You are on page 1of 12

Document type: Review/Account

Account/Review for Frontiers of Molecular Science

Iron Catalyzed C­C-Bond Formation: From Canonical Cross Coupling


to a Quest for New Reactivity#

Alois Fürstner

Max-Planck-Institut für Kohlenforschung, 45470 Mülheim/Ruhr, Germany


E-mail: fuerstner@kofo.mpg.de

Received: October 9, 2020; Accepted: November 9, 2020; Web Released: November 17, 2020

Alois Fürstner
Alois Fürstner obtained his doctoral degree in 1987 from the Technical University of Graz, Austria. After a
postdoctoral stint in Geneva, Switzerland, and a Habilitation in Graz, he joined the Max-Planck-Institut für
Kohlenforschung, Mülheim, Germany, in 1993. In 1998, he was promoted to the rank of Director. His
research is focused on organometallic catalysis, ranging from fundamental mechanistic studies to appli-
cations in target-oriented synthesis. He received numerous awards, including the Thieme/IUPAC Prize
(2000), the ACS Arthur C. Cope Scholar Award (2002), the Mukaiyama Award (2005), the Otto-Bayer-Prize
(2006), the Janssen Pharmaceutica Prize (2008), the Prelog Medal (2011), the Karl-Ziegler-Prize (2013), the
Kitasato Medal (2013), the ACS H. C. Brown Award (2016), and the Prix Mondial Nessim Habif (2019).

Abstract questions. As a consequence, research into base metal catalysis


This Account summarizes our work in the area of organoiron has massively grown and become a mega-trend in recent years.2
chemistry during the last two decades, with special emphasis This important development notwithstanding, it is fair to say
on iron catalyzed C-C-bond formation. Specifically, it is shown that the dominance of the noble metals over large parts of
that iron catalysts can emulate reactivity more befitting noble homogeneous catalysis has not yet been challenged to the
metals in that they allow various cross coupling, cycloaddition extent one might expect.3
and cycloisomerization reactions to be carried out with sur- This conclusion may seem perplexing at first sight since
prising ease. At the same time, this base metal opens oppor- almost all noble metals are not just expensive but subject to
tunities for the discovery of genuinely new transformations. “political” price fixing as a consequence of their uneven global
distribution; the supply is un-sustained on the longer run, and
Keywords: Iron catalysis j Cross coupling j their physiological and toxicological properties are not partic-
Natural product synthesis ularly favorable either. These ® a priori ® significant draw-
backs are obviously outweighed by a number of most attractive
chemical attributes: the quest for truly competitive base metal
1. Introduction catalysts must therefore start by asking what these virtues of
There is no biocatalyst known to date that incorporates a the noble metals are and how they can be mimicked by cata-
noble metal;1 it would have been a significant evolutionary lysts based on iron and its neighbors from the fourth period
handicap if any such scarce element were indispensable for life. of the transition metals. Short such analyses have been pub-
Although the boundary conditions for chemical synthesis, as lished and are not reiterated here.4,5 Rather, the present Account
practiced by mankind since the advent of the industrial age, are outlines how the iron catalysis project in my laboratory has
certainly much different, it is arguably relevant to contemplate started with the ambition to emulate conventional cross
how metalloenzymes are able to effect so many different coupling chemistry, but has evolved since then into a search
chemistries with breathtaking ease even though they rely on for new reactivity.
base metals only. Can their chemical properties be emulated in
industrially viable settings? Can the range of transformations 2. Canonical Cross Coupling
even be extended beyond the repertoire of nature? The scien- The overwhelming success of palladium catalyzed C­C and
tific community worldwide is well aware of these and related C­X bond forming reactions may hide the fact that the term

666 | Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan
“cross coupling” had originally been coined in the context of reagents of choice.14 Many reactions proceed with excep-
iron catalysis. It was used in Kochi’s pioneering reports on tionally high rates at or (much) below room temperature.
reactions of certain Grignard reagents with alkenyl halides in Because of this favorable kinetic profile, various functional
the presence of simple iron salts;6 the first publications of this groups remain untouched, which one would expect to react
series pre-date the seminal discoveries of nickel- and palladium with Grignard reagents otherwise. Electron deficient aryl- or
cross coupling that triggered the avalanche shortly thereafter.7 heteroaryl chlorides are the most adequate substrates, whereas
Interestingly enough, scattered examples of what would nowa- triflates are typically required as the leaving group in case of
days certainly be called iron catalyzed “homo coupling” and electron-rich reaction partners. Unfortunately, the analogous
“cross coupling” had already been published prior to Kochi’s aryl-aryl coupling turned out to be of limited scope;14 since the
work, although the impact was arguably marginal.8 original report, however, much progress has been achieved in
The potential advantages notwithstanding, the scope of early this important field, even though a completely general proce-
iron catalyzed cross coupling was narrow. Largely variable dure is still awaited.24
outcomes were recorded even for alkenyl halides,6,9 and only For the many favorable attributes, iron catalyzed aryl/alkyl
few other types of electrophiles were found suitable.10 This cross coupling was rapidly embraced by the community and its
limitation was perhaps a major reason for a “latency period” scope was further extended: for example, many additional and
with little activity in the field, which lasted nearly until the turn rather “unconventional” leaving groups proved viable, which
of the millennium. Only at this point in time was the field traditional palladium catalysts fail to activate (aryl sulfamates,
thoroughly revisited: several independent reports showcased carbamates, carboxylates, phosphonates, (thio)ethers, trialkyl-
that the potential of iron catalysis for C­C-bond formation in ammonium salts etc.).18­23 Moreover, certain nucleophiles
general and cross coupling in particular had been largely other than Grignard reagents can be used, even though there
underestimated.11­14 is room for further improvement.12,18­23 Applications to natural
Our group’s first systematic foray pursued the simple but product synthesis corroborate the utility of this chemistry.25
important goal of extending iron catalyzed cross coupling to This notion is exemplified by the serial cross-coupling of the
aryl halides. Though arguably the most important type of elec- bifunctional pyridine derivative 3 with two different Grignard
trophiles, they had neither been addressed by Kochi nor anyone reagents in one pot (Scheme 2); the resulting product 5 was
else prior to our work. Gratifyingly, many aryl chlorides, then swiftly transformed into muscopyridine by a one-pot ring
-triflates and even -tosylates were found to react well with closing metathesis/hydrogenation cascade.26,27 Since iron
Grignard reagents in the presence of iron salts as simple as catalyzed cross coupling reactions scale well,28 several appli-
Fe(acac)3 (Scheme 1).14 External ligands can be used but are cations from industrial laboratories to various bioactive com-
not mandatory, whereas co-solvents or additives such as N- pounds have already been disclosed (Figure 1).29,30
methylpyrrolidone (NMP) or TMEDA usually prove advanta- In view of the foregoing, it may not come as a surprise
geous; they are believed to control the speciation in solution, that we found other electrophiles such as alkenyl triflates
even though their exact role is hardly understood at the molec- (­tosylates) or acid chlorides to be well-behaved reaction part-
ular level.15 NMP was later classified as potentially reprotoxic, ners too;31 many additional types of substrates were added to
but many benign alternatives of similar efficacy were found the list by other research groups.18­23 Once again, a growing
since.16 number of applications to natural product synthesis bears
Several lines of evidence prove that it is the iron that witness to the power of this chemistry. The straightforward
accounts for the cross coupling and not trace impurities in the synthesis of cubebene by a sequence of π-acid catalyzed
precatalyst (Cu, Pd etc.).17 Specifically: (i) iron salts of the cycloisomerization of an enyne with concomitant 1,2-acyl
highest available purity show unchanged performance; (ii) aryl shift32 followed by an iron catalyzed cross coupling is repre-
chlorides are typically much better substrates than the corre- sentative (Scheme 3).25,33­36
sponding bromides or iodides,14 very much opposite to what is
observed with copper- or palladium catalysts; (iii) the reactions 3. Steps Beyond the Canon: Alkyl Electrophiles
proceed with unprecedented rates at temperatures as low as Where applicable, iron catalysis has notable merits for being
¹78 °C in some cases, at which copper or palladium catalysts fast, productive, mild, benign and cost-effective. From a con-
usually fail to activate aryl chlorides; (iv) structurally well- ceptual perspective, however, the transformations described in
defined and fully characterized iron complexes were found that
are even better (pre)catalysts than simple iron salts.
BrMg
Because several excellent reviews on this topic are available, Cl N 4
MgBr
a summary in “al fresco” style may suffice at this point.18­23 In Cl N OTf Fe-salene (5 mol%)
80%
the original format, alkylmagnesium halides proved to be the 3 THF/NMP, 0°C 4

n-Hexyl-MgBr O N
O
N Muscopyridine
Fe(acac)3 (5 mol%) OMe
OMe

Cl THF/NMP, 0°C, <10 min


1 2 5
84% [15 g scale]

Scheme 1. Prototypical iron catalyzed aryl/alkyl coupling Scheme 2. Sequential “one-pot” cross coupling en route to
reaction. muscopyridine.

Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan | 667
O CF3 O OMe
HO N N Ph2 Zn(tmeda) 2MgX2
O Br O
AcO [L FeCl 2] (3 mol% AcO Ar2 P PAr2
N
56 g scale 210 g scale AcO OAc THF, 0°C AcO OAc Ar = 3,5(TMS)C6H3-
@ 0.1% Fe loading @ 0.1% Fe loading 9 x 245 g scale 2 kg scale OAc 96%, / = 73:27 OAc L = TMS-SciOPP
NBn2 10 11
N N(PMB)2
N N H
N N OTES OTES
Br N F3C MgBr
N
O OBn O OBn
3.8 kg scale MeO MeO
2.9 kg scale 6.9 kg scale
O OTES O OTES
FeCl 3 (10 mol%)
F3C MeO tmeda, THF
I
O
84%
OTES 12 OTES 13
O O NMe2

0.8 - 60 kg scale 6 x 50 kg scale


Me
P P
Figure 1. Larger-scale applications of iron catalyzed cross Me
O
MgBr (6 mol%)
coupling reactions of alkenyl-X and aryl-X derivatives. Cl
O
+ Fe(acac)3 (3 mol%) O
O
THF, 0°C
MeMgBr rac- 14 15 (82%, er = 90:10)
AcO PtCl2 (2 mol%) Fe(acac)3 (10 mol%)

92% H Scheme 4. Examples of advanced “inverse” alkyl/aryl cou-


H THF, -30°C Me
RO H
90%
H
pling reactions from the literature.40h,42,43
6 7 R = Ac α-cubebene (9)
8 R = OTf MesMgBr
Et2 Cl Et 2
depe (2.5 eq.) P P (2.2 eq.) Et 2 P PEt 2
Me FeCl2(THF)1.2 Fe
Me Fe
OTBS THF, RT P P THF, 70°C
H H Et2 Cl Et2
MOMO OTBS O H Mes Mes
88% 81%
Me H O O 16 17
O
Ph O
O O O
O H H H H OTBS
72% O 2,5-Me2C6H3MgBr (vycor filter)
H
95% TBSO
O 89% I cyclohexane
17 (5 mol%), THF, 70°C
18 73% 19
Scheme 3. Examples of iron catalyzed cross coupling reac-
tions of alkenyl triflates as applied to natural product
H retigeranic acid
synthesis. H
H
HO
20 21 O
the previous Section replicate no more than a small subset of
cross coupling reactions which palladium is perfectly able to Scheme 5. Iron catalyzed coupling of a sterically hindered
effect. Grignard reagent: formal synthesis of retigeranic acid.
During our initial studies, however, we became increasingly
aware that there is an orthogonal agenda in that iron catalysts as mesitylmagnesium bromide, which are low-yielding other-
are also capable of promoting transformations (largely) beyond wise.44,45 The straightforward preparation of 19 illustrates this
reach of their noble cousins. A first step in this direction was aspect (Scheme 5): this particular product had previously served
taken when we learned that aryl-Grignard reagents react as substrate of an elegant total synthesis of retigeranic acid and
exceedingly well with alkyl halides as electrophilic partners. silphiperfol-6-ene via alkene/arene meta-photocycloaddition as
In our case, this discovery was made during investigations into the key step.46
the reactivity of structurally well-defined low-valent iron com- The net propargylic substitution observed in reactions of 1-
plexes (see below).37 At the same time, other research groups alkynylcyclopropyl tosylates with alkylmagnesium halides and
had made similar observations using much simpler and hence catalytic Fe(acac)3 is deemed another noteworthy advance
more practical catalyst systems.38,39 (Scheme 6). It represents the first example of successful cross
Since then, this “inverse” alkyl/aryl coupling has become coupling of a tertiary alkyl electrophile under iron catalysis.47
a focal point of research activities, and important advances The method was used in the context of natural product chem-
were reached in preparative as well as mechanistic terms istry and the resulting structural motif is present in a number of
(Scheme 4).40,41 Somehow reciprocal to what had been found bioactive ingredients too.48,49
in reactions with aryl chlorides, the coupling of alkyl halides To properly assess this result, one has to keep in mind that a
often benefits from the use of ancillary ligands. This fact pro- variety of other propargylic substrates undergoes vinylogous
vides opportunities for fine-tuning and for asymmetric catalysis substitution with formation of the corresponding allenes.50­54
as well. The example of the racemic α-chloropropionate 14 that Such transformations allow stereochemical information to be
is converted in an enantioconvergent manner into optically relayed from a stereogenic center to a chiral axis. In view of the
active 15 is a landmark.42 high reactivity of allenes, this information content can be used
Ligand-tuning allowed yet another limitation to be over- in many different ways, as illustrated by the formation of the
come: complex 16, for example, endowed with commercially tetrasubstituted chiral center of the tetrahydrofuran derivative
available bis(diethylphosphino)ethane (depe) is able to cata- 28 (Scheme 7).55 Specifically, propargyl epoxide 26 was treat-
lyze cross coupling reactions of Grignard reagents as hindered ed with n-PrMgCl in the presence of catalytic Fe(acac)3 to

668 | Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan
MeMgCl
OTs MeLi
Fe(acac)3 (5 mol%) Me
MeLi TBSO O TBSO O
TBSO O
THF, -20°C Fe(acac) 3 cat.
Cl Me
70% X [M] [M]
Ph 22 Ph 23
29 X = O CCl 4, PPh3 C
B
30 X = CCl2 70%
MgBr OH
OTs H O
Fe(acac)3 (5 mol%)
OH O
TBSO HO
THF, -20°C 75%
H
SiMe3 24 92% SiMe3 25 31 Brefeldin A

Scheme 6. Cross coupling at tertiary propargylic centers. 47,48 Scheme 8. Iron catalyzed synthesis of internal alkynes;
application to the total synthesis of brefeldin A.
n-PrMgBr RO
RO [M] RO OH
Fe(acac) 3 cat. O
O
R = TBDPS E
Me [Fe]
MeMgBr O O[M]
Me
26 A 27 (62%, syn:anti = 8:1)
[M] COOH
O O Fe(acac)3 cat. O O Me
F
O -30°C, 93%
32 D O O 33 (dr > 20:1)
Me
AgNO3 RO O HO O
CaCO3
HO O
O Scheme 9. Iron catalyzed formal ring opening/cross
90% coupling.
28 Amphidinolide Y

Scheme 7. Iron catalyzed allene synthesis by opening of a The new entry into non-terminal alkynes has served natural
propargyl epoxide; application to the total synthesis of product synthesis well. Whereas the copper-catalyzed proce-
amphidinolide Y. dure furnished key building blocks for our recent total syn-
theses of tulearin A and C,61,62 kendomycin,65 and 5,6-
furnish allenol 27 in less than five minutes reaction time. The dihydrocineromycin B,66 the iron-catalyzed variant proved
fact that the syn-isomer was formed preferentially (dr = 8:1) is instrumental for our approaches to callyspongiolide67 and
thought to reflect a “directed delivery” of the nucleophile via brefeldin A (Scheme 8);68 all of these projects relied on ring
pre-coordination (A) of the (loaded) oxophilic iron catalyst to closing alkyne metathesis (RCAM) as an exquisite means to
the epoxide O-atom; for this stereochemical course, the reac- forge macrocyclic frameworks.69,70 The reactions were per-
tion is complementary to the anti-selective opening of propar- formed on multigram scale and show that chiral centers α to the
gyl epoxides with organocopper reagents. The axial chirality of former lactone carbonyl remain intact.
27 was then transmitted to the tertiary ether center upon acti-
vation of the allene with the aid of a carbophilic Lewis acid.56 5. Iron-Catalyzed Ring-Opening/Cross Coupling
The resulting product 28 served as key building block for the In recognition of the fact that different types of alkenyl elec-
total syntheses of amphidinolide X and Y.55,57,58 trophiles are amenable to iron catalyzed cross coupling, we
explored whether compounds comprising such a motif as part
4. Internal Alkynes Derived from Lactones of a heterocyclic ring react analogously. 2-Pyrones fall into this
Because of their activated carbonyl character, many lactones category, not least because they show little aromatic character
are amenable to Wittig-type olefination, which ordinary esters and can be viewed as cyclic enol ester derivatives.
do not undergo. This propensity can be used to access gem- In line with this notion, treatment of 32 with MeMgBr and
dichloro olefin derivatives such as 30.59 On treatment with catalytic amounts of Fe(acac)3 in toluene or Et2O at low temper-
lithium sand, these compounds afford the corresponding ature furnished acid 33 in good yield and excellent selectivity
terminal alkynes.60 We saw an opportunity to extend the scope (Scheme 9).71 For its “non-thermodynamic” Z,E-configuration,
and reach internal alkynes simply upon replacement of lithium the product is highly isomerization-prone; actually, it sufficed
by organolithium reagents RLi.61 to warm the mixture to ambient temperature prior to work-up
The desired reaction takes place even in the absence of a to transform the kinetic product into the more stable E,E-
catalyst, but it is much accelerated by either copper or iron isomer.71 To the best of our knowledge, this kind of formal “ring
salts; in the latter case, a mixture of Fe(acac)3 and 1,2- opening/cross coupling” of pyrones has no precedent in palla-
diaminobenzene proved optimal.62 This variant works partic- dium catalysis; it is enabling in that it provides access to
ularly well for exo-dichloroolefins derived from five-membered highly substituted di-unsaturated acid derivatives that are
lactones (Scheme 8). The catalyst likely promotes the initial difficult to make otherwise.
metal/halogen exchange with formation of a metal vinylidene; From the conceptual viewpoint, this transformation marks a
intermediates of type B are known to be electrophilic63 and new frontier for cross coupling in general, because the former
capable of intercepting a second equivalent of RLi to give C “leaving group” is retained as an integral part of the result-
prior to reductive elimination.62 Note that the substituent R ing product. Moreover, it shows that heterocycles may be an
attached to the triple bond is derived from the nucleophile, underutilized pool of substrates that deserves more attention.
which distinguishes this transformation from a classical Corey- Because of the exceptionally mild conditions, it was hoped
Fuchs alkynylation/alkylation sequence.64 that pyrone ring opening/cross coupling would be applicable

Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan | 669
S S
OPMB OPMB
N N
OTBDPS OTBDPS (CO)3Fe
MeMgBr HOOC Ph
BocHN Fe(acac)3 (30 mol%) BocHN O O O O
73% (97% brsm) 37
O O 22
O O O O Me
34 35

S S
OPMB NMe 2
N N
O O

O O
BocHN H2N
Pateamin A
O O
O 36 O

Figure 2. Structure of a model complex in the solid state.77


Scheme 10. Iron catalyzed pyrone ring opening/cross cou-
pling as the key step of a total synthesis of pateamin A.
dienoate but concomitantly installs the methyl branch at C22
provides opportunities to explore structure/activity relation-
to more elaborate substrates comprising functionality that ships by using nucleophiles other than MeMgBr. Moreover the
might not subsist to a Grignard reagent otherwise. The marine underlying synthesis blueprint also brought the sister com-
macrolide pateamin A72 and the synthetic analogue desmethyl- pound DMDA Pat A into reach.78
desamino-pateamine A (DMDA-Pat A)73 were identified as a An important aspect yet to be discussed is the mechanism
particularly relevant testing ground. These potently cytotoxic of this new formal ring opening/cross coupling manifold. The
compounds exhibit differential activity in xenograft melanoma fact that the R group delivered by the Grignard reagent replaces
models in vivo.74 Moreover, pateamin A is the only potential the ester leaving group with retention of configuration of the
lead compound known to date which shows activity against enol double bond corresponds to the pattern of a true cross
cachexia, a muscle wasting syndrome that often takes a lethal coupling process (Scheme 9). Much evidence, however, speaks
course with immunosuppressed or cancer patients. Moreover, against a canonical course, in which the iron catalyst ® before
it is a very valuable probe molecule for chemical biology as or after reaction with the Grignard reagent ® would oxidatively
it brings protein biosynthesis to a hold by blocking the initial insert into the C­O bond. Rather, it is thought that the reaction
translation steps; its target protein is the eukaryotic initiation proceeds via π-complex formation (D) followed by delivery of
factor 4A.74 the nucleophile in what is best described as a 1,6-addition
Since pateamin A derives from a marine sponge, it is in reaction (Scheme 9).71,77 This proposal gains credence from the
extremely short supply. A total synthesis that is “scalable” to fact that iron catalyzed 1,6-additons to dienoates and related
the extent that it allows in-depth biological and pharmacolog- substrates are known to be facile.79 Finally, the resulting
ical profiling of this lead compound to be carried out meets a enolate of type E (or F) succumbs either to a torquoselective
real demand. Any such project faces several difficulties, but the electrocyclic ring-opening or is broken via an ionic mechanism.
formation of the non-thermodynamic Z,E-dienoate spanning Although no intermediate could be characterized in situ or
the macrodiolide ring is critical because the integrity of this ex situ, the structure of the model iron carbonyl complex 37
substructure is innately linked to the biological activity of the is deemed informative (Figure 2):77 it shows the purported η4-
compound. binding of the metal fragment to the pyrone π-system. Substan-
We were pleased to learn that a sequence of gold catalyzed tial back-bonding from iron into the antibonding orbitals of the
pyrone formation, which was developed in our laboratory in ligand entails re-hybridization and forces the heterocyclic ring
parallel work,75,76 followed by iron catalyzed pyrone opening to bend; as a result, orbital overlap between oxygen and the
served the purpose very well (Scheme 10). Thus, treatment of double bond is disrupted and the enol ester character largely
the N-Boc protected derivative 34 with excess MeMgBr and lost. Equally if not more important is the fact that C5, as the
Fe(acac)3 (30 mol%) at ¹60 °C furnished the desired product center to which the nucleophile will be delivered, gains signif-
35 in 79% yield (97% brsm) with an isomer ratio of ²17:1.77 icant “carbenoid” character in that it carries a carbon-metal
This example illustrates the compatibility of the reaction with bond and a leaving group. The electron back-donation referred
various electrophilic substituents as well as its insensitivity vis- to above alters the bond lengths throughout the ring: most
à-vis the donor sites of the adjacent thiazole. Since the reaction notably, the C3­C4 bond is shortened as if it “anticipates” the
directly furnished the free acid after work-up, the further bonding situation of the 1,6-addition product to be formed as a
elaboration into the desired macrocycle 36 by subsequent transient intermediate prior to ring opening. Collectively, these
lactonization was greatly facilitated and the risk of epimeriza- structural features are thought to render delivery of the R-group
tion minimized. Product 36 thus formed then paved the way to of RMgX to C5 not only possible but actually facile, very
pateamin A.77 much in line with what is experimentally observed.71,77
The strategy to encode the isomerization-prone Z,E- Whether the subsequent ring opening of the resulting enolate
configured dienoate in the form of a robust heterocycle, from proceeds in an electrocyclic or ionic manner remains unclear at
which it is revealed only at late stage, paid valuable dividends this point; indirect evidence speaks for the latter mechanism.77
in that it rendered the overall synthesis short, robust and pro- This mechanistic picture is consistent with additional data
ductive.77 The fact that iron catalysis does not only unmask the from the 2-pyridone series (Scheme 11).80 For the poorer leav-

670 | Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan
PhMgBr E E
Fe(acac)3 (5 mol%) PhMgBr
Et2O, -30°C Fe(acac)3 (5 mol%)
[Fe]
Ph O NHTs then DMF N O Ph N O
THF, -45°C E E
R Bn R E E E
R = Ts, 60% R = Bn, 96%
40 38 39 G E
[CpFe(ethylene)2]Li

Scheme 11. Iron catalyzed 1,6-addition versus formal ring R


cat.
E E
C
[Fe] R
opening/cross coupling. 45 42
E
E
R
[Fe]
42
67% ([4+2])
H

E E Scheme 13. The iron catalyzed [4+2] cycloaddition com-


E E mences at the enyne rather than the diene subunit.

E E
E E
lary alkenes by the substrate, followed by oxidative cyclization
Ph
[CpFe(olefin)2]Li with formation of the corresponding metallacycles which then
Ph E E E = COOEt
E E
Ph evolve into the observed products. For the Alder-ene reaction,
41 43 detailed labelling studies allowed conceivable alternative sce-
75% ([5+2]) 95% (Alder-ene)
narios to be ruled out (allylic C­H activation or metal hydride
Ph

E
addition/Heck reaction pathway).86 In the case of the metal-
E
E E catalyzed Diels-Alder cycloaddition, it is not intuitive whether
E 44
a substrate of type 45 initially forms a metallacycle G derived
E 89% ([2+2+2]) from the 1,3-diene entity, into which the alkyne then inserts;
45 could also behave as an enyne in the first place, that gives
Scheme 12. Prototypical cycloaddition and cycloisomeri-
rise to a metallacycle H which expands to give the final product
zation reactions catalyzed by CpFe(0) complexes.
(Scheme 13). Control experiments showed that 1,3-butadiene
ing group properties of an amide, it was possible to intercept is incapable of replacing the ethylene ligands of [CpFe-
and isolate 1,6-addition products such as 39. Upon activation of (alkene)2]Li, which in turn suggests that a pathway via G is
the 2-pyridone by N-tosylation, formal ring opening/cross cou- unlikely to be operative; rather, the alkyne is needed for the
pling with formation of 40 takes over, as one might expect.80 reaction to take place.89

6. Iron Catalyzed Cycloisomerization 7. Vignette: Perplexing Redox Events


and Cycloaddition Reactions As mentioned above, CpFe(0) complexes such as 43 and
The ability of low-valent iron to ligate π-bonds of all sorts is 44 originally used in our laboratory to study iron catalyzed
textbook knowledge;81,82 the affinity to alkynes and 1,3-dienes cycloisomerization and cycloaddition chemistry are formed on
is particularly high. Therefore, and in consideration of some treatment of ferrocene (Cp2Fe) with lithium in the presence of
early literature precedent,83 it seemed likely that iron cata- the appropriate olefin (Scheme 14).85,86 Under slightly more
lysts are able to promote various types of cycloaddition and forcing conditions, both Cp-rings of Cp2Fe are reductively
cycloisomerization reactions known from the realm of noble cleaved off, leading to the formation of the tetraethylene ferrate
metal catalysis otherwise.84 complex 45 or the analogous complex 46, each comprising an
Complexes of type [CpFe(alkene)2]Li, which were used at iron center of the formal oxidation state ¹2 and a formal d10
the outset of our investigations, are readily formed on scale electron count.85,90 As such, they are cousins of Collman’s rea-
from ferrocene by reductive cleavage of one cyclopentadienyl gent Na2[Fe(CO)4];91 for the much more labile ligand sphere,
ring in the presence of either ethylene or 1,5-cyclooctadiene however, they are inherently more reactive.
(COD) (see below).85 As can be deduced from the structure in Though obviously very sensitive to air and moisture, such
the solid state, the very electron-rich, formally zero-valent iron “exotic” complexes are routinely made on º10 g scale from
center is strongly engaged with the π*-orbitals of the ligand cheap starting materials and therefore provide a unique plat-
through electron back-donation, even though the olefins remain form for the study of “ultra-low-valent” iron.86,90 Iron species
kinetically labile.86 These properties readily explain why such of formally negative oxidation states were originally thought to
complexes catalyze [4+2], [5+2] and [2+2+2] cycloaddition play a key role in cross coupling,14 because Grignard reagents
as well as Alder-ene reactions of appropriate substrates had previously been reported to transform iron salts into such
(Scheme 12);86,87 a number of reducible functional groups, highly reduced entities.92 Although later (spectroscopic) results
basic sites and C­H acidic terminal alkynes are tolerated. It is speak against a systematic involvement,93 it remains a matter of
of particular note in the present context that these types of fact that 45 and 46 are perhaps the most active catalysts for
transformations had originally been developed with noble various iron catalyzed C­C bond forming reactions known to
metal catalysts;84 our results prove that low-valent iron is a date.37,90
serious alternative.88 Reduction of FeX2 ligated to phosphine ligands with lithium
Compelling evidence suggests that the reactions catalyzed under ethylene atmosphere takes a more conventional course in
by [CpFe(alkene)2]Li commence by exchange of the ancil- that it stops at the Fe(0) stage. In our hands, the use of EtMgBr

Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan | 671
[M]
Fe(acac)3
(tmeda)Li Fe Li(tmeda) X X
RMgX
45 [Fe] K H L
X [Fe] X
Li, ethylene, (5 atm) H
43% H
RT, then tmeda
RMgX
[Fe] XMgX
MgX
I X
Li, cod Li, ethylene (1 atm) J
Li (dme) Li (tmeda)
Fe Cp2Fe Fe
DME, -50°C to RT -50°C to 0°C [Fe] R
44 43 [M]
97% then tmeda, 45% M N
R

22% Li, ethylene, (5 atm)


DME, then cod, 70°C
Scheme 16. A novel cycloisomerization/cross coupling
cascade, merging bond-forming and bond-breaking events.
[(cod)2Fe] [Li(dme)]2 (46)
OH
O MgX
Scheme 14. Formation of low-valent iron complexes from COOMe COOMe

ferrocene. [Fe(acac)3] cat.


Me
Me THF, -30°C 51
50
Cl 90%, E/Z > 99:1
R Fe EtMgBr (excess) R Fe0
P Cl P two C-C-bonds are formed
E
P R P R E E BrMg O
R ethylene, THF, -5°C R one C-C-bond is broken!
R R O E
COOMe
47 R = iPr 77-97% COOMe
48
Fe(acac)3 cat. O
63%, E:Z > 98:2
52 53 O
C6H11 MgCl (excess) +4 E = COOMe
Fe
FeCl 2
THF, -35°C Scheme 17. Iron catalyzed cycloisomerization/cross
20-25% (+ [Fe] colloid] 49
coupling.
Scheme 15. The strikingly different outcome of reactions of
iron salts with Grignard reagents bearing β-hydrogen bond-breaking/bond-making events downstream of metalla-
atoms. cycle formation.96
Specifically, placement of a potential leaving group X next
instead of the alkali metal proved much more reliable.94 The to the alkene as shown in I might give ring opening a chance to
reduction of 47 to 48 on (multi)gram-scale is representative outperform β-hydride elimination because of the favorable
(Scheme 15). This well-characterized complex constitutes alignment of the C­X σ*-antibonding orbital with the electron-
another valuable starting point for forays into the underex- rich C­M bond within the emerging rigid bicycle J
plored organometallic chemistry of low-valent iron.89,94 (Scheme 16). However, such a bond-breaking event does not
To illustrate the intricacies that one may encounter, reference regenerate a catalytically active low-valent metal species in the
is made to the seemingly analogous reaction of FeCl2 with way β-hydride elimination ultimately would (J¼K¼L); there-
cyclohexylmagnesium chloride, which leads to a completely fore the envisaged ring opening step must be juxtaposed with
different outcome: [Fe(C6H11)4] (49) was the only defined prod- an alternative process that regenerates the active catalyst and
uct that could be isolated from the mixture (Scheme 15).95 This ensures turn-over.
complex represents an extremely scarce example of a homo- Apprehensive of the tendency of (low-valent) iron to form
leptic Fe(+4) species; it is diamagnetic (!) despite the tetra- ate-complexes on exposure to Grignard reagents or related
hedral ligand environment. The formation of 49 can only be nucleophiles (see the next Vignette), we planned to make use of
explained by disproportionation of an Fe(+2) intermediate, cross coupling for this very purpose.96 If a “loaded” iron spe-
with formation of colloidal iron as the second product.95 The cies M transfers the R-group to the emerging product N, the
meta-stability of 49 is perplexing since the central iron atom is reductive bond formation should adjust the oxidation state and
surrounded by ligands bearing no less than 20 H-atoms amena- hence close a catalytic cycle. At the same time, the Mg(+2)
ble to α- or β-hydride elimination. In any case, the comparison cation might assist the overall process by lowering the barriers
shown in Scheme 15 illustrates that extrapolations from one upon coordination to the (heteroatom) leaving group X in-
reaction to a seemingly similar reaction can be misleading. This scribed into the frame of M.
conclusion, in turn, implies that our knowledge of some of the The conceived scenario was reduced to practice with the
rather basic chemistry of iron is still in its infancy. aid of structurally well-defined iron complexes but also with
iron catalysts generated in situ from Fe(acac)3 and Grignard
8. Merging Iron Catalyzed Cycloisomerization reagents (Scheme 17).96 Unsurprising perhaps in view of the
and Cross Coupling results previously obtained with iron catalyzed canonical cross
As outlined above, all available evidence speaks for the coupling,14 alkylmagnesium halides perform better than their
intervention of ferracycles as key intermediates in the cyclo- aromatic siblings. They afford functionalized 1,3-diene prod-
addition and cycloisomerization reactions shown in ucts such as 51 comprising stereodefined tetrasubstituted
Schemes 12 and 13. In consideration thereof, we envisaged a alkenes in good yields and generally excellent selectivity.96
conceptually new reaction manifold based on unprecedented Not only heteroatoms were found to be suitable leaving groups;

672 | Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan
rather, the malonate derivative 52 proves that even activated C­ and only three organic residues in the first coordination
C-bonds can be cleaved by this catalytic process: in the case of sphere.44,98
product 53, two new C­C bonds were formed while one C­C More recent studies using advanced spectroscopies to study
bond was concomitantly broken during this unorthodox cyclo- freeze-trapped intermediates generated in situ confirmed the
isomerization process.96 intervention of ate-complexes in several iron catalyzed C­C-
It is appropriate to mention here that palladium catalysts bond forming processes. Moreover, it has been convincingly
can serve similar purposes. Actually, they allow aryl groups to shown that ate-complexes are subject to facile ligand exchange
be transferred, which the iron-based system does not do well. and evolve with time to lower oxidation states.99 They can
What is remarkable though is the fact that iron catalysis undergo either two-electron reduction with formation of Fe(0)
proceeds rapidly at ¹30 °C, whereas the palladium catalyzed or single-electron reduction with formation of Fe(+1). Strik-
reactions mandate heating to +90 °C.96 This differential illus- ingly, the latter process is neither the result of homolytic cleav-
trates the exceptional driving force pooled in the iron catalyst, age or disproportionation, as one might tend to think, but rather
which might well be harnessed in other settings too. a consequence of the transient formation of dinuclear entities,
which must be accessible on steric ground for the reaction to
9. Vignette: Iron Ate-Complexes proceed. This insight illustrates the intricacy of this chemistry
Early observations suggested that reactions of iron salts with and arguably provides guidance for future catalyst design.99,100
higher alkyl-Grignard reagents lead to the formation of low- A particular highlight is the study into the ferrate species
valent iron complexes by β-hydride elimination reactions formed on treatment of FeX3 with excess RMgBr (R = Me, Et).
and/or reductive coupling processes (compare complex 48, In the absence of NMP, ferrate cluster species of the formal
Scheme 15).14,31 In contrast, nucleophiles that cannot undergo composition [R12Fe8][MgCl(thf )] are generated;101,102 in the
either reduction mechanism were supposed to furnish iron-ate presence of NMP, in contrast, [R3Fe][Mg(nmp)6] is formed
complexes in analogy to the well-known cuprates invoked in in situ (even in case of R = Et, where β-hydide elimination is
much of organocopper chemistry.14,31 expected to take place).15,103 These latter complexes almost
The first such ferrate species that was shown to be relevant in certainly represent the reactive intermediates accountable for
the context of iron cross coupling was a “super-ate” complex the seminal results described by Kochi and coworkers on the
formed on reaction of FeCl2 with excess MeLi.97 The resulting iron catalyzed cross coupling of alkenyl halides.6 One must
homoleptic species [(Me4Fe)(MeLi)][Li¢(OEt2)]2 (54) carries appreciate the difficulty in characterizing reactive complexes of
four methyl groups in a tetrahedral arrangement about the this level of complexity, which are extremely air- and moisture
Fe(+2) center (Figure 3); three of them entertain additional sensitive, and, in case of [R12Fe8][MgCl(thf )], also paramag-
contacts to three escorting lithium counterions, which form a netic. This formidable “analytical frontier”4 is a major hurdle
tetrahedral metal framework with the iron center. Strikingly and the progress in the understanding of iron catalysis achieved
though, the cluster incorporates a fifth methyl group (C5) in recent year therefore all the more impressive.
devoid of any direct interaction with the iron atom.90,97 This
unusual complex proved competent, inter alia, for the alkyla- 10. A Case Study in C­H Activation
tion of alkenyl halides, alkenyl triflates, and acid chlorides, as Base metal catalysts in general and iron catalysts in par-
well as for pyrone ring opening/cross coupling.71,90 PhLi ticular show great promise in C­H activation reactions.104­106
afforded the related ate-complex [Ph4Fe][Li¢(OEt2)2][Li¢(1,4- As this topic is beyond the scope of the present Account, a
dioxane)], even though it lacks a fifth RLi entity.90 In con- single case study must suffice to illustrate the point. Specif-
trast, more bulky aryl-Grignard reagents give “ordinary” ate- ically, ligand exchange between the Fe(0) ethylene complex 48
complexes of type [Ar3Fe]¹ carrying a single negative charge with 1,3-cyclohexadiene takes a rather intricate course in that
it leads to the clean formation of 55 (Scheme 18).89,94 This
outcome is best rationalized by assuming that replacement of
one ethylene ligand by the diene is followed by instant acti-
vation of the allylic C­H bond. Migratory insertion of the
remaining ethylene into the iron hydride transiently formed
gives an iron ethyl intermediate; reductive ligand coupling

R Fe P R P R
P R P Fe H R P Fe H
P R R R R R
R
R 48

Figure 3. Structure of the dianionic “super-ate” complex R Fe R


Fe
P P R
[(Me4Fe)(MeLi)][Li¢(OEt2)]2 (54) in the solid state;97 note P R P R
H
P
Fe
H
R
R
R P R 55
that only C1, C2, C3 and C4 are bound to the Fe(+2) R R
R
center, whereas C5 has no direct contact; this additional
MeLi “ligand” likely serves to complete a stable tetra- Scheme 18. Iron-mediated allylic C-H activation/coupling
hedral metal core comprised of Li1, Li2, Li3 and Fe1. cascade.

Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan | 673
entails C­C bond formation and regenerates a low-valent iron inorganic Chemistry - Inorganic Elements in the Chemistry of Life:
center poised for the next C­H activation event that forms the An Introduction and Guide, 2nd Ed., Wiley, 2013.
final product 55.89,94 The endo-orientation of the ethyl sub- 2 R. M. Bullock, ed., Catalysis without Precious Metals.
stituent in 55, which was ascertained by X-ray diffraction, Wiley-VCH: Weinheim, 2010.
supports this involved mechanistic scenario. 3 Polymer synthesis is arguably the most important excep-
The elementary steps underlying the cascade comprised of tion.
two consecutive C­H activation events and an interwoven 4 A. Fürstner, ACS Cent. Sci. 2016, 2, 778.
5 R. M. Bullock, J. G. Chen, L. Galiardi, P. J. Chirik, O. K.
stereoselective C­C-bond forming step are deemed enabling:
Farha, C. H. Hendon, C. W. Jones, J. A. Keith, J. Klosin, S. D.
the overall reaction can also be seen as the prototype of an
Minteer, R. H. Morris, A. T. Radosevich, T. B. Rauchfuss, N. A.
unconventional “cross coupling” process employing cheap
Strotman, A. Vojvodic, T. R. Ward, J. Y. Yang, V. Surendranath,
ethylene as pre-nucleophile in lieu of a preformed organo-
Science 2020, 369, 786.
metallic reagent. 6 a) M. Tamura, J. K. Kochi, J. Am. Chem. Soc. 1971, 93,
11. Conclusion 1487. b) S. M. Neumann, J. K. Kochi, J. Org. Chem. 1975, 40,
599. c) R. S. Smith, J. K. Kochi, J. Org. Chem. 1976, 41, 502.
This personal Account article does not provide a compre- d) J. K. Kochi, Acc. Chem. Res. 1974, 7, 351.
hensive treatise of iron catalyzed C­C bond formation. Rather 7 For a historical perspective, see: C. C. C. Johansson
it is meant to showcase that the use of this base metal (and its Seechurn, M. O. Kitching, T. J. Colacot, V. Snieckus, Angew.
neighbors from the fourth period) provides valuable oppor- Chem., Int. Ed. 2012, 51, 5062.
tunities in “conventional” cross coupling; at the same time, it 8 a) M. S. Kharasch, E. K. Fields, J. Am. Chem. Soc. 1941,
opens intriguing perspectives for the discovery of genuinely 63, 2316. b) M. S. Kharasch, R. Morrison, W. H. Urry, J. Am.
novel transformations. Chem. Soc. 1944, 66, 368. c) M. S. Kharasch, F. L. Lambert, W. H.
The awareness that sustained catalysis mandates the devel- Urry, J. Org. Chem. 1945, 10, 298. d) G. Vavon, C. Chaminade, G.
opment of systems operating with base metals in general and Quesnel, C. R. Hebd. Seances Acad. Sci. 1945, 220, 850. e) W. C.
iron in particular is rapidly increasing. Yet, the understanding Percival, R. B. Wagner, N. C. Cook, J. Am. Chem. Soc. 1953, 75,
of many of their fundamental properties lags behind: What are 3731.
the best ways to characterize fleeting paramagnetic intermedi- 9 G. A. Molander, B. J. Rahn, D. C. Shubert, S. E. Bonde,
ates?41,107 Where and how does the spin-state come into play Tetrahedron Lett. 1983, 24, 5449.
10 a) T. Mukaiyama, T. Takeda, M. Osaki, Chem. Lett. 1977,
and where does spin crossover intervene (“two-state reactiv-
1165. b) J. L. Fabre, M. Julia, J.-N. Verpeaux, Tetrahedron Lett.
ity”)?108­110 How can ligands be designed that allow the spin-
1982, 23, 2469. c) C. Cardellicchio, V. Fiandanese, G. Marchese,
state to be controlled? Will a given complex favor single-
L. Ronzini, Tetrahedron Lett. 1985, 26, 3595. d) A. Yanagisawa,
electron or two-electron transfer?111 In which cases is dispro- N. Nomura, H. Yamamoto, Tetrahedron 1994, 50, 6017.
portionation a shortcut for “regular” redox events? How can 11 a) G. Cahiez, H. Avedissian, Synthesis 1998, 1199. b) G.
“non-innocent” ligands and/or other metal/ligand cooperativ- Cahiez, S. Marquais, Pure Appl. Chem. 1996, 68, 53.
ity be used to advantage?112,113 12 The first example from our laboratory is contained in a
Certain answers to these and related questions can be found study on organomanganese reagents, see: A. Fürstner, H. Brunner,
in the pertinent literature, but a fully comprehensive picture Tetrahedron Lett. 1996, 37, 7009.
does not emerge from the current state-of-the-art; some of the 13 a) M. Nakamura, A. Hirai, E. Nakamura, J. Am. Chem. Soc.
examples discussed above corroborate this notion. Although 2000, 122, 978. b) M. Nakamura, K. Matsuo, T. Inoue, E.
the challenges in this field of research are tremendous, the Nakamura, Org. Lett. 2003, 5, 1373.
ingenuity of the next generation of scientists is certainly up to 14 a) A. Fürstner, A. Leitner, M. Méndez, H. Krause, J. Am.
the task; it is my conviction that an interdisciplinary approach Chem. Soc. 2002, 124, 13856. b) A. Fürstner, A. Leitner, Angew.
using ever-improved analytical tools in concert with much Chem., Int. Ed. 2002, 41, 609.
refined computational methods114 is the most promising way 15 For an important exception, see: S. B. Munoz, S. L.
forward. Daifuku, J. D. Sears, T. M. Baker, S. H. Carpenter, W. W.
Brennessel, M. L. Neidig, Angew. Chem., Int. Ed. 2018, 57, 6496.
I thank all coworkers and collaboration partners who 16 a) E. Bisz, M. Kardela, M. Szostak, ChemCatChem 2019,
participated in the iron catalysis project for their invaluable 11, 5733. b) E. Bisz, M. Szostak, J. Org. Chem. 2019, 84, 1640.
c) G. Cahiez, G. Lefèvre, A. Moyeux, O. Guerret, E. Gayon, L.
intellectual and experimental contributions; their names appear
Guillonneau, N. Lefevre, Q. Gu, E. Zhou, Org. Lett. 2019, 21,
in the references. Generous financial support by the Max-
2679.
Planck-Society and the Fonds der Chemischen Industrie is
17 Putative iron catalyzed C­N, C­O, and C­S coupling
gratefully acknowledged. reactions were later shown to be caused by copper impurities in the
chosen iron salts, see: S. L. Buchwald, C. Bolm, Angew. Chem.,
References Int. Ed. 2009, 48, 5586.
# Dedicated to Prof. Eiichi Nakamura on the occasion of his 18 E. Nakamura, T. Hatakeyama, S. Ito, K. Ishizuka, L. Ilies,
70th birthday and in recognition of his seminal contributions to M. Nakamura, Org. React. 2014, 83, 1.
science, including pioneering work on iron catalysis. 19 I. Bauer, H.-J. Knölker, Chem. Rev. 2015, 115, 3170.
1 a) I. Bertini, H. B. Gray, S. J. Lippard, J. S. Valentine, ed., 20 A. Fürstner, Org. Synth. 2019, 96, 1.
Bioinorganic Chemistry, University Science Books: Mill Valley, 21 a) B. D. Sherry, A. Fürstner, Acc. Chem. Res. 2008, 41,
CA, 1994. b) W. Kaim, B. Schwederski, A. Klein, ed., Bio- 1500. b) A. Fürstner, Angew. Chem., Int. Ed. 2009, 48, 1364. c) A.

674 | Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan
Fürstner, R. Martin, Chem. Lett. 2005, 34, 624. 35 A. Hamajima, M. Isobe, Org. Lett. 2006, 8, 1205.
22 a) A. Piontek, E. Bisz, M. Szostak, Angew. Chem., Int. 36 Q. Liu, Y. Deng, A. B. Smith, J. Am. Chem. Soc. 2017, 139,
Ed. 2018, 57, 11116. b) D. Lübken, M. Saxarra, M. Kalesse, 13668.
Synthesis 2019, 51, 161. c) G. Cahiez, A. Moyeaux, J. Cossy, 37 R. Martin, A. Fürstner, Angew. Chem., Int. Ed. 2004, 43,
Adv. Synth. Catal. 2015, 357, 1983. d) O. M. Kuzmina, A. K. 3955.
Steib, A. Moyeux, G. Cahiez, P. Knochel, Synthesis 2015, 47, 38 M. Nakamura, K. Matsuo, S. Ito, E. Nakamura, J. Am.
1696. e) T. L. Mako, J. A. Byers, Inorg. Chem. Front. 2016, 3, Chem. Soc. 2004, 126, 3686.
766. f ) T. Mesganaw, N. K. Garg, Org. Process Res. Dev. 2013, 39 T. Nagano, T. Hayashi, Org. Lett. 2004, 6, 1297.
17, 29. g) W. M. Czaplik, M. Mayer, J. Cvengroš, A. 40 See the following for leading references and literature cited
Jacobi von Wangelin, ChemSusChem 2009, 2, 396. h) E. B. therein: a) R. B. Bedford, D. W. Bruce, R. M. Frost, M. Hird,
Bauer, Curr. Org. Chem. 2008, 12, 1341. Chem. Commun. 2005, 4161. b) G. Cahiez, V. Habiak, C. Duplais,
23 a) B. Plietker, ed., Iron Catalysis. Fundamentals and A. Moyeux, Angew. Chem., Int. Ed. 2007, 46, 4364. c) T.
Applications, Springer: Heidelberg, 2011; Top. Organomet. Chem., Hatakeyama, S. Hashimoto, K. Ishizuka, M. Nakamura, J. Am.
Vol. 33. b) E. B. Bauer, ed., Iron Catalysis II; Ed.; Springer Chem. Soc. 2009, 131, 11949. d) R. B. Bedford, M. Betham, D. W.
International Publishing: Cham, CH, 2015; Top. Organomet. Bruce, A. A. Danopoulos, R. M. Frost, M. Hird, J. Org. Chem.
Chem., Vol. 50. 2006, 71, 1104. e) R. R. Chowdhury, A. K. Crane, C. Fowler, P.
24 For leading references, see the following and literature Kwong, C. M. Kozak, Chem. Commun. 2008, 94. f ) A. Guérinot,
cited therein: a) T. Hatakeyama, S. Hashimoto, K. Ishizuka, M. S. Reymond, J. Cossy, Angew. Chem., Int. Ed. 2007, 46, 6521.
Nakamura, J. Am. Chem. Soc. 2009, 131, 11949. b) I. Sapounzis, g) G. Bauer, M. D. Wodrich, R. Scopelliti, X. Hu, Organometallics
W. Lin, C. C. Kofink, C. Despotopoulou, P. Knochel, Angew. 2015, 34, 289. h) L. Adak, S. Kawamura, G. Toma, T. Takenaka,
Chem., Int. Ed. 2004, 43, 1454. c) L. Wang, Y.-M. Wei, Y. Zhao, K. Isozaki, H. Takaya, A. Orita, H. C. Li, T. K. M. Shing, M.
X.-F. Duan, J. Org. Chem. 2019, 84, 5176. Nakamura, J. Am. Chem. Soc. 2017, 139, 10693.
25 a) J. Legros, B. Figadère, Nat. Prod. Rep. 2015, 32, 1541. 41 a) R. Agata, H. Takaya, H. Matsuda, N. Nakatani, K.
b) J. E. Zweig, D. E. Kim, T. R. Newhouse, Chem. Rev. 2017, 117, Takeuchi, T. Iwamoto, T. Hakateyama, M. Nakamura, Bull. Chem.
11680. c) P. DaBell, S. P. Thomas, Synthesis 2020, 52, 949. Soc. Jpn. 2019, 92, 381. b) J. D. Sears, P. G. N. Neate, M. L.
26 A. Fürstner, A. Leitner, Angew. Chem., Int. Ed. 2003, 42, Neidig, J. Am. Chem. Soc. 2018, 140, 11872. c) T. Parchomyk, K.
308. Koszinowski, Synthesis 2017, 49, 3269.
27 For other early applications from this laboratory, see: a) B. 42 M. Jin, L. Adak, M. Nakamura, J. Am. Chem. Soc. 2015,
Scheiper, F. Glorius, A. Leitner, A. Fürstner, Proc. Natl. Acad. Sci. 137, 7128.
U.S.A. 2004, 101, 11960. b) A. Fürstner, D. De Souza, L. Parra- 43 C. Gregg, C. Gunawan, A. W. Y. Ng, S. Wimala, S.
Rapado, J. Jensen, Angew. Chem., Int. Ed. 2003, 42, 5358. c) G. Wickremasinghe, M. A. Rizzacasa, Org. Lett. 2013, 15, 516.
Seidel, D. Laurich, A. Fürstner, J. Org. Chem. 2004, 69, 3950. 44 C.-L. Sun, H. Krause, A. Fürstner, Adv. Synth. Catal. 2014,
d) A. Fürstner, D. De Souza, L. Turet, M. D. B. Fenster, L. Parra- 356, 1281.
Rapado, C. Wirtz, R. Mynott, C. W. Lehmann, Chem.®Eur. J. 45 For a different diphosphine ligand that enables similar
2007, 13, 115. e) A. Fürstner, D. Kirk, M. D. B. Fenster, C. Aissa, transformations, see: T. Hatakeyama, Y. Fujiwara, Y. Okada, T.
D. De Souza, C. Nevado, T. Tuttle, W. Thiel, O. Müller, Chem.® Itoh, T. Hashimoto, S. Kawamura, K. Ogata, H. Takaya, M.
Eur. J. 2007, 13, 135. Nakamura, Chem. Lett. 2011, 40, 1030.
28 A. Fürstner, A. Leitner, G. Seidel, Org. Synth. 2004, 81, 33. 46 a) P. A. Wender, S. K. Singh, Tetrahedron Lett. 1990, 31,
29 a) N. Tewari, N. Maheshwari, R. Medhane, H. Nizar, M. 2517. b) P. A. Wender, S. K. Singh, Tetrahedron Lett. 1985, 26,
Prasad, Org. Process Res. Dev. 2012, 16, 1566. b) P. J. Rushworth, 5987.
D. G. Hulcoop, D. J. Fox, J. Org. Chem. 2013, 78, 9517. c) C. 47 D. J. Tindall, H. Krause, A. Fürstner, Adv. Synth. Catal.
Risatti, K. J. Natalie, Jr., Z. Shi, D. A. Conlon, Org. Process Res. 2016, 358, 2398.
Dev. 2013, 17, 257. d) F. Bartoccini, G. Piersanti, S. Armaroli, A. 48 C. L. Hugelshofer, V. Palani, R. Sarpong, J. Org. Chem.
Cerri, W. Cabri, Tetrahedron Lett. 2014, 55, 1376. e) S. M. 2019, 84, 14069.
Andersen, M. Bollmark, R. Berg, C. Fredriksson, S. Karlsson, C. 49 For a very recent report on the cross coupling of a tert-
Liljeholm, H. Sörensen, Org. Process Res. Dev. 2014, 18, 952. iodide with a bicyclo[1.1.1]pentane backbone, see: J. Nugent,
f ) S. Gangula, U. K. Neelam, S. R. Baddam, V. H. Dahanukar, R. B. R. Shire, D. F. J. Caputo, H. D. Pickford, F. Nightingale, I. T. T.
Bandichhor, Org. Process Res. Dev. 2015, 19, 470. g) P. Mullens, Houlsby, J. J. Mousseau, E. A. Anderson, Angew. Chem., Int. Ed.
E. Cleator, M. McLaughlin, B. Bishop, J. Edwards, A. Goodyear, 2020, 59, 11866.
T. Andreani, Y. Jin, J. Kong, H. Li, M. Williams, M. Zacuto, Org. 50 D. J. Pasto, S.-K. Chou, A. Waterhouse, R. H. Shults, G. F.
Process Res. Dev. 2016, 20, 1075. h) G. Cahiez, O. Guerret, A. Hennion, J. Org. Chem. 1978, 43, 1385.
Moyeux, S. Dufour, N. Lefèvre, Org. Process Res. Dev. 2017, 21, 51 A. Fürstner, M. Méndez, Angew. Chem., Int. Ed. 2003, 42,
1542. i) P. Chourreu, O. Guerret, L. Guillonneau, E. Gayon, G. 5355.
Lefèvre, Org. Process Res. Dev. 2020, 24, 1335. 52 B. D. Sherry, A. Fürstner, Chem. Commun. 2009, 7116.
30 For iron-catalyzed coupling in flow, see: Y. Deng, X.-J. 53 a) S. N. Kessler, J.-E. Bäckvall, Angew. Chem., Int. Ed.
Wei, X. Wang, Y. Sun, T. Noel, Chem.®Eur. J. 2019, 25, 14532. 2016, 55, 3734. b) S. N. Kessler, F. Hundemer, J. E. Bäckvall, ACS
31 B. Scheiper, M. Bonnekessel, H. Krause, A. Fürstner, Catal. 2016, 6, 7448.
J. Org. Chem. 2004, 69, 3943. 54 For an exception, see: I. Manjon-Mata, M. T. Quiros, E.
32 V. Mamane, T. Gress, H. Krause, A. Fürstner, J. Am. Chem. Bunuel, D. J. Cardenas, Adv. Synth. Catal. 2020, 362, 146.
Soc. 2004, 126, 8654. 55 O. Lepage, E. Kattnig, A. Fürstner, J. Am. Chem. Soc.
33 A. Fürstner, P. Hannen, Chem.®Eur. J. 2006, 12, 3006. 2004, 126, 15970.
34 A. Fürstner, A. Schlecker, Chem.®Eur. J. 2008, 14, 9181. 56 a) A. Fürstner, P. W. Davies, Angew. Chem., Int. Ed. 2007,

Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan | 675
46, 3410. b) A. Fürstner, Acc. Chem. Res. 2014, 47, 925. d) J. M. Takacs, P. W. Newsome, C. Kuehn, F. Takusagawa,
57 A. Fürstner, E. Kattnig, O. Lepage, J. Am. Chem. Soc. Tetrahedron 1990, 46, 5507.
2006, 128, 9184. 84 P. A. Wender, M. P. Croatt, N. M. Deschamps, in
58 See also: T. Kang, S. B. Song, W.-Y. Kim, B. G. Kim, H.-Y. Comprehensive Organometallic Chemistry III, (I. Ojima, Ed.),
Lee, J. Am. Chem. Soc. 2014, 136, 10274. Elsevier, 2007, Vol. 10, pp. 603­648.
59 a) M. Suda, A. Fukushima, Tetrahedron Lett. 1981, 22, 759. 85 K. Jonas, L. Schieferstein, C. Krüger, Y.-H. Tsay, Angew.
b) M. Lakhrissi, Y. Chapleur, J. Org. Chem. 1994, 59, 5752. Chem., Int. Ed. Engl. 1979, 18, 550.
60 J. S. Yadav, V. Prahlad, M. C. Chander, J. Chem. Soc., 86 A. Fürstner, K. Majima, R. Martin, H. Krause, E. Kattnig,
Chem. Commun. 1993, 137. R. Goddard, C. W. Lehmann, J. Am. Chem. Soc. 2008, 130, 1992.
61 K. Lehr, R. Mariz, L. Leseurre, B. Gabor, A. Fürstner, 87 A. Fürstner, R. Martin, K. Majima, J. Am. Chem. Soc. 2005,
Angew. Chem., Int. Ed. 2011, 50, 11373. 127, 12236.
62 K. Lehr, S. Schulthoff, Y. Ueda, R. Mariz, L. Leseurre, B. 88 For a leading review on an alternative ligand design that
Gabor, A. Fürstner, Chem.®Eur. J. 2015, 21, 219. enables yet other valuable iron catalyzed cycloaddition reactions,
63 a) M. Topolski, M. Duraisamy, J. Rachon, J. Gawronski, K. see: P. J. Chirik, Angew. Chem., Int. Ed. 2017, 56, 5170.
Gawronska, V. Goedken, H. M. Walborsky, J. Org. Chem. 1993, 89 A. Casitas, H. Krause, S. Lutz, R. Goddard, E. Bill, A.
58, 546. b) H. Yanagisawa, K. Miura, M. Kitamura, K. Narasaka, Fürstner, Organometallics 2018, 37, 729.
K. Ando, Bull. Chem. Soc. Jpn. 2003, 76, 2009. 90 A. Fürstner, R. Martin, H. Krause, G. Seidel, R. Goddard,
64 E. J. Corey, P. L. Fuchs, Tetrahedron Lett. 1972, 13, 3769. C. W. Lehmann, J. Am. Chem. Soc. 2008, 130, 8773.
65 L. Hoffmeister, P. Persich, A. Fürstner, Chem.®Eur. J. 91 J. P. Collman, Acc. Chem. Res. 1975, 8, 342.
2014, 20, 4396. 92 B. Bogdanovic, M. Schwickardi, Angew. Chem., Int. Ed.
66 S. M. Rummelt, J. Preindl, H. Sommer, A. Fürstner, Angew. 2000, 39, 4610.
Chem., Int. Ed. 2015, 54, 6241. 93 M. L. Neidig, S. H. Carpenter, D. J. Curran, J. C. DeMuth,
67 a) G. Mata, B. Wölfl, A. Fürstner, Chem.®Eur. J. 2019, 25, V. E. Fleichauer, T. E. Iannuzzi, P. G. N. Neate, J. D. Sears, N. J.
246. b) B. Wölfl, G. Mata, A. Fürstner, Chem.®Eur. J. 2019, 25, Wolford, Acc. Chem. Res. 2019, 52, 140.
255. 94 A. Casitas, H. Krause, R. Goddard, A. Fürstner, Angew.
68 M. Fuchs, A. Fürstner, Angew. Chem., Int. Ed. 2015, 54, Chem., Int. Ed. 2015, 54, 1521.
3978. 95 A. Casitas, J. A. Rees, R. Goddard, E. Bill, S. DeBeer, A.
69 A. Fürstner, Angew. Chem., Int. Ed. 2013, 52, 2794. Fürstner, Angew. Chem., Int. Ed. 2017, 56, 10108.
70 A. Fürstner, Science 2013, 341, 1229713. 96 a) P.-G. Echeverria, A. Fürstner, Angew. Chem., Int. Ed.
71 C.-L. Sun, A. Fürstner, Angew. Chem., Int. Ed. 2013, 52, 2016, 55, 11188. b) F. Gomes, P.-G. Echeverria, A. Fürstner,
13071. Chem.®Eur. J. 2018, 24, 16814.
72 P. T. Northcote, J. W. Blunt, M. H. G. Munro, Tetrahedron 97 A. Fürstner, H. Krause, C. W. Lehmann, Angew. Chem., Int.
Lett. 1991, 32, 6411. Ed. 2006, 45, 440.
73 D. Romo, N. S. Choi, S. Li, I. Buchler, Z. Shi, J. O. Liu, 98 R. B. Bedford, P. B. Brenner, E. Carter, P. M. Cogswell,
J. Am. Chem. Soc. 2004, 126, 10582. M. F. Haddow, J. N. Harvey, D. M. Murphy, J. Nunn, C. H.
74 L. Shen, J. Pelletier, Nat. Prod. Rep. 2020, 37, 609. Wondall, Angew. Chem., Int. Ed. 2014, 53, 1804.
75 A. Fürstner, Angew. Chem., Int. Ed. 2018, 57, 4215. 99 a) L. Rousseau, C. Herrero, M. Clémancey, A. Imberdis, G.
76 a) W. Chaladaj, M. Corbet, A. Fürstner, Angew. Chem., Int. Blondin, G. Lefèvre, Chem.®Eur. J. 2020, 26, 2417. b) M.
Ed. 2012, 51, 6929. b) J. Preindl, K. Jouvin, D. Laurich, G. Seidel, Clémancey, T. Cantat, G. Blondin, J.-M. Latour, P. Dorlet, G.
A. Fürstner, Chem.®Eur. J. 2016, 22, 237. c) J. Preindl, S. Lefèvre, Inorg. Chem. 2017, 56, 3834.
Schulthoff, C. Wirtz, J. Lingnau, A. Fürstner, Angew. Chem., Int. 100 See also: a) T. Parchomyk, K. Koszinowski, Chem.®Eur. J.
Ed. 2017, 56, 7525. 2018, 24, 16342. b) T. Parchomyk, S. Demeshko, F. Meyer, K.
77 C.-X. Zhuo, A. Fürstner, J. Am. Chem. Soc. 2018, 140, Koszinowski, J. Am. Chem. Soc. 2018, 140, 9709.
10514. 101 S. B. Munoz, S. L. Daifuku, W. W. Brennessel, M. L.
78 C.-X. Zhuo, A. Fürstner, Angew. Chem., Int. Ed. 2016, 55, Neidig, J. Am. Chem. Soc. 2016, 138, 7492.
6051. 102 T. Parchomyk, K. Koszinowski, Chem.®Eur. J. 2017, 23,
79 a) K. Fukuhara, H. Urabe, Tetrahedron Lett. 2005, 46, 603. 3213.
b) S. Okada, K. Arayama, R. Murayama, T. Ishizuka, K. Hara, N. 103 J. D. Sears, S. B. Munoz, S. L. Daifuku, A. A. Shaps, S. H.
Hirone, T. Hata, H. Urabe, Angew. Chem., Int. Ed. 2008, 47, 6860. Carpenter, W. W. Brennessel, M. L. Neidig, Angew. Chem., Int. Ed.
c) G. Sugano, K. Kawada, M. Shigeta, T. Hata, H. Urabe, 2019, 58, 2769.
Tetrahedron Lett. 2019, 60, 885. 104 a) E. Nakamura, N. Yoshikai, J. Org. Chem. 2010, 75,
80 L. Huang, Y. Gu, A. Fürstner, Chem.®Asian J. 2019, 14, 6061. b) N. Yoshikai, Isr. J. Chem. 2017, 57, 1117.
4017. 105 M. C. White, Science 2012, 335, 807.
81 H. Kurosawa, A. Yamamoto, ed., Fundamentals of Molec- 106 a) P. Gandeepan, T. Müller, D. Zell, G. Cera, S. Warratz, L.
ular Catalysis, Elsevier, 2003. Ackermann, Chem. Rev. 2019, 119, 2192. b) C.-L. Sun, B.-J. Li,
82 M. F. Semmelhack, in Organometallics in Synthesis. A Z. J. Shi, Chem. Rev. 2011, 111, 1293.
Manual (M. Schlosser, Ed.), 2nd Ed.; Wiley: Chichester, 2002; 107 S. H. Carpenter, N. L. Neidig, Isr. J. Chem. 2017, 57, 1106.
pp. 1003. 108 a) P. L. Holland, Acc. Chem. Res. 2015, 48, 1696. b) P. L.
83 a) J. P. Genet, J. Ficini, Tetrahedron Lett. 1979, 20, 1499. Holland, Acc. Chem. Res. 2008, 41, 905.
b) H. tom Dieck, R. Diercks, Angew. Chem., Int. Ed. Engl. 1983, 109 H. Nagashima, Bull. Chem. Soc. Jpn. 2017, 90, 761.
22, 778. c) K.-U. Baldenius, H. tom Dieck, W. A. König, D. 110 For an instructive case in cycloaddition chemistry, see: L.
Icheln, T. Runge, Angew. Chem., Int. Ed. Engl. 1992, 31, 305. Hu, H. Chen, J. Am. Chem. Soc. 2017, 139, 15564.

676 | Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan
111 R. R. Arevalo, P. J. Chirik, J. Am. Chem. Soc. 2019, 141, 113 P. J. Chirik, K. Wieghardt, Science 2010, 327, 794.
9106. 114 F. Neese, Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012,
112 A. Quintard, J. Rodriquez, Angew. Chem., Int. Ed. 2014, 2, 73.
53, 4044.

Bull. Chem. Soc. Jpn. 2021, 94, 666–677 | doi:10.1246/bcsj.20200319 © 2021 The Chemical Society of Japan | 677

You might also like