaBArF (5 mol%) [Rh]
[Rh]
H 1,4-Dioxane H
O H
110 °C, 24 h H Ph
Ph Ph
56 57 58 59
95%, 70/30 d.r.
96/4 e.r.

b Ph
[Ir(cod)Cl] 2 (2.5 mol%)
(R )-BINAP* (5 mol%) MeO
Ph OMe NaBArF (10 mol%) Ph OMe N

2-Phenylpyridine, 80 °C
[Ir] Ar
Toluene, 24 h

60 61 62
87%, 97/3 e.r.

c
(phen)PdMeCl (2.5 mol%)
EtO2C NaBArF (3 mol%) EtO 2C H2 /PtO2 EtO 2C
EtO2C EtO 2C [Pd] EtO 2C
DCE, r.t., 24 h

63 64 65
1/4 E /Z 83%, 95/5 d.r.

d s BuLi(1.2 equiv.)
67 (1.2 equiv.)
H Et2O, –78 °C [Pd]Ph Ph
O O O
nPr
then, ZnCl2 (1.2 equiv.) nPr n
Pr
n N H then, PhBr n N H n N H
Pr Pr Pr
Boc Pd2(dba) 3 (2.5 mol%) Boc Boc
68 (5 mol%)
Toluene, 80 °C
66 69 70
72%, 96.5/3.5 e.r.

MeO PAr2 PAr2


H P(iPr)2
MeO PAr2 PAr2 N

N N
H

(R )-DM-MeO-BIPHEP (R )-BINAP* 67 68
Ar = 3,5-di-Me-C6H3 Ar = 3,5-di- t Bu-4-OMe-C6H2 (–)-Sparteine

Fig. 6 | Formation of stereocentres at the termination site through alkylmetal intermediates. a, Rh-catalysed asymmetric hydroacylation involving
alkene isomerization. The alkylmetal intermediate that leads to five-membered ring formation is selectively trapped. b, Ir-catalysed isomerization–
hydroarylation of a ω-alkenyl ether. Isomerization to an enol ether is followed by enantioselective hydroarylation. c, Pd-catalysed remote
cycloisomerization of 1,n-dienes, for example, 63. Hydropalladation is followed by reversible chain-walking, which is shifted by a 5-exo-trig cyclization. d,
Enantioselective C5 functionalization of Boc-1,3-oxazinanes through Pd-catalysed alkene isomerization and Negishi cross-coupling of an enantioenriched
organozinc species. The enantiospecificity of the isomerization process stems from its non-dissociative nature. dba, dibenzylideneacetone; BIPHEP,
2,2′-bis(diphenylphosphino)biphenyl; phen, phenanthroline; Boc, tert-butyloxycarbonyl; (R)-DM-MeO-BIPHEP, (R)-di-Me-MeO-BIPHEP.

to form a kinetically favoured five-membered ring (Fig. 6c)69,70. The palladium catalyst equipped with the phosphine ligand 68, known
process is terminated by a non-regioselective β-hydride elimina- to favour chain-walking, were added. After transmetallation to
tion, which mandates the hydrogenation of the cycloisomeriza- palladium, monoisomerization resulted in a new alkylpalladium
tion product to obtain isomerically pure products 65. Similar work species 69, which retained the stereogenic information due to the
regarding hydrosilylative cycloisomerizations was performed by the non-dissociative nature of the isomerization. Reductive elimination
same group71. afforded the final product 70, which can be easily cleaved under
A synthetically and mechanistically exciting study by Baudoin oxidative conditions to afford enantioenriched β-amino acids.
and co-workers demonstrates the preparation of functionalized Notably, isomerization can be avoided using a different phosphine
β-amino acid derivatives via a sequence of enantioselective lithia- ligand, which gives rise to a different regioisomer of the product.
tion, transmetallation, stereospecific alkene isomerization and The authors also demonstrated that enantioselective lithiation can
cross-coupling (Fig. 6d)72. Using (−)-sparteine (67) to induce enan- be performed using substoichiometric amounts of (+)-sparteine
tioselective lithiation α to the protected amine of 66, followed by surrogate along with the achiral diamine bispidine. Previous work
transmetallation with ZnCl2, an enantioenriched alkylzinc spe- from the same group demonstrated an example in which the stereo-
cies was obtained. Immediately following, an electrophile and a chemistry established by enantioselective lithiation can be retained

Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth 43


Review Article NATure SynThesis
a Ir-catalysed alkene isomerization-asymmetric carbonyl allylation

[Ir(cod)2]BF4 (5 mol%)
OH
PCy3 (12.5 mol%)
H2 activation PhCHO
BPin BPin
(R)-TRIP (5 mol%)
1,2-DCE, 4 Å MS
71 73
72
91%, 98/2 d.r.
98/2 e.r.
b syn-selective allylation through Z-selective Ni-catalysed alkene isomerization

NiCl2(dppp) (10 mol%)


Zn powder (20 mol%) OH
ZnI2 (20 mol %) BPin
PhCHO
Bpin
HPPh2 (5 mol %)
74
4 h to r.t. (o/n) 76
75
72%, 91/9 d.r.
c Ir-catalysed alkene isomerization followed by asymmetric aldol reaction

MesCu (5 mol%)
[Ir(cod)2(PPh2Me)2]PF6 (R)-DTBM-SEGPHOS
(0.5 mol%) (5 mol%) OH OH
O H2 activation O iPrOH (1 equiv.)

B B
O O O O
PhCHO;
77 then LiBH4 79
78
91%, 94/6 d.r.
98.5/1.5 e.r.
d Ir-catalysed alkene isomerization followed by Mukaiyama aldol reaction

[Ir(cod)Cl]2 (0.5 mol%) PhCH(OMe)2 OMe O


PCy3 (3 mol%) Sc(OTf)3
NaBArF (1.25 mol%) H
OSiMe2Ph OSiMe2Ph
H2 activation

80 syn-82
76%, 94/6 d.r.
E-81

OMe O

As above H
OSiMe2Ph
OSiMe2Ph
83 anti-82
Z-81 67%, >95/5 d.r.

Fig. 7 | Formation of stereocentres at the termination site through intermolecular reactions of alkene intermediates. a, Ir-catalysed alkene isomerization
equilibrates readily available vinylboronate esters with their E-allylboron counterparts. The latter are trapped by reaction with an aldehyde present in
situ, catalysed by the chiral phosphoric acid (R)-TRIP, which resulted in diastereo- and enantiomerically enriched allylation products. b, Ni-catalysed Z-
selective isomerization leads to syn-allylation products through a similar pathway. c, Ir-catalysed alkene isomerization affords E-boron enolates, which
are subsequently engaged in a Cu-catalysed asymmetric aldol reaction. d, Ir-catalysed alkene isomerization proceeds through allylmetal intermediates
and allows the stereoselective formation of fully substituted aldehyde-derived silyl enol ethers. The two stereoisomeric silyl enol ethers participate in a
stereospecific Mukaiyama aldol reaction in the same pot. dppp, 1,3-bis(diphenylphosphino)propane; Mes, mesityl; o/n, overnight.

during a haptotropic isomerization process to enable another exam- alcohols 73. Furthermore, the Ir catalyst is compatible with the chiral
ple of an enantioselective remote cross-coupling reaction73. phosphoric acid (R)-TRIP ((R)-3,3′-bis(2,4,6-triisopropylphenyl)-
1,1′-binaphthyl-2,2′-diyl hydrogenphosphate), which exerts enan-
Alkene intermediates. Rather than directly trapping the organo- tiocontrol over the allylation. The key E-allylboronates could also
metallic intermediates involved in chain-walking, one can generate be accessed through long-range isomerization of ω-ene alkylboro-
and subsequently trap stereodefined alkenes, which are challenging nates, using the same catalytic system. This concept was extensively
to generate otherwise. explored by the Murakami group, and resulted in a suite of isom-
This idea is exemplified by the work from the Murakami group, erization–allylation reactions, which provided allylated products that
in which vinylboronate esters, such as 71 (readily prepared through featured various substitution patterns and functionality75–78. Hilt and
hydroboration of terminal alkynes), were subjected to isomeriza- co-workers recently reported a complementary method, in which a
tion with an Ir-based catalyst in the presence of aldehydes74. The Ni catalyst isomerized a remote alkene (74) to stereoselectively gen-
unfavourable isomerization from vinyl- to allylboronate 72 is erate Z-allylboronate 75, which furnished syn-homoallylic alcohol 76
driven by the reaction of the latter with the aldehyde (Fig. 7a). As (Fig. 7b)79. This study serves as an impressive synthetic application of
the reaction proceeds through a cyclic chair-like transition state, the the Z-selective isomerization catalysts developed by the Hilt group80.
E-stereochemistry of the allylboronate 72 generated in situ is trans- In a similar vein, Kanai and collaborators developed an ele-
lated into complete diastereocontrol over the product homoallylic gant stereoselective strategy towards the cross-aldol reaction of

44 Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth


NATure SynThesis Review Article
a
[Ir(coe)2Cl] 2 (1 mol%)
PCy3 (6 mol%), NaBPh 4 (2 mol%) O
O DCE, 75 °C, 60 h O
Bu
Bu Bu

84 85 86
75%, 95/5 d.r.

b O
[Ir(cod)Cl] 2 (1 mol%) PPh 3 (6 mol%)
O PCy3 (6 mol%), NaBArF (2 mol%) O 85 °C, 3 h O

O Ph DCE, r.t, 30 min O Ph or MgBr2 (80 mol%) Ph


Et2O, TBM, 10 min
87 89
88 85 °C: 87%,
2/1 E/Z, 96/4 d.r.
MgBr2: 86%
95/5 E/Z, 89/11 d.r.

c O O O

MeO [Ir(cod)Cl] 2 (1 mol%) MeO


[3,3] Me O
PCy3 (6 mol%), NaBArF (2.5 mol%)
9
DCE, TBM to 85 °C
Ph Ph 9
Ph 9

90 91 92
52%, >95/5 d.r.

d O Ph O Ph
MeO O
[Ir(cod)Cl] 2 (2 mol%) O
PCy3 (12 mol%), NaBArF (5 mol%) Hydrolysis
H
OMe Ph
PhCl, TBM to 140 °C
OMe O
O O
93 94 95
91%, >95/5 d.r. 85%, >95/5 d.r.

Fig. 8 | Accessing sigmatropic rearrangement substrates through alkene isomerization. a, A cationic Ir-based catalyst site- and stereoselectively
isomerizes the primary allyl fragment of bis-allylic ethers, which constitutes a powerful approach towards the aliphatic Claisen rearrangement to
generate quaternary carbon stereocentres. b, Diallyl acetals derived from α,β-unsaturated aldehydes undergo isomerization to the corresponding divinyl
acetals, which serve as substrates for the Coates–Claisen rearrangement. c, In an alkene isomerization–Cope rearrangement, the same Ir-based catalytic
system forms reactive divinyl cyclopropanes which rearrange to yield valuable cycloheptadiene products with complete diastereoselectivity. d, Di-ω-
ene epoxides similarly undergo isomerization and Cope rearrangement that leads to dihydrooxepines, which serve as a source of stereodefined acyclic
1,6-dicarbonyl compounds.

aldehydes81,82. By generating E-configured boron enolate 78 through aldehyde-derived silyl enol ethers 81, which underwent a stereo-
Ir-catalysed alkene isomerization, problems of stereoselectivity and specific Sc-catalysed Mukaiyama aldol reaction with benzaldehyde
self-condensation typical of aldehyde-derived enolates were side- dimethyl acetal in the same pot87, which resulted in two different
stepped. These nascent enolates were reacted with aldehydes in diastereomers of aldehyde 82 that bears adjacent quaternary and
the presence of a Cu-based chiral catalyst, which resulted in enan- tertiary stereocentres.
tio- and diastereopure aldol products, which were isolated as the [3,3]-Sigmatropic rearrangements, such as those named after
respective 1,3-diols 79 (Fig. 7c). As the aldehyde product of the first Claisen and Cope, are strikingly general transformations, as the
aldol reaction is similarly reactive to the initial aldehyde, the authors parent reactive scaffold can be decorated with various substitution
found that by variation of additives, double, triple and even qua- patterns and functionality without compromising the success of the
druple aldol products could be obtained with high stereoselectivity. sigmatropic event. Indeed, the main ‘bottleneck’ when such sigma-
Alkene isomerization is an attractive strategy to generate enolates tropic rearrangements are employed in synthesis is often to find an
because it can circumvent the issues of regio- and stereoselectivity efficient and selective access to the substrates, as many pose signifi-
often encountered with classical deprotonation-based approaches. cant synthetic challenges that result from the positioning of unsatu-
Building on previously reported Ir-based catalytic systems that ration in the molecule. It is therefore fitting that substantial efforts
operate through allylmetal intermediates83–85, our group developed have been dedicated to access difficult sigmatropic rearrangement
an approach towards fully substituted aldehyde-derived silyl enol substrates through alkene isomerization of their more readily avail-
ethers86. By virtue of the 1,3-hydride shift mechanism and the con- able regioisomeric counterparts.
formational preferences of the allyliridium hydride intermediate, In the case of the Claisen rearrangement, alkene isomerization
either stereoisomer of a given enolate was cleanly generated, with can be leveraged to access relatively tricky allyl vinyl ether substrates
the stereochemistry determined by the regiochemistry of the start- from simple diallyl ethers. This idea has been explored in numerous
ing material, rather than by the steric properties of the substituents iterations that utilized various catalytic systems, but the most syn-
at the α-carbon (Fig. 7d). Thus, two regioisomeric allyl silyl ethers thetically relevant example was reported by Nelson, who employed
80 and 83 were isomerized into two stereoisomeric fully substituted a highly site- and stereoselective Ir-based isomerization catalyst to

Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth 45


Review Article NATure SynThesis

isomerize substituted bis-allyl ethers, such as 84, into their reactive regiochemistry as well as stereochemistry. Finally, the development
allyl vinyl ether counterparts (85), which then rearranged in the of catalytic isomerization systems based on Earth-abundant tran-
same pot to selectively and efficiently furnish diastereopure Claisen sition metals, such as Mn, Fe, Co and Ni, without compromising
rearrangement products that possess quaternary carbon stereo- efficiency and selectivity is a crucial long-term goal.
centres (86) (Fig. 8a)88–91. The strategic use of alkene isomerization One can envisage exciting prospects for stereoselective remote
thus greatly enhances the synthetic appeal of the aliphatic Claisen functionalization reactions. We hope this overview of the topic will
rearrangement. inform and inspire further progress in the field.
The same concept could be demonstrated to an even greater
effect when diallyl acetals of α,β-unsaturated aldehydes, such as 87, Received: 29 June 2021; Accepted: 19 October 2021;
were employed, which yielded the corresponding divinyl acetals Published online: 12 January 2022
(88) (Fig. 8b)92. The allylic divinyl acetals thus obtained serve as
substrates for the Coates–Claisen rearrangement93–95, a relatively References
little-known variant that has suffered from limited applicabil- 1. Vasseur, A., Bruffaerts, J. & Marek, I. Remote functionalization through
ity owing to the inaccessibility of the key vinyl acetals. The acetal alkene isomerization. Nat. Chem. 8, 209–219 (2016).
2. Sommer, H., Juliá-Hernández, F., Martin, R. & Marek, I. Walking metals for
moiety confers two attractive features to this rearrangement: con- remote functionalization. ACS Cent. Sci. 4, 153–165 (2018).
siderable rate enhancement due to the electron-donating alkoxy 3. Molloy, J. J., Morack, T. & Gilmour, R. Positional and geometrical
moiety, which also enables the mediation of this rearrangement isomerisation of alkenes: the pinnacle of atom economy. Angew. Chem. Int.
by mild Lewis acids, such as MgBr2, as well as the presence of a Ed. 58, 13654–13664 (2019).
vinyl ether in the product 89, which serves as a protected aldehyde 4. Janssen‐Müller, D., Sahoo, B., Sun, S. & Martin, R. Tackling remote sp3 C−H
functionalization via Ni‐catalyzed ‘chain‐walking’ reactions. Isr. J. Chem. 60,
and offers synthetic possibilities distinct from those available with 195–206 (2020).
classical alkenes generated through other variants of the Claisen 5. Massad, I. & Marek, I. Alkene isomerization through allylmetals as a strategic
rearrangement. tool in stereoselective synthesis. ACS Catal. 10, 5793–5804 (2020).
Ir-catalysed alkene isomerization can also be applied to tran- 6. Fiorito, D., Scaringi, S. & Mazet, C. Transition metal-catalyzed alkene
siently generate stereodefined divinyl cyclopropane intermediates, isomerization as an enabling technology in tandem, sequential and domino
processes. Chem. Soc. Rev. 50, 1391–1406 (2021).
such as 91, which immediately undergo a Cope rearrangement 7. Zhang, F. G. et al. Remote fluorination and fluoroalkyl(thiol)ation reactions.
to yield cycloheptadienes that contain two contiguous stereocen- Chem. Eur. J. 26, 15378–15396 (2020).
tres (92) (Fig. 8c)96. Unlike the case of the isomerization–Claisen 8. Juliá-Hernández, F., Moragas, T., Cornella, J. & Martin, R. Remote
rearrangement, which is performed sequentially, the isomeriza- carboxylation of halogenated aliphatic hydrocarbons with carbon dioxide.
Nature 545, 84–88 (2017).
tion–Cope cascade effectively funnels the equilibrium towards the
9. Kumobayashi, H., Akutagawa, S. & Otsuka, S. Metal-assisted terpenoid
divinyl cyclopropane through the ensuing rearrangement, even syntheses. 6. Enantioselective hydrogen migration in prochiral allylamine
when ten different energetically degenerate isomers are present in systems by chiral cobalt catalysts. J. Am. Chem. Soc. 100, 3949–3950 (1978).
equilibrium through isomerization. 10. Tani, K. et al. Highly enantioselective isomerization of prochiral allylamines
A similar transformation can be performed on epoxides, such catalyzed by chiral diphosphine rhodium(I) complexes. Preparation of
optically active enamines. J. Am. Chem. Soc. 106, 5208–5217 (1984).
as 93, rather than cyclopropanes, to form reactive dialkenyl epox- 11. Tani, K. Asymmetric isomerization of allylic compounds and the mechanism.
ides, which rearrange into stereodefined dihydrooxepines (94) (Fig. Pure Appl. Chem. 57, 1845–1854 (1985).
8d)97. The rearrangement products are amenable to hydrolysis, and 12. Akutagawa, S. Asymmetric synthesis by metal BINAP catalysts. Appl. Catal. A
yield diastereomerically pure 1,6-dicarbonyl compounds that fea- 128, 171–207 (1995).
ture two adjacent stereocentres. This combined isomerization–rear- 13. Yoshimura, T. et al. Exploring the full catalytic cycle of rhodium(I)–BINAP-
catalysed isomerisation of allylic amines: a graph theory approach for path
rangement–hydrolysis sequence constitutes an attractive strategy optimisation. Chem. Sci. 8, 4475–4488 (2017).
towards acyclic 1,6-dicarbonyl compounds 95. 14. Tanaka, K., Qiao, S., Tobisu, M., Lo, M. M.-C. & Fu, G. C. Enantioselective
isomerization of allylic alcohols catalyzed by a rhodium/phosphaferrocene
Outlook complex. J. Am. Chem. Soc. 122, 9870–9871 (2000).
Work in the area of remote functionalization through alkene isom- 15. Tanaka, K. & Fu, G. C. A versatile new catalyst for the enantioselective
isomerization of allylic alcohols to aldehydes: scope and mechanistic studies.
erization is diverse and spans investigations into the mechanistic J. Org. Chem. 66, 8177–8186 (2001).
intricacies of chain-walking, the design of new catalytic systems 16. Fu, G. C. in Modern Rhodium(I)‐Catalyzed Organic Reactions (ed. Evans, P.
and the application of existing technologies in complex molecular A.) Ch. 4 (Wiley‐VCH, 2005).
settings. The development of new remote functionalization systems 17. Mantilli, L., Gerard, D., Torche, S., Besnard, C. & Mazet, C. Iridium‐catalyzed
can lead to their extensive use in industry, as illustrated by Takasago asymmetric isomerization of primary allylic alcohols. Angew. Chem. Int. Ed.
48, 5143–5147 (2009).
industry’s menthol synthesis. 18. Mantilli, L., Gérard, D., Torche, S., Besnard, C. & Mazet, C. Improved
The goal of this review was to highlight an additional facet and catalysts for the iridium‐catalyzed asymmetric isomerization of
source of opportunity in remote functionalization reactions, namely, primary allylic alcohols based on Charton analysis. Chem. Eur. J. 16,
the control and leverage of stereochemical information. Tackling 12736–12745 (2010).
issues of stereocontrol at all stages of the remote functionalization 19. Li, H. & Mazet, C. Iridium-catalyzed selective isomerization of primary allylic
alcohols. Acc. Chem. Res. 49, 1232–1241 (2016).
process can lead to synthetically and conceptually valuable results, 20. Cahard, D., Gaillard, S. & Renaud, J.-L. Asymmetric isomerization of allylic
as well as deepen our mechanistic understanding of such processes. alcohols. Tetrahedron Lett. 56, 6159–6169 (2015).
Specifically, designing isomerization catalysts that can form 21. Arai, N., Sato, K., Azuma, K. & Ohkuma, T. Enantioselective isomerization of
highly substituted alkenes with complete stereoselectivity can lead primary allylic alcohols into chiral aldehydes with the tol-binap/dbapen/
ruthenium(II) catalyst. Angew. Chem. Int. Ed. 52, 7500–7504 (2013).
to control over the diastereoselectivity of subsequent stereospecific
22. Wu, R., Beauchamps, M. G., Laquidara, J. M. & Sowa, J. R. Ruthenium-
transformations, such as aldol and allylation reactions, sigmatropic catalyzed asymmetric transfer hydrogenation of allylic alcohols by an
rearrangements and more. Alternatively, the migrating metal cata- enantioselective isomerization/transfer hydrogenation mechanism. Angew.
lyst can actively participate in stereocentre-forming steps, such as Chem. Int. Ed. 51, 2106–2110 (2012).
migratory insertion and reductive elimination. Identifying chiral 23. Lin, L., Romano, C. & Mazet, C. Palladium-catalyzed long-range
deconjugative isomerization of highly substituted α,β-unsaturated carbonyl
ligands that enable long-distance isomerization and also achieve effi- compounds. J. Am. Chem. Soc. 138, 10344–10350 (2016).
cient stereocontrol is thus an important undertaking. Furthermore, 24. Liu, T. L., Ng, T. W. & Zhao, Y. Rhodium-catalyzed enantioselective
modulating the catalyst to precisely control the position at which isomerization of secondary allylic alcohols. J. Am. Chem. Soc. 139,
termination occurs can result in products that feature challenging 3643–3646 (2017).

46 Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth


NATure SynThesis Review Article
25. Huang, R.-Z., Lau, K. K., Li, Z., Liu, T.-L. & Zhao, Y. Rhodium-catalyzed 52. Kochi, T., Kanno, S. & Kakiuchi, F. Nondissociative chain walking as a
enantioconvergent isomerization of homoallylic and bishomoallylic secondary strategy in catalytic organic synthesis. Tetrahedron Lett. 60, 150938 (2019).
alcohols. J. Am. Chem. Soc. 140, 14647–14654 (2018). 53. Li, J., Qu, S. & Zhao, W. Rhodium-catalyzed remote C(sp3)−H borylation of
26. Martinez-Erro, S. et al. Base-catalyzed stereospecific isomerization of silyl enol ethers. Angew. Chem. Int. Ed. 59, 2360–2364 (2020).
electron-deficient allylic alcohols and ethers through ion-pairing. J. Am. 54. Cohen, Y., Cohen, A. & Marek, I. Creating stereocenters within acyclic systems
Chem. Soc. 138, 13408–13414 (2016). by C–C bond cleavage of cyclopropanes. Chem. Rev. 121, 140–161 (2020).
27. Frauenrath, H., Reim, S. & Wiesner, A. Asymmetric double-bond 55. Masarwa, A. et al. Merging allylic carbon–hydrogen and selective carbon–
isomerization of cyclic allyl acetals by using diop and chiraphos modified carbon bond activation. Nature 505, 199–203 (2014).
nickel complexes. Tetrahedron Asymmetry 9, 1103–1106 (1998). 56. Bruffaerts, J. et al. Zirconocene-mediated selective C–C bond cleavage
28. Mei, T. S., Patel, H. H. & Sigman, M. S. Enantioselective construction of of strained carbocycles: scope and mechanism. J. Org. Chem. 83,
remote quaternary stereocentres. Nature 508, 340–344 (2014). 3497–3515 (2018).
29. Werner, E. W., Mei, T. S., Burckle, A. J. & Sigman, M. S. Enantioselective 57. Singh, S., Bruffaerts, J., Vasseur, A. & Marek, I. A unique Pd-catalysed
Heck arylations of acyclic alkenyl alcohols using a redox-relay strategy. Heck arylation as a remote trigger for cyclopropane selective ring-opening.
Science 338, 1455–1458 (2012). Nat. Commun. 8, 14200 (2017).
30. Mei, T.-S., Werner, E. W., Burckle, A. J. & Sigman, M. S. Enantioselective 58. Kou, X., Shao, Q., Ye, C., Yang, G. & Zhang, W. Asymmetric aza-Wacker-type
redox-relay oxidative Heck arylations of acyclic alkenyl alcohols using cyclization of N-Ts hydrazine-tethered tetrasubstituted olefins: synthesis of
boronic acids. J. Am. Chem. Soc. 135, 6830–6833 (2013). pyrazolines bearing one quaternary or two vicinal stereocenters. J. Am. Chem.
31. Xu, L. et al. Mechanism, reactivity, and selectivity in palladium-catalyzed Soc. 140, 7587–7597 (2018).
redox-relay Heck arylations of alkenyl alcohols. J. Am. Chem. Soc. 136, 59. Bruffaerts, J., Pierrot, D. & Marek, I. Efficient and stereodivergent synthesis
1960–1967 (2014). of unsaturated acyclic fragments bearing contiguous stereogenic elements.
32. Zhang, C., Santiago, C. B., Crawford, J. M. & Sigman, M. S. Enantioselective Nat. Chem. 10, 1164–1170 (2018).
dehydrogenative Heck arylations of trisubstituted alkenes with indoles to 60. Cohen, A., Chagneau, J. & Marek, I. Stereoselective preparation of distant
construct quaternary stereocenters. J. Am. Chem. Soc. 137, 15668–15671 (2015). stereocenters (1,5) within acyclic molecules. ACS Catal. 10, 7154–7161 (2020).
33. Chen, Z. M., Hilton, M. J. & Sigman, M. S. Palladium-catalyzed 61. Jiang, X. et al. Nickel-catalysed migratory hydroalkynylation and
enantioselective redox-relay Heck arylation of 1,1-disubstituted homoallylic enantioselective hydroalkynylation of olefins with bromoalkynes. Nat.
alcohols. J. Am. Chem. Soc. 138, 11461–11464 (2016). Commun. 12, 3972 (2021).
34. Race, N. J., Schwalm, C. S., Nakamuro, T. & Sigman, M. S. Palladium- 62. Yu, R., Rajasekar, S. & Fang, X. Enantioselective nickel‐catalyzed migratory
catalyzed enantioselective intermolecular coupling of phenols and allylic hydrocyanation of nonconjugated dienes. Angew. Chem. Int. Ed. 59,
alcohols. J. Am. Chem. Soc. 138, 15881–15884 (2016). 21436–21441 (2020).
35. Avila, C. M. et al. Enantioselective Heck–Matsuda arylations through chiral 63. Larock, R. C., Wang, Y., Lu, Y. & Russell, C. A. Synthesis of aryl-substituted
anion phase‐transfer of aryl diazonium salts. Angew. Chem. Int. Ed. 56, allylic amines via palladium-catalyzed coupling of aryl iodides, nonconjugated
5806–5811 (2017). dienes, and amines. J. Org. Chem. 59, 8107–8114 (1994).
36. Race, N. J., Yuan, Q. & Sigman, M. S. Enantioselective C2-alkylation of 64. Zhu, D. et al. Asymmetric three-component Heck arylation/amination of
indoles via a redox-relay Heck reaction of 2-indole triflates. Chem. Eur. J. 25, nonconjugated cyclodienes. Angew. Chem. Int. Ed. 59, 5341–5345 (2020).
512–515 (2019). 65. Liu, C., Yuan, J., Zhang, Z., Gridnev, I. D. & Zhang, W. Asymmetric
37. Chen, Z. M., Nervig, C. S., DeLuca, R. J. & Sigman, M. S. Palladium- hydroacylation involving alkene isomerization for the construction of
catalyzed enantioselective redox-relay Heck alkynylation of alkenols to access C3-chirogenic center. Angew. Chem. Int. Ed. 60, 8997–9002 (2021).
propargylic stereocenters. Angew. Chem. Int. Ed. 56, 6651–6654 (2017). 66. Park, J. W., Kou, K. G. M., Kim, D. K. & Dong, V. M. Rh-catalyzed
38. Yuan, Q. & Sigman, M. S. Palladium-catalyzed enantioselective relay Heck desymmetrization of α-quaternary centers by isomerization–hydroacylation.
arylation of enelactams: Accessing α, β-unsaturated δ-lactams. J. Am. Chem. Chem. Sci. 6, 4479–4483 (2015).
Soc. 140, 6527–6530 (2018). 67. Ebe, Y., Onoda, M., Nishimura, T. & Yorimitsu, H. Iridium-catalyzed
39. Yuan, Q. & Sigman, M. S. Palladium-catalyzed enantioselective regio- and enantioselective hydroarylation of alkenyl ethers by olefin
alkenylation of enelactams using a relay Heck strategy. Chem. Eur. J. 25, isomerization. Angew. Chem. Int. Ed. 56, 5607–5611 (2017).
10823–10827 (2019). 68. Ebe, Y. & Nishimura, T. Iridium-catalyzed branch-selective hydroarylation
40. Liu, J., Yuan, Q., Toste, F. D. & Sigman, M. S. Enantioselective construction of of vinyl ethers via C–H bond activation. J. Am. Chem. Soc. 137,
remote tertiary carbon–fluorine bonds. Nat. Chem. 11, 710–715 (2019). 5899–5902 (2015).
41. Ross, S. P., Rahman, A. A. & Sigman, M. S. Development and mechanistic 69. Kochi, T., Hamasaki, T., Aoyama, Y., Kawasaki, J. & Kakiuchi, F. Chain-
interrogation of interrupted chain-walking in the enantioselective relay Heck walking strategy for organic synthesis: catalytic cycloisomerization of
reaction. J. Am. Chem. Soc. 142, 10516–10525 (2020). 1,n-dienes. J. Am. Chem. Soc. 134, 16544–16547 (2012).
42. Prater, M. B. & Sigman, M. S. Enantioselective synthesis of alkyl allyl ethers 70. Hamasaki, T., Aoyama, Y., Kawasaki, J., Kakiuchi, F. & Kochi, T. Chain
via palladium‐catalyzed redox‐relay Heck alkenylation of O‐alkyl enol ethers. walking as a strategy for carbon–carbon bond formation at unreactive sites in
Isr. J. Chem. 60, 452–460 (2020). organic synthesis: catalytic cycloisomerization of various 1, n-dienes. J. Am.
43. Zhang, C. et al. Palladium-catalyzed enantioselective Heck alkenylation of Chem. Soc. 137, 16163–16171 (2015).
trisubstituted allylic alkenols: a redox-relay strategy to construct vicinal 71. Kochi, T., Ichinose, K., Shigekane, M., Hamasaki, T. & Kakiuchi, F.
stereocenters. Chem. Sci. 8, 2277–2282 (2017). Metal-catalyzed sequential formation of distant bonds in organic molecules:
44. Nakamura, H., Yasui, K., Kanda, Y. & Baran, P. S. 11-step total synthesis of palladium-catalyzed hydrosilylation/cyclization of 1,n-dienes by chain
teleocidins B-1–B-4. J. Am. Chem. Soc. 141, 1494–1497 (2019). walking. Angew. Chem. Int. Ed. 58, 5261–5265 (2019).
45. Lessard, S., Peng, F. & Hall, D. G. α-Hydroxyalkyl heterocycles via chiral 72. Lin, W., Zhang, K.-F. & Baudoin, O. Regiodivergent enantioselective C–H
allylic boronates: Pd-catalyzed borylation leading to a formal enantioselective functionalization of Boc-1, 3-oxazinanes for the synthesis of β2- and β3-amino
isomerization of allylic ether and amine. J. Am. Chem. Soc. 131, 9612–9613 acids. Nat. Catal. 2, 882–888 (2019).
(2009). 73. Royal, T. & Baudoin, O. Pd-catalyzed γ-arylation of γ,δ-unsaturated
46. Kim, Y. R. & Hall, D. G. Optimization and multigram scalability of a catalytic O-carbamates via an unusual haptotropic rearrangement. Chem. Sci. 10,
enantioselective borylative migration for the synthesis of functionalized chiral 2331–2335 (2019).
piperidines. Org. Biomol. Chem. 14, 4739–4748 (2016). 74. Miura, T., Nishida, Y., Morimoto, M. & Murakami, M. Enantioselective
47. Clement, H. A. & Hall, D. G. Synthesis of α-hydroxyalkyl dehydroazepanes synthesis of anti homoallylic alcohols from terminal alkynes and aldehydes
via catalytic enantioselective borylative migration of an enol nonaflate. based on concomitant use of a cationic iridium complex and a chiral
Tetrahedron Lett. 59, 4334–4339 (2018). phosphoric acid. J. Am. Chem. Soc. 135, 11497–11500 (2013).
48. Clement, H. A., Estaitie, M., Kim, Y.-R., Hall, D. G. & Legault, C. Y. 75. Miura, T., Oku, N. & Murakami, M. Diastereo- and enantioselective synthesis
Mechanism of the palladium-catalyzed asymmetric borylative migration of of (E)-δ-boryl-substituted anti-homoallylic alcohols in two steps from
enol perfluorosulfonates: insights into an enantiofacial-selective terminal alkynes. Angew. Chem. Int. Ed. 58, 14620–14624 (2019).
transmetalation. ACS Catal. 11, 8902–8914 (2021). 76. Miura, T., Nishida, Y. & Murakami, M. Construction of homoallylic alcohols
49. Larionov, E., Lin, L., Guenee, L. & Mazet, C. Scope and mechanism in from terminal alkynes and aldehydes with installation of syn-stereochemistry.
palladium-catalyzed isomerizations of highly substituted allylic, homoallylic, J. Am. Chem. Soc. 136, 6223–6226 (2014).
and alkenyl alcohols. J. Am. Chem. Soc. 136, 16882–16894 (2014). 77. Miura, T. et al. Enantioselective synthesis of anti-1,2-oxaborinan-3-enes from
50. Ida, A., Hoshiya, N. & Uenishi, J. Alkene migration to the end-terminal aldehydes and 1,1-di(boryl)alk-3-enes using ruthenium and chiral phosphoric
carbon bearing a phenyl group over a chiral siloxy carbon center in Heck acid catalysts. J. Am. Chem. Soc. 139, 10903–10908 (2017).
reaction. Tetrahedron 71, 6442–6448 (2015). 78. Miura, T., Nakahashi, J. & Murakami, M. Enantioselective synthesis of
51. Chen, X., Cheng, Z., Guo, J. & Lu, Z. Asymmetric remote CH borylation of (E)‐δ‐boryl‐substituted anti homoallylic alcohols using palladium and a chiral
internal alkenes via alkene isomerization. Nat. Commun. 9, 3939 (2018). phosphoric acid. Angew. Chem. Int. Ed. 56, 6989–6993 (2017).

Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth 47


Review Article NATure SynThesis
79. Weber, F., Ballmann, M., Kohlmeyer, C. & Hilt, G. Nickel-catalyzed double 92. Massad, I. & Marek, I. Alkene isomerization revitalizes the Coates–Claisen
bond transposition of alkenyl boronates for in situ syn-selective allylboration rearrangement. Angew. Chem. Int. Ed. 60, 18509–18513 (2021).
reactions. Org. Lett. 18, 548–551 (2016). 93. Coates, R. M., Shah, S. K. & Mason, R. W. Stereoselective total synthesis of
80. Weber, F. et al. Double-bond isomerization: highly reactive nickel catalyst (+/–)-gymnomitrol via reduction–alkylation of α-cyano ketones. J. Am.
applied in the synthesis of the pheromone (9Z,12Z)-tetradeca-9,12-dienyl Chem. Soc. 104, 2198–2208 (1982).
acetate. Org. Lett. 17, 2952–2955 (2015). 94. Coates, R. M. & Hobbs, S. J. α-Alkoxyallylation of activated carbonyl
81. Lin, L., Yamamoto, K., Matsunaga, S. & Kanai, M. Rhodium-catalyzed compounds. A novel variant of the Michael reaction. J. Org. Chem. 49,
cross-aldol reaction: in situ aldehyde–enolate formation from allyloxyboranes 140–152 (1984).
and primary allylic alcohols. Angew. Chem. Int. Ed. 51, 10275–10279 (2012). 95. Coates, R. M., Rogers, B. D., Hobbs, S. J., Peck, D. R. & Curran, D. P.
82. Lin, L. et al. Catalytic asymmetric iterative/domino aldehyde cross-aldol Synthesis and Claisen rearrangement of alkoxyallyl enol ethers. Evidence for a
reactions for the rapid and flexible synthesis of 1,3-polyols. J. Am. Chem. Soc. dipolar transition state. J. Am. Chem. Soc. 109, 1160–1170 (1987).
137, 15418–15421 (2015). 96. Sommer, H., Weissbrod, T. & Marek, I. A tandem iridium-catalyzed
83. Baudry, D., Ephritikhine, M. & Felkin, H. Isomerisation of allyl ethers ‘chain-walking’/Cope rearrangement sequence. ACS Catal. 9,
catalysed by the cationic iridium complex [Ir(cyclo-octa-1,5-diene) 2400–2406 (2019).
(PMePh2)2]PF6. A highly stereoselective route to trans-propenyl ethers. J. 97. Suresh, R., Massad, I. & Marek, I. Stereoselective tandem iridium-catalyzed
Chem. Soc. Chem. Commun. 1978, 694–695 (1978). alkene isomerization–Cope rearrangement of ω-diene epoxides: efficient
84. Oltvort, J. J., Van Boeckel, C. A. A., De Koning, J. H. & Van Bloom, J. H. Use access to acyclic 1,6-dicarbonyl compounds. Chem. Sci. 12, 9328–9332 (2021).
of the cationic iridium complex 1,5-cyclooctadiene-
bis[methyldiphenylphosphine]-iridium hexafluorophosphate in carbohydrate Acknowledgements
chemistry: smooth isomerization of allyl ethers to 1-propenyl ethers. Synthesis This project has received funding from the European Union’s Horizon 2020 research and
1981(305), 308 (1981). innovation program under grant agreement no. 786976.
85. Moriya, T., Suzuki, A. & Miyaura, N. A stereoselective preparation of
γ-alkoxyallylboronates via catalytic isomerization of pinacol [(E)-3-alkoxy-1-
propenyl] boronates. Tetrahedron Lett. 36, 1887–1888 (1995). Author contributions
86. Massad, I., Sommer, H. & Marek, I. Stereoselective access to fully substituted I. Massad, R.S., L.S. and I. Marek conceived and wrote the manuscript. All the authors
aldehyde-derived silyl enol ethers by iridium-catalyzed alkene isomerization. contributed to discussions.
Angew. Chem. Int. Ed. 59, 15549–15553 (2020).
87. Wang, P. Y., Massad, I. & Marek, I. Stereoselective Sc(OTf)3-catalyzed aldol Competing interests
reactions of disubstituted silyl enol ethers of aldehydes with acetals. Angew. The authors declare no competing interests.
Chem. Int. Ed. 60, 12765–12769 (2021).
88. Nelson, S. G., Bungard, C. J. & Wang, K. Catalyzed olefin isomerization
leading to highly stereoselective Claisen rearrangements of aliphatic allyl Additional information
vinyl ethers. J. Am. Chem. Soc. 125, 13000–13001 (2003). Correspondence should be addressed to Ilan Marek.
89. Stevens, B. D., Bungard, C. J. & Nelson, S. G. Strategies for expanding Peer review information Nature Synthesis thanks Guangbin Dong, Wanbin Zhang and
structural diversity available from olefin isomerization−Claisen the other, anonymous, reviewer(s) for their contribution to the peer review of this work.
rearrangement reactions. J. Org. Chem. 71, 6397–6402 (2006). Alison Stoddart managed the editorial process and peer review in collaboration with the
90. Wang, K., Bungard, C. J. & Nelson, S. G. Stereoselective olefin isomerization rest of the editorial team.
leading to asymmetric quaternary carbon construction. Org. Lett. 9, Reprints and permissions information is available at www.nature.com/reprints.
2325–2328 (2007).
91. Kerrigan, N. J., Bungard, C. J. & Nelson, S. G. Pd(II)-catalyzed aliphatic Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
Claisen rearrangements of acyclic allyl vinyl ethers. Tetrahedron 64, published maps and institutional affiliations.
6863–6869 (2008). © The Author(s), under exclusive licence to Springer Nature Limited 2022

48 Nature Synthesis | VOL 1 | January 2022 | 37–48 | www.nature.com/natsynth

You might also like