1
Introduction
Despite
being largely absent from natural products and biological
processes, fluorine plays a conspicuous and increasingly important
role within pharmaceuticals and agrochemicals, as well as in materials
science.
1a−1c
Indeed, as many as 35% of agrochemicals and 20% of
pharmaceuticals on the market contain fluorine.
1d
Fluorine is the most electronegative element in the periodic
table, and the introduction of one or more fluorine atoms into a molecule
can result in greatly perturbed properties. Fluorine substituents
can potentially impact a number of variables, such as the acidity
or basicity of neighboring groups, dipole moment, and properties such
as lipophilicity, metabolic stability, and bioavailability. The multitude
of effects that can arise from the introduction of fluorine in small
molecules in the context of medicinal chemistry has been extensively
discussed elsewhere.
2
For these reasons,
methods to introduce fluorine into small organic
molecules have been actively investigated for many years by specialists
in the field of fluorine chemistry. However, particularly in the past
decade, a combination of the increasing importance of fluorine-containing
molecules and the successful development of bench stable, commercially
available fluorine sources has brought the expansion of fluorine chemistry
into the mainstream organic synthesis community. This has resulted
in an acceleration in the development of new fluorination methods
and consequently in methods for the asymmetric introduction of fluorine.
3
Catalytic asymmetric fluorination methods have
inevitably lagged somewhat behind their nonasymmetric counterparts
as understanding of the modes of reactivity of new fluorinating reagents
must generally be developed and understood before they can be extended
to enantioselective catalysis.
3b
Indeed,
the last special issue of Chemical Reviews dedicated
to fluorine chemistry, in 1996, contained no articles addressing asymmetric
fluorine chemistry, and the editor of the issue noted that “although
fluorine chemistry is much less abstruse now than when I entered the
field a generation ago, it remains a specialized topic and most chemists
are unfamiliar, or at least uncomfortable, with the synthesis and
behavior of organofluorine compounds.”
4
The field has undoubtedly undergone great change within the last
two decades.
As with the incorporation of the fluorine atom,
the introduction
of the trifluoromethyl (CF3) group into organic molecules
can substantially alter their properties. As with fluorine, the prevalence
of CF3 groups in pharmaceuticals and agrochemicals coupled
with the development of new trifluoromethylating reagents also has
led to a recent surge in the development of asymmetric trifluoromethylation
and perfluoroalkylation. Although the fluorine and trifluoromethyl
moieties are often found on the aromatic rings of many pharmaceutical
and agrochemicals rather than in aliphatic regions, this may be a
result of the lack of efficient methods for the asymmetric introduction
of C–F and C–CF3 bonds into molecules; it
could be the case that lack of chemical methods is restricting useful
exploration of such molecules. However, there are still encouraging
examples of drug candidates containing chiral fluorine and trifluoromethyl-bearing
carbons (Figure 1).
Figure 1
Molecules of medicinal
interest bearing C–F and C–CF3 stereocenters.
The asymmetric synthesis of fluorine-containing
organic compounds
using catalytic methods is of particular topical interest and is a
vibrant area of chemical research. Since the beginning of the 21st
century, much progress has been made in this field with significant
advances being made since the preceding Chemical Reviews article in 2008 by Ma and
Cahard.
3a
In
this Review, we aim to comprehensively cover advances in catalytic
enantioselective fluorination, trifluoromethylation, and perfluoralkylation
reactions up to May 2014. Additionally, we will also include sections
covering the introduction of mono- and difluoromethyl groups as well
as trifluoromethylthiolation. In contrast to the 2008 review, due
to the desire to focus on catalytic asymmetric processes, we will
not include diastereoselective processes (with the arguable exception
of tandem processes), noncatalytic reactions, or asymmetric functionalization
of fluorine-containing compounds, unless these can explicitly lead
to a mono- or difluoromethyl group.
2
Catalytic
Enantioselective Fluorination
2.1
Electrophilic Fluorination
Electrophilic
fluorination reactions with highly oxidizing fluorinating reagents
such as fluorine gas, hypofluorites, and fluoroxysulfates can be challenging
to perform without special equipment and precautions, due to their
high reactivity. This largely precluded the development of catalytic
asymmetric methods until the development in the 1990s of bench-stable,
easily handled electrophilic fluorinating reagents such as N-fluorobenzene-sulfonimide
(NFSI), the family of N-fluoropyridinium salts, and 1-chloromethyl-4-fluoro-1,4-diazoniabicyclo-[2.2.2]octane
bis(tetrafluoroborates) (Selectfluor). It is the development of these
“tamed” electrophilic reagents that allowed researchers
to develop catalytic enantioselective fluorination of nucleophilic
substrates (Figure 2).
Figure 2
A selection of commonly
used electrophilic fluorinating reagents.
2.1.1
Metal-Catalyzed Fluorination Involving Enolates
The earliest advances in catalytic asymmetric fluorination were
made by exploiting transition metal enolates, capable of a bidentate
mode of coordination to a metal. This approach provided activation
of the substrate through enolate formation, together with the ability
to impose a rigid chiral environment by virtue of chiral ligands bound
to the transition metal.
2.1.1.1
Ti/TADDOL Catalysts
The first
such reaction was developed by Hintermann and Togni in 2000.
5a
The authors reasoned that a catalytic amount
of a Lewis acid would accelerate fluorination of β-ketoesters
by catalyzing the enolization process. Fluorination of acyclic β-ketoesters 1 with
Selectfluor and Ti(TADDOLato) catalyst 2 in MeCN at room temperature afforded the
desired products 3 in high yield (>80%) and up to 90% ee (Scheme 1). The authors have
recently advanced a steric model
explaining
the facial selectivity of the fluorination, which was arrived at by
combining X-ray data with molecular modeling (Figure 3).
5b,5c
While only a few examples in
the original disclosure achieved the highest levels of enantioselectivity,
this was an extremely influential study and set the stage for much
work that followed.
Scheme 1
Figure 3
Proposed steric model explaining the facial selectivity
in the
titanium-TADDOLate-catalyzed fluorination. Reproduced with permission
from ref (5b). Copyright
2011 Beilstein-Institut zur Foerderung der Chemischen Wissenschaften.
In 2003, Togni and co-workers
applied the same catalytic system
to the one-pot enantioselective heterodihalogenation of β-ketoesters
with Selectfluor and NCS to afford α-chloro-α-fluoro-β-ketoesters
(5, 6) in moderate to good yield (Scheme 2).
6
The sequence of addition
of the halogenating agents was found to determine the sense of asymmetric
induction, although the enantioselectivities were moderate.
Scheme 2
In 2006, Togni and co-workers applied the Ti(TADDOLato)
catalyst 2 to the asymmetric fluorination of α-acyl
lactams 7 (Table 1).
7
They found that using NFSI as fluorinating reagent
gave superior
enantioselectivity to Selectfluor; however, only one substrate (R1 = Me, R2 = Ph)
afforded high enantioselectivity
(75% yield, 87% ee).
Table 1
Asymmetric Fluorination
of α-Acyl
Lactams
R1
R2
yield, %
ee, %
Me
CH2Ph
75
26
Ph
CH2Ph
78
15
Ph
Me
75
6
Me
Ph
75
87
Cy
Me
60
50
Me
Cy
nd
46
t-Bu
Me
40
20
2.1.1.2
Metal/BINAP Catalysts
In 2002,
Sodeoka and co-workers reported the enantioselective fluorination
of β-ketoesters catalyzed by a chiral palladium complex (Scheme 3).
8
The reaction was
carried out with NFSI and 2.5 mol % of cationic palladium catalyst
(12b and 12c) to afford fluorinated β-ketoesters 10 (both cyclic and acyclic) in high
enantioselectivity. In
2003, the same group reported the use of an ionic liquid immobilized
palladium complex as catalyst for the same reaction. Similarly high
enantioselectivity was achieved, and the immobilized catalyst could
be reused up to 10 times with levels of efficiency comparable to those
in conventional organic solvents.
9
Scheme 3
The proposed mechanism involves the β-ketoester
coordinated
in a bidentate manner to the cationic palladium catalyst. This coordination
increases the acidity of the α-proton, allowing a nucleophilic
metal enolate to be easily generated and to react with the NFSI (Figure 4). The enantioselectivity
was rationalized by considering
the structure of the square-planar chiral Pd enolate complex, which
was proposed to arrange itself to minimize steric interactions between
the ligand aryl groups and the ester tert-butyl group.
This results in the Si face being shielded by the
ester tert-butyl group, requiring the NFSI to approach
from the less hindered Re face (Figure 5).
10
Figure 4
Proposed catalytic cycle
for enantioselective Pd-catalyzed fluorination
of β-ketoesters.
Figure 5
Proposed structure of the Pd–enolate complex to account
for enantioselectivity. Reproduced with permission from ref (10). Copyright 2006 Elsevier.
In 2005, the same group reported
the enantioselective fluorination
of oxindoles using a similar approach (Table 2).
11
Treatment of 3-substituted oxindoles 13 with NFSI and 2.5 mol % palladium catalyst
12b in 2-propanol gave the fluorinated products 14 in good
yield with high to excellent enantioselectivities (75–96% ee).
The method was applied to the synthesis of BMS 204352 (MaxiPost),
a promising agent for the treatment of stroke. They also demonstrated
the asymmetric fluorination of unsubstituted oxindoles; by changing
the solvent to ClCH2CH2Cl/MeOH, 16 was obtained with high enantioselectivity.
Table 2
Enantioselective
Fluorination of Oxindoles
Using Palladium Catalysis
R1
R2
temp, °C
yield, %
ee, %
Ph
H
0
96
90
p-MeC6H4
H
rt
97
86
p-MeC6H4
H
0
92
88
p-FC6H4
H
rt
94
84
o-MeC6H4
CF3
rt
80
75
Me
H
rt
86
95
Me
H
0
85
96
Et
H
rt
85
92
CH2COCH3
H
rt
85
86
Bn
H
rt
72
80
i-Bu
H
rt
85
75
In 2005, Kim and co-workers
reported the enantioselective fluorination
of α-cyano esters, catalyzed by cationic palladium complex 19a (Scheme 4).
12
Treatment of substrates 17 with NFSI as fluorine
source under mild conditions afforded the α-cyano α-fluoro
esters 18 in high yields with excellent enantiomeric
excesses (85–99% ee).
Scheme 4
Soon after, Kim
13a
and Sodoeka
10,14
reported the
catalytic fluorination of β-ketophosphonates
catalyzed by chiral palladium complexes, with high enantioselectivity
for both cyclic and acyclic β-ketophosphonate substrates 20 (Scheme 5). Kim and co-workers
later
reported this reaction in ionic liquids, with the aim of simplified
product isolation and catalyst recycling.
13b
Scheme 5
Kim and co-workers also developed enantioselective fluorination
of α-chloro-β-keto phosphonates 22 catalyzed
by chiral palladium complex 19c, which gave the corresponding
α-chloro-fluoro-β-keto phosphonates 23 with
excellent enantioselectivity (up to 95% ee) (Scheme 6).
15
Scheme 6
In 2007, Kim and co-workers
reported employing the chiral Pd(II)
complex 19a in the catalytic enantioselective α-fluorination
of α-chloro-β-ketoesters 24 with moderate
enantioselectivity (Scheme 7).
16
Scheme 7
In 2007, Kim
17
and
Sodeoka
18
simultaneously reported the enantioselective
fluorination of α-aryl-α-cyano-phosphonates 26 (Scheme 8). Because of the lower reactivity
of this substrate class, an organic base was required to accelerate
abstraction of the acidic proton. Kim selected 2,6-di-tert-butyl-4-methylpyridine
(DTBMP) as base with 81–91% ee; Sodeoka
selected 2,6-lutidine as base with 24–78% ee. In both cases,
the aryl substituent was required; aliphatic substrates gave no reaction.
Scheme 8
In 2007, Sodeoka and co-workers reported an efficient
enantioselective
fluorination of tert-butoxycarbonyl lactones and
lactams 28 with excellent enantioselectivities (94–99%
ee) (Scheme 9).
19
In the case of the less acidic lactam substrates, 2,6-lutidine (0.5
equiv) was required to function as a non-nucleophilic, weak base.
Scheme 9
In an important advance demonstrating that this strategy
could
be applied to less acidic substrates, Sodeoka and co-workers in 2007
reported the NiCl2–BINAP-catalyzed direct asymmetric
fluorination of α-aryl acetic acid derivatives 30 (Table 3 and Figure 6).
20
Among the acid derivatives tested,
thiazolidin-2-one was optimal. In addition, 2,6-lutidine and triethylsilyl
triflate were essential for efficient asymmetric fluorination. Triethylsilyl
triflate was proposed to play an important role both in generating
a dicationic nickel triflate complex and in rendering NSFI more reactive.
2,6-Lutidine was assumed to promote enolization of the substrate.
Excellent yields and good enantioselectivities (up to 88% ee) were
achieved with α-arylacetic acid derivatives; however, the alkyl
derivative gave both poor reactivity and enantioselectivity.
Table 3
Asymmetric Fluorination of α-Aryl
Acetic Acid Derivatives Using Nickel Phosphine Complex 32
R
X
cat, mol %
yield, %
ee, %
Ph
S
5
99
88
p-FC6H4
S
5
90
83
p-MeOC6H4
S
5
92
81
m-MeOC6H4
S
10
95
82
o-MeOC6H4
S
10
87
78
2-naphthyl
S
10
99
83
1-naphthyl
S
5
94
87
Ph
O
10
95
87
n-propyl
S
10
15
11
Figure 6
Proposed model to account for sense of
asymmetric induction using
catalyst 32. Reproduced with permission from ref (20). Copyright 2007 John Wiley
and Sons.
In 2012, Sodeoka and
co-workers reported the catalytic enantioselective
monofluorination of α-ketoesters 33 using a chiral
palladium μ-hydroxo complex 12d (Table 4).
21
As the resulting
monofluorinated α-ketoesters spontaneously converted into the
hydrate during purification, in situ reduction with lithium tri-sec-butylborohydride
(L-Selectride) was carried out to give
products 34. Aryl-substituted substrates all gave excellent
enantioselectivities (up to 95% ee), and substrates with a longer
alkyl chain and a benzyloxy-substituted compound gave the desired
fluorinated product with reduced but acceptable enantioselectivity
(83% ee). A stereochemical model related to that proposed earlier
for β-ketoesters was invoked to rationalize the sense of enantioinduction.
Table 4
Asymmetric Monofluorination of α-Ketoesters
R
yield, %
ee, %
syn:anti
Ph
89
94
6.9:1
p-MeC6H4
83
91
4.6:1
p-MeOC6H4
75
94
4.7:1
p-FC6H4
68
95
6.5:1
p-ClC6H4
69
95
7.6:1
PhCH2
65
83
4.2:1
BnOCH2
66
83
1.4:1
2.1.1.3
Metal/Bis(oxazoline)
Catalysts
In 2004, Ma and Cahard reported the catalytic enantioselective
fluorination
of both cyclic and acyclic β-ketoesters employing a chiral bis(oxazoline)–copper
complex derived from 37 and Cu(OTf)2 (Scheme 10).
22
They found the
use of HFIP (1,1,1,3,3,3-hexafluoroisopropanol) as additive was crucial
for achieving high enantioselectivity, although only one substrate
ultimately gave satisfactory enantiomeric excess (>80% ee).
Scheme 10
In the same year, Shibata and co-workers reported fluorination
of similar substrates using nickel and copper complexes of 37 (Scheme 11), with the
use of MTBE as solvent
leading to higher enantioselectivies.
23
An intriguing outcome of this report is that the (S,S)-bis(oxazoline)-Ph-Cu(II) complex
provided the
fluorination product 10c with opposite configuration
to that obtained by the use of (S,S)-bis(oxazoline)-Ph-Ni(II), both with high enantioselectivity.
While
use of different solvents with the different metal catalysts was required
for optimal enantioselectivity, control experiments using the same
solvent for both demonstrated that the solvent choice was not responsible
for the switch. The authors speculate that this could be a consequence
of a change in metal center geometry and provide plausible transition
states to support this hypothesis.
Scheme 11
In 2005, Shibata and co-workers reported enantioselective
chlorination
and fluorination of carbonyl compounds capable of two-point binding
(Scheme 12).
24
As
the previously reported fluorination of β-ketoesters catalyzed
by Cu(II) and Ni(II) complexes 37 still left room for
improvement, they turned to the DBFOX-Ph ligand (40),
which was highly effective in other reactions. A catalyst obtained
from DBFOX-Ph and Ni(ClO4)2·6H2O gave fluorinated compounds 39 with extremely high
levels of enantioselectivity. The reaction scope was demonstrated
on cyclic β-ketoesters (93–99% ee), acyclic β-ketoesters
(83% ee), and 3-substituted oxindoles (93–96% ee). The methodology
was showcased with a catalytic enantioselective preparation of Maxipost.
They proposed an octahedral nickel complex as key intermediate, ligated
by both substrate and ligand, and put forward a model to rationalize
the absolute stereochemistry observed leading to compound 39a (Figure 7).
Scheme 12
Figure 7
Proposed model to account for fluorination leading to 39a. Reproduced with permission
from ref (24). Copyright 2005 John Wiley and Sons.
In 2008, the same authors reported the enantioselective
fluorination
of malonates catalyzed by a DBFOX-Ph/Zn(OAc)2 complex (Table 5).
25
In contrast to β-ketoesters,
malonates 41 are relatively symmetrical, being differentiated
by sterics only. They are also generally less acidic than β-ketoesters;
however, reflux of the substrates with NFSI, Zn(OAc)2,
and DBFOX-Ph in CH2Cl2 afforded the fluorinated
malonates 42 with excellent levels of enantioselectivity.
The scope of the reaction was broad, with a wide range of functional
groups such as alkyl, aryl, oxygen, sulfur, and amino substitution
being tolerated. Ni(ClO4)2·6H2O could be used in place of Zn(OAc)2, but generally
the
latter gave superior results.
Table 5
Enantioselective
Fluorination of Malonates
entry
R
yield, %
ee, %
a
CH2Ph
90
98
b
Et
94
96
c
Me
90
99
d
Bu
93
99
e
Ph
95
99
f
OPh
85
98
g
SPh
81
90
h
NPht
91
93
i
NPht(4-Br)
93
97
In 2008, Shibata and co-workers reported the enantioselective
fluorination
of 3-(2-arylacetyl)-2-thiazolidinones 43 with NFSI employing
DBFOX-Ph/metal complexes as catalysts (Table 6).
26a
The best results were obtained using
DBFOX-Ph (11 mol %), Ni(ClO4)2·6H2O (10 mol %), and 2,6-lutidine (1.0 equiv) in CH2Cl2
at 0 °C. Later in 2009, they reported modified conditions
in which HFIP (0.3 equiv) activated the DBFOX-Ph/Ni(II) catalyst system,
permitting lower reaction temperatures.
26b
With the help of catalytic HFIP, the reaction proceeded smoothly
at −60 °C to afford the desired fluorinated product in
high yield with improved enantioselectivity.
Table 6
Enantioselective
Fluorination of α-Aryl
Acetic Acid Derivatives Using a Nickel–DBFOX Complex
entry
R
yield, %
ee, %
a
Ph
91
98
b
C6H4-m-OMe
93
96
c
C6H4-p-OMe
93
96
d
C6H4-m-Me
93
98
e
C6H4-p-Me
90
96
f
C6H4-p-CF3
94
94
g
C6H4-p-F
90
94
h
C6H4-p-Br
96
93
i
1-naphthyl
87
92
j
2-naphthyl
94
95
As a further
demonstration of the utility of this catalytic system,
3-butenoyl derivatives 45 were tested under these conditions,
which gave the desired fluorinated products 46 in high
yield and with good enantioselectivity (78–91% ee) (Scheme 13).
Scheme 13
In 2007, Iwasa and co-workers
reported the design and synthesis
of a series of hybrid chiral oxazoline ligands 49, incorporating
an axially chiral binaphthyl unit. The Ni(II) complex of this tridentate
ligand catalyzed the enantioselective fluorination of β-ketoesters
with excellent yields and enantioselectivities (94% ee) (Scheme 14).
27a
Later, they found
the enantiomeric excess of this reaction could be dramatically improved
to 99% ee when a solution of NFSI was slowly added to the mixture
of substrate and catalyst.
27b
A strong
matched–mismatched effect was observed, arising from the two
sources of chirality contained within the ligand backbone.
Scheme 14
In 2011, Gade and co-workers described the synthesis of
a new class
of chiral tridentate N-donor pincer ligands, bis(oxazolinyl-methyldiene)isoindolines
52. These ligands were subsequently applied in the Ni(II)-catalyzed
enantioselective fluorination of oxindoles 50 and β-ketoesters 47 to afford the corresponding
products with enantioselectivities
of up to 99% ee (Scheme 15).
28
Scheme 15
In 2013, Kesavan and co-workers reported the
enantioselective fluorination
of aliphatic cyclic and acyclic β-ketoesters in excellent yield
with moderate enantioselectivities using tartrate derived bidentate
bisoxazoline–Cu(II) complex 53 (Scheme 16).
29
In this case, the
bisoxazoline forms a five-membered chelate with the metal.
Scheme 16
In 2011, Shibatomi and co-workers reported the asymmetric
α-fluorination
of α-chloro-β-ketoesters using Cu(OTf)2 and
chiral spiro-oxazoline ligand 54, which was developed
into a one-pot tandem chlorination–fluorination transformation,
leading to products 6.
30a
This
could also be extended to asymmetric gem-chlorofluorination
of β-ketophosphonates. In both cases, high enantioselectivities
could be achieved (Scheme 17).
Scheme 17
In 2013, the same authors reported the highly enantioselective
fluorination of α-alkyl-β-ketoesters 9 using
Cu(OTf)2 and the same catalyst system (Scheme 18).
30b
The fluorination
proceeded in a highly enantioselective manner both when cyclic and
acyclic substrates were applied. Fluorination of α-alkyl-malonate 55 was also performed
to afford the corresponding product 56 in good enantioselectivity (83% ee).
Scheme 18
2.1.1.4
Miscellaneous Catalysts
In 2006,
Inanaga and co-workers reported the synthesis of a novel chiral rare
earth perfluorinated binaphthyl phosphate, Sc[(R)-F8BNP]3 (57), and its application
to
the α-fluorination of cyclic and acyclic β-ketoesters
with NFPY-OTf (Scheme 19).
31
This report is distinct as being a rare report of successful
catalytic asymmetric fluorination using an N-fluoropyridinium
salt as an electrophilic fluorine source.
Scheme 19
In 2007, Shibata and co-workers reported the enantioselective
electrophilic
fluorination of β-ketoesters using Selectfluor as fluorine source,
and catalyzed by DNA and an achiral copper–bipyridine complex
(Scheme 20).
32
DNA
is thought to act as a chiral scaffold, and the chirality transfer
from DNA (in this case Salmon testis DNA, st-DNA)
to the substrate appears to occur through intercalation or groove
binding of the substrate–ligand–Cu(II) complex to DNA.
The pH of the reaction needed to be carefully controlled, but impressive
enantioselectivities were achieved for such a novel and innovative
approach to asymmetric catalysis.
Scheme 20
In 2007, Togni and co-workers
reported the enantioselective electrophilic
fluorination of 1,3-dicarbonyl compounds catalyzed by ruthenium(II)
complex [RuCl(PNNP)] (58) upon activation with (Et3O)PF6 (2 equiv) (Scheme 21).
33
Oxygen donors, in particular Et2O as cosolvent, increased the activity of the catalyst
and,
in some case, the enantioselectivity.
Scheme 21
In 2009, Frings and Bolm reported the enantioselective
fluorination
of β-ketoesters catalyzed by Cu(OTf)2 and amino sulfoximine
ligands of the general structure 59 (Scheme 22).
34
Starting from both
cyclic and acyclic substrates 35, the corresponding products 36 were formed with moderate
to good enantioselectivities.
Scheme 22
In 2009, Queneau, Billard, and co-workers reported the
synthesis
of a series of new carbohydrate-substituted bipyridines 60 and their application in
the asymmetric fluorination of β-ketoesters
(Scheme 23).
35
However,
disappointingly low enantioselectivities were obtained with both cyclic
and acyclic substrates.
Scheme 23
In 2010, Kang and Kim reported
the use of chiral nickel–diamine
complexes such as 63 to catalyze the electrophilic fluorination
of α-chloro-β-ketoesters 61, which allowed
access to chiral gem-chlorofluoro products 62 with high
enantioselectivities (Scheme 24).
36
Scheme 24
In 2010, Itoh and co-workers
reported the enantioselective α-fluorination
of β-ketoesters catalyzed by Co(acac)2 and Jacobsen’s
salen ligand (R,R)-(64) (Scheme 25).
37
Both cyclic and acyclic substrates 35 gave the corresponding
products 36 in high yield with good enantioselectivities.
Scheme 25
In 2012, van Leeuwen and co-workers reported the synthesis
of a
series of new enantiopure wide-bite-angle diphosphanes 67 and their use in the palladium-catalyzed
asymmetric fluorination
of α-cyanoacetates 65 (Scheme 26).
38
However, the substrate scope
was rather limited, with only one substrate (65a) giving
greater than 90% ee.
Scheme 26
In 2012, Feng and co-workers
reported the enantioselective fluorination
of 3-substitued oxindoles 68 catalyzed by Sc(III)/N,N′-dioxide complex 70 (Scheme
27).
39
Under mild reaction conditions, a series of 3-aryl and 3-alkyl-3-fluoro-2-oxindoles
69 were obtained with high enantioselectivity. Employing this
method, MaxiPost was synthesized with 96% ee.
Scheme 27
In 2014, Peng and Du reported an efficient highly enantioselective
fluorination of β-keto esters/amides catalyzed by diphenylamine-linked
bis(thiazoline) 74–Cu(OTf)2 complex
(Scheme 28).
40
Scheme 28
Very recently in 2014, Xu, Che, and co-workers reported
the enantioselective
fluorination of β-keto esters and N-Boc oxindoles
using iron(III)–salen complexes.
41
2.1.2
Metal-Catalyzed Fluorination Not Involving
Enolates
In 2013, Gagne and co-workers reported the enantioselective
cyclization/fluorination of polyenes catalyzed by (Xylyl-phanephos)Pt2+ in combination
with XeF2, a rare example of metal-catalyzed
asymmetric fluorination not proceeding through an enolate intermediate
(Table 7).
42
The
authors noted that electrophilic Pt(II) complexes are highly effective
for initiating cation-olefin cascades and set out to investigate if
the final Pt–C bond of a cascade could be converted to a C–F
bond. Accordingly, the isolated complex [(triphos)Pt-alkyl][BF4] (75) reacted rapidly
with XeF2 to
yield the C–F containing product 76. Under the
optimized conditions, a variety of alcohol and phenol terminated dienes
and trienes 77 were converted into corresponding C3-fluorinated
bicyclic and tricyclic products 76 in moderate to good
yield with good enantioselectivities (up to 87% ee). The proposed
mechanism is as shown in Figure 8. NMR data
studies suggested that 75 is the catalyst resting state
and either undergoes β-hydride elimination to give the nonfluorinated
byproduct 78 or oxidation (with XeF2) to give
the dicationic Pt(IV) complex, which reductively eliminates to give 76.
Table 7
Platinum-Catalyzed Enantioselective
Cyclization and C3-Fluorination of Polyenes
Figure 8
Proposed mechanism for
the platinum-catalyzed cyclization/fluorination
of polyenes, adapted from ref (42).
Recently, Toste and
co-workers reported the palladium-catalyzed
three-component coupling of Selectfluor, styrenes, and boronic acids,
to provide chiral monofluorinated compounds in good yield and in high
enantiomeric excess (Scheme 29).
43
They hypothesized that the aromatic amide could
act as a directing group to control regioselectivity and to stabilize
a high-valent Pd(IV) intermediate in conjunction with the bipyridine
ligand. This would disfavor an oxidative Heck-type coupling reaction,
with the intention of favoring C–F bond formation. A chiral
ligand (pyridyl-oxazolidine ligand, 81) was used for
the asymmetric version of this reaction (81% ee); however, with the
quinolone-based directing group, the yield was low (15%). Encouragingly,
they found the simplified anilide derivatives 79 gave
improved yield without significant negative effect on the enantioselectivity.
Following these results, a range of boronic acids were tested on the
4-methoxyaniline substrate, and the desired products 80 were obtained in good yield
with high enantioselectivities with
the exception of the bulky 2,6-dimethylboronic acid (79k).
Scheme 29
A proposed mechanism for the reaction is outlined in Figure 9. The catalytic cycle
is initiated through formation
of N,N-ligated palladium(II) intermediate.
Transmetalation with the boronic acid and thereafter coordination
and insertion yield a β-arylated Pd(II) species. The coordination
of the directing group and N,N-ligand
may stabilize this intermediate and retard the competing β-hydride
elimination process. Oxidation to the high-valent Pd(IV) intermediate
is achieved using Selectfluor, and subsequent reductive elimination
yields the desired monofluorinated product and the catalyst complex.
Figure 9
Proposed
mechanism for the palladium-catalyzed three-component
coupling of Selectfluor, a styrene, and a boronic acid.
2.1.3
Organocatalytic Electrophilic
Fluorination
2.1.3.1
Tertiary Amine Catalysis
Chiral N-fluoro reagents have been investigated
for some time in
a stoichiometric sense, based on either camphorsultams
44
or other scaffolds, although the preparation
of such reagents commonly involves multiple steps and the use of challenging
fluorination methods.
45
A significant step
forward was made in 2000 when Shibata and co-workers
46a
and Cahard and co-workers
46b
independently disclosed that combining cinchona alkaloids with Selectfluor
results in N-fluoroammonium salts of the cinchona
alkaloids, which are stable and isolable. These reagents were demonstrated
as being remarkably effective for the enantioselective fluorination
of preformed enolates equivalents, such as silyl enol ethers and metal
enolates. While they are not catalytic reactions and thus will not
be discussed in detail herein, these studies laid important groundwork
for subsequent advances. Accordingly, in 2006, Shibata and co-workers
reported transition to a catalytic method for the enantioselective
fluorination of acyl enol ethers employing a cinchona alkaloid/Selectfluor
combination (Scheme 30).
47
The authors found acetyl enol ethers were not significantly
reactive toward Selectfluor at room temperature, and this allowed
formation of the reactive fluorinated cinchona alkaloid to occur.
Addition of NaOAc as base was necessary for enolate activation. While
the desired α-fluoroketones 83 were only afforded
in moderate enantioselectivities (up to 54% ee), this proved that
this approach is viable in a catalytic sense.
Scheme 30
By 2008, Shibata and co-workers modified their approach
to realize
for the first time a highly enantioselective catalytic process (Table 8).
48
They employed NFSI
as fluorinating reagent in combination with catalytic amounts of bis-cinchona
alkaloids in the presence of excess base. Allyl silanes and silyl
enol ethers 85 undergo efficient enantioselective fluorodesilylation
to provide the corresponding fluorinated compounds 86 bearing an enantioenriched fluorine-substituted
quaternary carbon
center, although the requirement for a bulky substituent on the substrate
is a limitation on the enantioselectivity of this method. NFSI was
chosen as fluorinating reagent because of its low background reaction.
The authors advance a mechanistic hypothesis in which excess base
would form an N-fluoroammonium KCO3
– salt intermediate, in which the KCO3
– triggers fluorodesilylation of the substrate and would
be followed by the enantioselective transfer of fluorine to the substrate.
Additionally, they propose a transition state for the reaction, shown
below, to account for the observed absolute stereochemistry (Figure 10).
Figure 10
Proposed transition state for fluorination of silyl enol
ether.
Reproduced with permission from ref (48). Copyright 2008 John Wiley and Sons.
Table 8
Enantioselective Fluorination of Allyl
Silanes and Silyl Enol Ethers Using Bis-cinchona Alkaloid Catalysts
X
R
n
catalyst
T (°C)
yield, %
ee, %
CH2
C6H5CH2
1
a
–40
75
94
CH2
p-Me-C6H4CH2
1
a
–20
75
95
CH2
p-Cl-C6H4CH2
1
a
–20
81
94
CH2
p-Me-C6H4CH2
1
a
–20
65
90
CH2
o-Me-C6H4CH2
1
a
–20
58
93
CH2
2-naphthylmethyl
1
a
–20
69
91
CH2
Me
1
a
–40
73
72
CH2
H
1
a
–20
58
51
CH2
C6H5CH2
2
a
–20
74
81
CH2
p-Me-C6H4CH2
2
a
–20
71
81
O
C6H5CH2
2
b
–40
82
82
O
p-Me-C6H4CH2
2
b
–40
79
86
O
p-Cl-C6H4CH2
2
b
–40
74
86
O
p-MeO-C6H4CH2
2
b
–40
84
85
O
2-naphthylmethyl
2
b
–40
88
84
O
Et
2
b
–40
95
67
To demonstrate
the further synthetic utility of this catalytic
approach, they also investigated the catalytic enantioselective fluorination
of oxindoles 13a (Scheme 31).
An ee value of 85% was able to be obtained when CsOH·H2O was used as a base at −80
°C, although this required
use of the modified catalyst (DHQD)2AQN.
Scheme 31
In 2011, Gouverneur and co-workers reported the enantioselective
fluorocyclization of prochiral indoles 87 catalyzed by
cinchona alkaloids (Table 9).
49
The best results were afforded when using a catalytic amount
of (DHQ)2PHAL at −78 °C in acetone with NFSI
and an excess of K2CO3. The yields and enantiomeric
excesses of the catalytic reaction were comparable to the corresponding
stoichiometric reactions, which were also reported, although the relatively
few results delivering the highest selectivities (>80%) demonstrate
the challenging nature of the transformation. The process installs
the fluorine substituent on a quaternary benzylic stereogenic carbon
center and leads to new fluorinated analogues of natural products
featuring the hexahydropyrrolo[2,3-b]indole or the
tetrahydro-2H-furo-[2,3-b]indole
skeleton 88.
Table 9
Enantioselective
Fluorocyclization
of Prochiral Indoles Catalyzed by Cinchona Alkaloids
entry
R1
R2
XH
yield, %
ee, %
a
H
Me
OH
72
66
b
H
Et
OH
69
52
c
H
allyl
OH
76
60
d
OMe
Me
OH
65
74
e
OBn
Me
OH
78
74
f
OEt
Me
OH
78
72
g
Oallyl
Me
OH
65
78
h
Ph
Me
OH
61
62
i
Mes
Me
OH
55
84
j
H
Me
NHTs
59
64
k
OMe
Me
NHTs
51
70
l
Ph
Me
NHTs
70
70
m
Mes
Me
NHTs
80
84
n
H
Me
NHCOMe
95
80
o
Mes
Me
NHCOMe
65
92
p
H
Me
NHCO2Me
76
74
q
H
Me
NHCO2BN
47
77
r
H
Me
NHBoc
70
78
Prochiral indole 89 was
also subjected to the optimized
fluorocyclization conditions, which afforded the difluorinated tricyclic
tetrahydrooxazolo[3,2-a] indole 90 in
50% yield, 60% ee (Scheme 32).
Scheme 32
In 2012, Tu and co-workers reported the asymmetric fluorination/semipinacol
rearrangement of 2-oxa allylic alcohols 91 to afford
the corresponding chiral β-fluoroketones 92, catalyzed
by cinchona-alkaloid derivatives (Scheme 33).
50
Moderate to good enantioselectivities
of the β-fluoroketones 92 were afforded using an
NFSI (1.2 equiv)/(DHQD)2PYR (0.2 equiv) combination with
K2CO3 as base at −10 °C. Comparable
enantioselectivities of the antipodes were afforded when (DHQ)2PYR was used as catalyst.
Scheme 33
In 2013, He and co-workers
reported using structurally modified N-fluorobenzenesulfonimides
(NFSI) in the enantioselective
fluorination of oxindoles 93 in the presence of a bis-cinchona
alkaloid, (DHQD)2PHAL, as the catalyst (Scheme 34).
51a,51b
They investigated a range of
NSFI analogues 95 and found that the analogue bearing tert-butyl groups at the para-position
of the phenyl rings led to an enhanced enantioselectivity in most
cases, although with relatively slow reaction rate and low yield.
Scheme 34
2.1.3.2
Enamine Catalysis
The rapid advances
in enantioselective enamine catalysis seen in the early 2000s were
quickly extended to fluorination. In 2005, several groups almost concurrently
reported the enantioselective fluorination of aldehydes catalyzed
by chiral secondary amines, proceeding via enamine intermediates.
Enders and Hüttl reported the organocatalytic direct α-fluorination
of aldehydes and ketones employing Selectfluor and proline-related
secondary amine 98 as catalyst (Scheme 35).
52
However, only low enantioselectivity
(36% ee) was afforded when the best conditions were applied to cyclohexanone
(96), and enantioselectivities for the aldehyde substrates
were not reported.
Scheme 35
Jørgensen and co-workers reported the application
of silylated
prolinol derivative 101 to tackle the enantioselective
α-fluorination of aldehydes. They used NFSI as fluorine source
in MTBE, with high enantioselectivities (91–97% ee) (Table 10).
53
Because of their
instability, the α-fluorinated aldehydes were reduced in situ
to afford the β-fluorinated alcohols 100 for subsequent
analysis.
Table 10
Scope of Reaction for Enantioselective
α-Fluorination of Aldehydes Using Catalyst 101
entry
R
yield, %
ee, %
a
Pr
95
96
b
Bu
>90
91
c
Hex
55
96
d
BnO(CH2)3
64
91
e
Bn
74
93
f
Cy
69
96
g
tBu
>90
97
h
1-Ad
75
96
The authors proposed a catalytic
cycle for this reaction (Figure 11), and rationalize
the stereochemical outcome by
invoking formation of the more stable E-configured
enamine, with the 3,5-di(trifluoromethyl)phenyl groups blocking the Re face. As a
consequence, the electrophilic fluorination
occurs from the Si face, with excellent stereocontrol.
Figure 11
Proposed
catalytic cycle for enantioselective α-fluorination
of aldehydes.
They also demonstrated
preliminary results on the extension of
the reaction scope to the formation of quaternary stereocenters (Scheme 36). The sterically
encumbered substrate 102 required less bulky catalyst 104 and a higher temperature,
allowing product formation with a moderate but encouraging 48% ee.
Scheme 36
At a similar time, Barbas and co-workers focused on the
enantioselective
fluorination of α-branched aldehydes 105 using
chiral secondary amine catalysts (Scheme 37).
54
After a catalyst screen in THF at
room temperature, moderate enantioselectivities of the fluorinated
products were obtained (45% ee for 106a; 66% ee for 106b). The authors found that
better results were obtained
when applying these conditions to nonbranched aldehydes. While formation
of the α,α-difluoro product was initially a problem, a
screen of the solvents revealed that DMF could inhibit formation of
this byproduct, and the desired product was formed with 88% ee when
imidazolidinone catalyst 111 (30 mol %) was employed,
although the yield was low (30%). Unfortunately, use of a stoichiometric
amount of the catalyst was required to achieve good yield.
Scheme 37
MacMillan and co-workers reported the use of imidazolidinone
dichloroacetate 112 for the fluorination of linear aldehydes 99 with NFSI to achieve
excellent enantioselectivity (up to
99% ee)
(Scheme 38).
55
The
authors found that addition of 10% i-PrOH as cosolvent
generally improved enantiocontrol and efficiency, although the origin
of this effect is not clear. A wide range of functional groups, including
olefins, esters, amine, carbamates, aryl rings, and sterically demanding
substituents, could be readily tolerated on the aldehyde substrates,
although α-branched aldehydes were not discussed. They also
evaluated the effect of the catalyst loading on reaction efficiency
and found that loadings as low as 2.5 mol % could be used without
loss of enantiocontrol.
Scheme 38
In 2006, Jørgensen and
co-workers reported using chiral 8-amino-2-naphthol
derivative 114, the product of an asymmetric Friedel–Craft
amination, as an organocatalyst in the asymmetric α-fluorination
of α-branched aldehydes.
56
This catalyst
delivered the products 113 in up to 90% ee (Scheme 39). In general, good enantioselectivities
were achieved
when R1 was an aromatic substituent, although the yields
were low to moderate due to suspected instability of the products
upon column chromatography. In the case of substrates bearing two
aliphatic substituents, the enantioselectivities of the reaction dropped
significantly, highlighting the challenging nature of these substrates.
The authors advanced a mechanistic hypothesis for asymmetric induction,
inspired by an X-ray structure of an acylated analogue of the catalyst,
which suggested that the catalyst naphthol substituents may sit at
right angles to the naphthol core and a hydrogen bond may be present
between the BOC group and enamine NH (Figure 12).
Scheme 39
Figure 12
Proposed structure of enamine intermediate showing one face blocked
and an intramolecular hydrogen bond. Reproduced with permission from
ref (56). Copyright
2006 John Wiley and Sons.
In 2005, MacMillan and co-workers reported a combination
of transfer
hydrogenation using Hantzsch ester and electrophilic fluorination
using NFSI on enal substrates, which would allow the formal asymmetric
addition of HF across trisubstituted olefin systems using a cascade-catalysis
approach (Scheme 40).
57
The authors assumed that the iminium and enamine steps might be
discretely controlled by separate catalysts, although a single catalyst
could also enable both activation cycles. Using imidazolidinone catalyst 118 (20 mol
%), the product 116 was obtained
in 60% yield with 99% ee, albeit with 3:1 anti:syn diastereoselectivity.
Implementation of catalyst combination A (118 7.5 mol
% and (S)-112 30 mol %) allowed the
formal addition of HF to the trisubstituted enal with 16:1 anti:syn
selectivity (99% ee). Remarkably, the syn HF addition
product 117 could be accessed with 9:1 selectivity and
in 99% ee by simply changing the enantiomeric series of 112 employed in this catalyst
combination (118 7.5 mol
% and (R)-112 30 mol %).
Scheme 40
In 2010, Brenner-Moyer and co-workers reported organocatalyzed
enantioselective aminofluorination of enals to afford chiral α-fluoro-β-amino
aldehydes using the Jørgensen–Hayashi diarylprolinol catalyst 121 (Table 11).
58
The consistently high diastereo- and enantioselectivities
are a testament to the power of the enamine/iminium modes of catalysis.
The presence of other olefins, ether protecting groups, and remote
reactive functional groups, such as cyano groups, was well tolerated.
Table 11
Scope of Reaction for Enantioselective
Aminofluorination of Enals
entry
R
yield, %
dr (syn:anti)
ee, %
a
Et–
73
95:5
99
b
n-Pr–
64
94:6
99
c
n-Bu–
66
95:5
99
d
PhCH2CH2–
51
95:5
98
e
(CH3)2CHCH2–
41
90:10
99
f
i-Pr–
24
98:2
80
g
CH2=CH(CH2)3–
61
93:7
99
h
PhCH2OCH2–
57
87:13
99
i
N=C(CH2)5–
41
91:9
99
In 2008, Yamamoto and Shibatomi
reported enantioselective fluorination
of α-chloro-aldehydes 122 to afford α,α-chlorofluoro
aldehydes in high enantioselectivity catalyzed by the catalyst (101) developed by
Jørgensen et al. (Scheme 41).
59
In situ reduction
of the aldehydes with NaBH4 afforded β,β-chlorofluoro
alcohols 123 in high enantioselectivity. The products
could be further elaborated to the corresponding α,α-chlorofluoro
ketones 124. Subsequent mechanistic studies suggest that
the reaction mechanism involves a kinetic resolution of the starting
α-chloro-aldehyde.
60
Scheme 41
In 2009, Jørgensen and co-workers reported an extension
to
their earlier published work: a simple, direct one-pot organocatalytic
approach to the formation of optically active propargylic fluorides 126 (Table 12).
61
This consists of organocatalytic α-fluorination of
aldehydes followed by homologation with the Ohira–Bestman reagent 125, providing optically
active propargylic fluorides 126 with enantioselectivities of up to 99% ee.
Table 12
One-Pot Organocatalytic Approach
to Enantioenriched Propargylic Fluorides
entry
R
yield, %
ee, %
a
–Bn
56
95
b
–(CH2)7CH3
67
93
c
–(CH2)13CH3
65
99
d
–(CH2)7CH=CH2
55
92
e
p-OMe-C6H4CH2–
65
91
f
p-Br-C6H4CH2–
69
92
g
o-Me-C6H4CH2–
58
99
h
o-Br-C6H4CH2–
47
94
i
–(CH2)3COOMe
45
93
The authors also extended this approach
further to the direct synthesis
of click adducts 127 from aldehydes in three steps in
one pot in 50% yield with 95% ee (Scheme 42). The Wittig reaction is also fully compatible
with the organocatalytic
asymmetric reaction. The one-pot reaction, organocatalytic formation
of α-fluoro aldehydes in combination with the commercially available
methyl (triphenylphosphoranylidene) acetate, furnished the corresponding
allylic fluorides 128 in moderate yield and high enantioselectivities.
Scheme 42
In 2009, Lindsley and co-workers reported the one-pot
catalytic
enantioselective synthesis of chiral β-fluoramines 129 via organocatalysis, again
proving the versatility of the α-fluorination
methodology (Scheme 43).
62
This was accomplished using NFSI with imidazolidinone catalyst 112 (20 mol %), at
−20 °C in 10% i-PrOH/THF, followed by direct addition of Boc-piperazine and NaBH(OAc)3.
The desired β-fluoroamine 129 was isolated
in 65% yield and with >96% ee.
Scheme 43
In 2011, MacMillan and co-workers
reported the first highly enantioselective
α-fluorination of cyclic ketones, catalyzed by a cinchona alkaloid-derived
primary amine (Scheme 44).
63a
This catalyst was identified after a high-throughput evaluation
of a library of chiral amine catalysts. Use of the dihydroquinidine
catalyst 132 with trichloroacetic acid (TCA) as cocatalyst
at −20 °C provided the α-fluoro cyclic ketones 131 in good yield with excellent enantioselectivities
(up
to 99% ee). The scope of this reaction was well examined; geminal
disubstituted cyclohexanones and a wide array of heterocycles were
well tolerated. The five-membered and seven-membered cyclic ketones
also worked well (88% and 98% ee), albeit with moderate yield. They
also successfully applied this methodology to the diastereoselective
fluorination of cyclic ketones bearing pre-existing stereocenters.
For example, in fluorination of (R)-3-methyl cycylohexanone
(130j), the diastereoselectivity of the fluorinated product
could be completely controlled by the choice of catalyst.
Scheme 44
They also employed the fluorination conditions to more
complex
substrates, including the hydrogenated Hajos–Parrish ketone 131o, allo-pregnanedione
131p, and cholestanone 131q (Scheme 45).
Scheme 45
Very recently, Lam and Houk reported a detailed computational
study
of MacMillan’s ketone fluorination, to determine the origin
of enantioselectivity.
63b
They proposed
that fluorination of the quinuclidine portion of the catalyst by NFSI
is fast and that the enantiodetermining step is intramolecular transfer
of fluorine to the enamine, the latter being formed from condensation
of the catalyst primary amine and the ketone substrate. They concluded
that there are two key determining factors in the enantiocontrol:
first the preferred chair conformation of the seven-membered ring
in the transition state and second the steric bulk of the C9-quinoline
of the catalyst giving it a strong equatorial preference (Figure 13).
Figure 13
Two views of calculated lowest energy transition structure
for
fluorination of cyclohexanone using amine catalyst 132. Reproduced with permission
from ref (63b). Copyright 2014 American Chemical Society.
Recently, Xu and co-workers reported
the enantioselective fluorination
of β-ketoesters catalyzed by a combination of chiral primary
amine catalysts (Scheme 46).
64
The desired fluorinated product 48 was afforded
in moderate enantioselectivities (39–55% ee) only when QN-NH2 and l-leucine were used
together catalysts, and
the exact function of each catalyst is unclear.
Scheme 46
2.1.3.3
Phase-Transfer Catalysis
2.1.3.3.1
Cationic Phase-Transfer Catalysis
In 2002, Kim and
Park reported the first catalytic enantioselective
fluorination of β-ketoesters employing phase-transfer catalysis
using chiral quaternary ammonium salts (Table 13).
65
They found that a bulky group at
the bridgehead nitrogen of cinchona alkaloids was crucial for high
stereoselectivity, as in catalysts 132. Several cyclic
β-ketoesters were submitted to fluorination using NFSI and 132a (10 mol %) as catalyst,
affording the desired fluorinated
products 48 in high yield with moderate enantioselectivities
(48–69% ee). The authors found that for different substrates,
tailoring of the base was required to obtain the best results. Fluorination
of an acylic substrate 133 required NaH as base, and
afforded the α-fluoro product 134 in only 40% ee.
Table 13
Enantioselective Fluorination of
β-Ketoesters by Phase-Transfer Catalysis
entry
n
R
base
yield, %
ee, %
a
1
Me
K2CO3
92
69
b
1
Et
Cs2CO3
91
63
c
2
Me
Cs2CO3
88
48
d
2
Et
CsOH
78
52
In 2010, Maruoka and co-workers
reported the highly enantioselective
fluorination of β-ketoesters catalyzed by chiral bifunctional
phase-transfer catalysts (Scheme 47).
66
Fluorination of various cyclic β-ketoesters 9 gave the desired α-fluorinated products
with excellent
enantioselectivities (up to 99% ee) under the catalysis of thiomorpholine-derived
catalyst 135. The authors found the presence of hydroxyl
groups in 135 is crucially important to obtain high enantioselectivity.
They proposed that in the transition state, the Z-enolate would be stabilized by both
the ionic interaction of the
ammonium enolate and the hydrogen bonding between the enolate oxygen
and one hydroxyl group of the catalyst (Figure 14). This would lead to approach of
the NFSI from the upper face of
the enolate, as depicted. Unfortunately, these conditions were unsuitable
for the fluorination of acyclic substrates.
Scheme 47
Figure 14
Proposed transition state structure. Reproduced with permission
from ref (66). Copyright
2010 The Royal Society of Chemistry.
In 2013, Lu and co-workers reported the enantioselective
fluorination
of indanone-derived β-ketoesters 47 catalyzed by
adamantoyl-derivatized cinchona alkaloid phase-transfer catalyst 136, affording the
α-fluorinated products 48 in high enantioselectivities (84–94% ee) (Scheme 48).
67
Scheme 48
Ma, Cahard, and co-workers have recently explored the
use of P-spiro phosphonium salts for the fluorination
of 3-substituted
benzofuran-2-(3H)-ones 137, but only
moderate enantioselectivities have been obtained thus far (Scheme 49).
68
Scheme 49
2.1.3.3.2
Anionic Phase-Transfer
Catalysis
The use of chiral cationic salts as phase-transfer
catalysts for
anionic reagents is well precedented; however, an analogous charge-inverted
strategy in which the salt of chiral anion brings an insoluble cationic
promoter into solution has been rather less explored.
69
In 2011, Toste and co-workers reported an advance in this
field, asymmetric fluorocyclization using an anionic chiral phase-transfer
catalyst.
70
Selectfluor is normally insoluble
in nonpolar media, but the authors hypothesized that lipophilic, bulky
chiral phosphate anions such as the conjugate base of acid 140 may exchange with the
tetrafluoroborate anions associated with Selectfluor
to bring the reagent into solution. The resulting chiral ion pair
could then mediate an asymmetric fluorination of substrate in solution.
Given the insolubility of Selectfluor, little background fluorination
of the substrate would be anticipated (Figure 15).
Figure 15
Proposed catalytic cycle for chiral anion phase-transfer catalysis,
leading to fluorocyclization.
Employing optimized conditions (5 mol % 140a, 1.25
equiv of Selectfluor, and 1.1 equiv of proton sponge in C6H5F at −20 °C) on electron-rich
enol ether
substrates 141 gave the desired fluorocyclization product 142 in high yield with excellent
enantio- and diastereoselectivities
(Scheme 50). Consistent with their hypothesis,
the hydrophobic alkyl chains attached to the backbone of the catalyst
proved beneficial for the phase-transfer aspect, improving enantioselectivity
from 87% with 140b to 92% with 140a.
Scheme 50
They also extended the methodology to less electron-rich
alkenes 143 (Scheme 51). Accordingly,
fluorocyclization
of dihydronaphthalene (143a–c) and
chromene (143d) substrates also gave excellent enantio-
and diastereoselectivities at room temperature.
Scheme 51
An unanticipated benefit of the phase-transfer protocol
was an
improved tolerance toward sensitive functionality. For example, when
treated with Selectfluor under homogeneous conditions, benzothiophene
substrates 145 were converted to a complex mixture of
products. However, when the chiral anion phase-transfer reaction conditions
were applied, fluorocyclization products 146 were isolated
in good yield and high optical purity (Scheme 52).
Scheme 52
In 2012, Toste and co-workers reported the asymmetric
fluorination
of enamides to access α-fluoroimines using the same chiral anion
phase-transfer catalysis strategy (Scheme 53).
71
Applying conditions similar to those
previously employed (Selectfluor, chiral phosphoric acid 140a, Na2CO3 in hexane),
enamides or ene-carbamates
gave the desired α-fluoroimine products, which were stable and
isolable. However, only the N-benzoylenamide substrates 147 gave very high enantioselectivities.
A number of cyclic
indanone and tetralone-derived enamides provided enantioenriched α-(fluoro)benzoylimines
148 bearing a variety of diversely substituted quaternary
fluorinated stereocenters. Under the phase-transfer fluorination conditions,
chloro- and bromo-substituted enamides also gave the novel β,β-chloro,fluoro
and bromo,fluoro compounds with high enantioselectivities. A beneficial
effect in enantioselectivity was observed for some substrates when
the reaction was run in the presence of 5 equiv of 3-hexanol, although
the exact role of this additive is unclear.
Scheme 53
Unsubstituted tetralone-derived enamides 149 also
delivered stable imine products that did not racemize (Scheme 54). Fluorination of
a racemic flavanone-derived
enamide 151, which has a pre-existing stereocenter, resulted
in approximately equal ratio of separable anti and syn diastereomers (152a, 152b),
both with excellent enantioselectivities, illustrating the high selectivity
of the catalytic system, regardless of the presence of an existing
adjacent stereocenter.
Scheme 54
On the basis of the experimentally
observed absolute stereochemistry,
they advanced a tentative model for interactions with the catalyst
(Figure 16). The phosphate anion is thought
to form an ion pair with the Selectfluor reagent on the phosphate
oxygen, while the phosphoryl oxygen activates the enamide through
hydrogen bonding. The absolute stereochemistry can be rationalized
by placing the aromatic ring of the tetralone in the “open”
quadrant of the catalyst.
Figure 16
Proposed model for selectivity in fluorination
of enamides using
chiral anion phase-transfer catalysis. Reproduced with permission
from ref (71). Copyright
2012 American Chemical Society.
Subsequently, Toste and co-workers developed an enantioselective
tandem oxyfluorination of enamides using a doubly axially chiral phosphoric
acid.
72
They proposed that the fluorination
of aldehyde-derived enamides 153 would generate, in the
first instance, a protonated α-fluoro-N-acyliminium
ion. This intermediate should exhibit hydrogen-bonding interactions
with the chiral phosphate anion, allowing catalyst-controlled addition
of an external oxygen nucleophile, constituting an oxyfluorination
of enamides. Optimization of the reaction revealed that only the Z-enamides gave high
diastereo- and enantioselectivities
when the novel doubly axially chiral catalyst 155 was
used. They anticipated this catalyst would generate a more rigid and
constrained pocket for the substrate, leading to higher selectivity.
Both aromatic and aliphatic substituted enamides 153 were
effective in the tandem hydroxyfluorination process (Scheme 55). In accord with the
hypothesis that catalyst
control is operative in the hydration step, product 154j was produced in high enantioselectivity,
in which only the N,O-aminal carbon is chiral. When the reaction
was run in the presence of alcohols, alcohol addition was observed
rather than hydration, the latter being thought to arise from moisture
in the Selectfluor (154k–m).
Scheme 55
They also explored substrates that would generate a chiral
quaternary
fluorine stereocenter (Scheme 56). Fluorination
of (E)-156 gave poor diastereoselectivity
but extremely high enantioselectivity for the anti-157 and low enantioselectivity
for syn-158, which was believed to be a result of double stereodifferentiation.
However, this effect was not as pronounced for (Z)-159.
Scheme 56
In 2013, Toste and co-workers
reported the catalytic asymmetric
1,4-aminofluorination of conjugated dienes using chiral anion phase-transfer
catalysis (Table 14).
73
The 6-endo-trig fluorination of diene substrate 160 would produce an allylic fluoride,
an important chemical
motif. Optimization revealed catalyst 140c to be superior.
Although far from the reactive center, substitution on the benzamide
arene exerted a strong influence on the selectivity, with tert-butyl substitution
at the para position
giving the best results.
Table 14
Asymmetric 1,4-Aminofluorination
of Conjugated Dienes Using Chiral Anion Phase-Transfer Catalysis
entry
Ar
R1
R2
X
yield, %
ee, %
dr
a
2-MeC6H4
H
H
CH2
91
96
>20:1
b
3-MeC6H4
H
H
CH2
92
92
5.9:1
c
C6H5
H
H
CH2
90
92
6.9:1
d
C6H5
H
OMe
CH2
90
93
6.9:1
e
C6H5
H
H
O
85
91
5.5:1
f
4-CF3C6H4
H
H
CH2
94
95
10:1
g
4-MeOC6H4
H
H
CH2
89
93
7.5:1
h
C6H5
nBu
H
CH2
85
94
>20:1
The fluorocyclization of less-reactive dienes 164 was
also examined (Scheme 57). When subjected to
the optimized reaction conditions using Selectfluor as the fluorine
source, dienes 164 (not tetralone-derived, as before)
reacted sluggishly, affording only racemic product. The authors hypothesized
that the reactivity of the fluorinating reagent may be increased by
attaching an electron-deficient aryl group in place of the chlorine
atom of Selectfluor. Different derivatives were accessed simply by
treating tetrafluoroborate salts 162 with XeF2. The more electron-deficient reagent
163 produced the
cyclized product 165c with the highest levels of enantioselectivity
(89% ee).
Scheme 57
In 2013, Phipps and Toste applied this chiral
anion phase-transfer
catalysis strategy to the asymmetric fluorinative dearomatization
of phenols.
74
They hypothesized that hydrogen
bonding of the phenolic hydroxyl group with the phosphoryl oxygen
of the catalyst might enable discrimination between the enantiotopic
faces of the phenol in a subsequent fluorination reaction. Accordingly,
they found that fluorination of 2,3-di- and 2,3,4-trisubstituted phenols
under optimized conditions using 5 mol % (S)-140c afforded ortho-fluorinated dearomatized
products with high enantioselectivities (Scheme 58).
Scheme 58
They also investigated substrates without substitution
at the 3-position
(Table 15). The isolated products 169 were found to be dimers resulting from [4 +
2] cycloaddition of
the chiral 2,4-cyclohexadienones produced after the initial fluorination.
They demonstrated the further transformation of the dimeric products
in the form of retro-[4 + 2]/[4 + 2] reactions with N-phenylmaleimide and cyclopentadiene
dimer to provide rapid access
to a diverse fluorinated scaffolds (170, 171) with no loss of enantioenrichment.
Table 15
Asymmetric
Fluorinative Dearomatization
of Phenols – Scope of Phenols Lacking 3-Substitution and Further
Transformation of the Products
entry
R1
R2
yield, %
ee, %
a
Bn
H
81
97
b
Ph
H
55
90
c
homoallyl
H
65
90
d
allyl
H
71
87
e
iPr
H
96
91
f
allyl
Me
51
92
g
Bn
Me
67
90
h
(CH2)3OTBS
H
77
91
A number of para-fluorinated products were also
demonstrated to be obtainable with good enantioselectivities by incorporating
the geminal 8,8′ disubstitution to increase the steric bulk.
Throughout the study, the authors noted that steric bulk on one side
of the phenol was required to achieve high selectivities. In the case
of the para-fluorination process, yields were moderate
due to competitive SEAr arene fluorination at the ortho position (Scheme 59).
Scheme 59
In 2013, Toste and co-workers reported the enantioselective
fluorination
of alkenes using chiral anion phase-transfer catalysis and employing
remote directing groups (Scheme 60).
75a
On the basis of their previous success with
amides as pendant nucleophiles, they posited that allylic amides may
be used as remote directing groups to direct alkene fluorination to
provide access to allylic fluoride products after elimination. Optimization
revealed the catalyst STRIP (176) gave superior enantioselectivity
and chemoselectivity to TRIP (140b), which may be caused
by its tighter binding pocket. An N-methylated analogue
of 174 was unreactive under the same conditions, supporting
their hypothesis that a substrate hydrogen-bond donor is required
to direct the fluorination. Investigation of the scope of compatible
amide directing groups revealed the best enantioselectivities were
observed with larger groups on the amide (174a–e). The scope of this reaction was
explored; substrates with
different ring sizes, benzamide substitution patterns, substitution
on the bicyclic core, and heterocyclic substrates were well tolerated.
Scheme 60
In addition to amides, 2-hydroxyphenyl was found to be
an effective
directing group to enable enantioselective fluorination of a number
of tethered alkenes under similar conditions (Scheme 61). As compared to related allyl
substrates that were previously
observed to undergo fluorinative dearomatization under similar conditions
(Table 15), increased substitution on the alkene
switches the chemoselectivity with fluorination at the alkene being
seen exclusively, rather than fluorination at the phenol. Both β-phenolic
tertiary and quaternary fluorides 177a–d with alkyl, aryl, or heteroaryl substituents
were obtained with
good to excellent enantioselectivities. This concept was extended
to allylic alcohols using a boronic acid as an in situ directing group.
75b
Scheme 61
In 2013, Alexakis and co-workers
reported the enantioselective
fluorination-induced Wagner–Meerwein rearrangement of strained
allylic alcohols using chiral anion phase-transfer catalysis (Scheme 62).
76
Fluorination of
substrates 178 with Selectfluor, Na3PO4 in C6H5F/hexane (1:1) at −20
°C with catalyst 180 afforded the ring expanded
products 179 with good diastereo- and enantioselectivities.
Interestingly, both the enantioselectivity and the diastereoselectivity
of the present transformation were controlled by the catalyst. Thus,
racemic reaction (using an achiral phosphoric acid) gave low diastereoselectivities.
The substrate scope encompasses both allylic cyclobutanols and allylic
cyclopropanols based on the tetralone scaffold, as well as the chromanone
scaffold. However, allylic alcohol substrates lacking the aromatic
ring gave good yields but only moderate stereoselectivities under
the optimized conditions.
Scheme 62
2.1.3.4
Miscellaneous
Catalysts
In 2012,
Sun and co-workers reported the enantioselective synthesis of β,γ-unsaturated
α-fluoroesters catalyzed by N-heterocyclic
carbenes (NHCs).
77
NHCs are well-known
for their unique capability to reverse the normal polarity of aldehydes.
The authors hypothesized that an enal bearing a leaving group in γ-position
would provide access to a chiral NHC-bound dienolate that may subsequently
react with an electrophilic fluorinating reagent and a nucleophile
to afford an enantioenriched fluorinated product (Scheme 63).
Scheme 63
Fluorination of γ-methyl-carbonate-substituted
α,β-unsaturated
enals 181 with NFSI using 183 as catalyst
afforded the desired fluorinated products 182 in good
yield with high enantioselectivities (Scheme 64). A variety of functional groups on
the substrates were well tolerated,
including ethers, halides, cyanides, alkenes, aryl aldehydes, ketones,
free alcohols esters, and silyl-protected alcohols. The presence of
a quaternary carbon atom in γ-position (181q) did
not significantly affect the reaction efficiency, and a trisubstituted
alkene could be obtained as a single E isomer. In
contrast, a substituent at the α-position (181r) significantly retarded the reaction
(<10% conversion) and led
to moderate enantioselectivity. Alkenes with alkyl substituents at
the γ position also participated smoothly in this reaction,
albeit with low E/Z ratio. The E/Z ratio could be improved by employing
bulkier alkyl groups, such as iPr and tBu.
Scheme 64
The authors also proposed a possible transition state
to account
for the control of enantioselectivity (Figure 17). In the proposed favored dienolate,
the chiral backbone of the
NHC blocks the Si face, leading to the observed enantiomer.
DFT calculations revealed that the less favored rotamer, in which
the Re face is blocked, is about 5.3 kca lmol–1 higher in energy than the favored
one due to an unfavorable
interaction between the enolate oxygen atom and a catalyst methyl
group when the system is fully conjugated.
Figure 17
Summary of DFT calculations
leading to hypothesis to account for
absolute stereochemistry.
In 2012, Niu and co-workers reported the thiourea-catalyzed
enantioselective
fluorination of β-ketoesters (Scheme 65).
78
They found that chiral bifunctional
thiourea catalyst 184 could efficiently catalyze the
fluorination of β-ketoesters 35 with NFSI with
the assistance of DMAP in MeOH at −60 °C to afford the
desired fluorinated product in high yield with good to excellent enantioselectivities.
The alkoxy group of the indanonecarboxylate derivatives had a great
influence on enantioselectivity, with Me and Bn being optimal. The
tetralone derivatives and acyclic β-ketoesters gave low enantioselectivities
in some cases.
Scheme 65
The authors also proposed a mechanism proceeding
via a dual-activation
process wherein the NFSI hydrogen bonds to the catalyst thiourea group
and the 1,3-dicarbonyl compound interacts with the basic nitrogen
of the catalyst (Figure 18).
Figure 18
Proposed intermediate
in dual-activation process. Reproduced with
permission from ref (78). Copyright 2012 John Wiley and Sons.
In 2014, Akiyama and co-workers reported enantioselective
fluorination
of β-ketoesters catalyzed by a chiral sodium phosphate derived
from acid 187 (Scheme 66).
79
The authors supposed that simultaneous formation
of the sodium enolate and sodium phosphate under basic conditions
is the key to achieving excellent selectivity. Indanone derivatives 185a–c as well
as benzofuranone derivatives 185d–f were employed in this reaction
to afford the fluorinated products in good yields with good to excellent
enantioselectivities.
Scheme 66
Very recently, Fu and co-workers
reported the asymmetric synthesis
of tertiary alkyl fluorides via a nucleophile-catalyzed α-fluorination
of ketenes (Scheme 67).
80
In this process, the planar-chiral nucleophilic catalyst 191 reacts with the ketene
189 to form a catalyst-derived
chiral enolate. This intermediate undergoes electrophilic fluorination,
and the resulting cationic intermediate is attacked by a phenoxide
additive (188), which releases the nucleophilic catalyst
and forms the product 190. The phenoxide additive needed
to be carefully chosen to ensure turnover, and mechanistic studies
ruled out an alternative pathway whereby the catalyst itself is directly
fluorinated (Figure 19). Using this approach,
a range of valuable tertiary α-fluoroesters could be accessed
from aryl alkyl ketenes, with very high levels of enantioselectivity.
Scheme 67
Figure 19
Proposed mechanism for α-fluorination of ketenes. Reproduced
with permission from ref (80). Copyright 2014 American Chemical Society.
2.1.4
Fluorination Using Multiple
Catalysts
In 2008, Lectka and co-workers reported a catalytic,
highly enantioselective
α-fluorination of acid chlorides.
81
This reaction exploits a “dual activation” strategy
in which a chiral Lewis base catalyst (a cinchona alkaloid derivative)
is combined with a transition metal-based Lewis acid cocatalyst (Pd
or Ni were particularly effective) to catalytically generate metal-coordinated,
chiral ketene enolates. These are fluorinated by NFSI at the α
position, followed by attack of the liberated dibenzenesulfonimide
anion that can be subsequently quenched with different nucleophiles
to afford fluorinated carboxylic acids, amides, esters, and even peptides
(Scheme 68).
Scheme 68
Treatment of acid chlorides 192 with NFSI,
Hünig’s
base, 10 mol % BQd catalyst (193), and trans-(Ph3P)2PdCl2 or (dppp)NiCl2, followed
by quenching with nucleophiles after 6–15
h afforded the desired α-fluorinated acid derivatives 194 in good yield with excellent
enantioselectivities. Acid
chlorides containing aromatic as well as heterocyclic substituents
proved to be compatible substrates (Table 16).
Table 16
Substrate Scope of the Fluorination
of Acid Chlorides Using a Dual Catalyst System
entry
R
catalyst
[M]
NuH
yield, %
ee/de, %
1
p-MeOPh
Ni
MeOH
83
99
2
p-MeOPh
Pd
l-NH2-Ph-OEt
68
>99
3
p-MeOPh
Pd
PhSH
67
98
4
p-MeOPh
Ni
N-Boc-l-prolinol
90
>99
5
Ph
Ni
MeOH
61
99
6
Ph
Ni
H2O
60
99
7
1-Np
Ni
MeOH
68
98
8
1-Np
Ni
N-Coc-l-Cys-OMe
80
>99
9
2-Np
Pd
MeOH
63
>99
10
2-thiophene
Pd
MeOH
69
99
11
2-(N-benzoylindolyl)
Pd
MeOH
58
94
12
2-(3-Ph-(ethylcinnamate)
Ni
MeOH
71
99
13
phthalimido-CH2
Pd
MeOH
72
>99
14
phthalimido-CH2
Pd
NH(CH2)5
79
>99
15
indo
Pd
MeOH
84
95
16
phthalimido-CH2
Pd
(+)-emetine
91
>99
In 2010, the same group reported
the quenching of the above reaction
with nucleophilic natural products to produce biologically relevant
α-fluorinated carbonyl derivatives (Scheme 69).
82
They found that the solubility
of the nucleophile used to quench the reaction proved critical. Glutathione,
morphine, and 6-aminopenicillanic displayed marginal solubility, leading
to drastically decreased yields. Some site selectivity was also observed;
for example, when p-methoxylphenylacetyl chloride
was fluorinated and quenched with taxol, the sole product (195c) resulted from acylation
at the hydroxyl shown.
Scheme 69
Although the fluorination of aryl and heteroaryl acetyl
chlorides
(192, R = aromatic) using the above approach proved to
be very successful, the related aliphatic substrates (196) worked poorly, delivering
low yields of fluorinated products. Following
initial mechanistic studies, Lectka and co-workers speculated that
addition of a second Lewis acid may coordinate NFSI selectively, thereby
increasing its electrophilicity and reactivity (Scheme 70).
83
After optimization, the authors
found that addition of 10 mol % LiClO4 combined with the
slow addition of Hünig’s base over 12 h could increase
the yield of the aliphatic substituted products 197 while
still delivering excellent enantioselectivities (>99% ee).
Scheme 70
After a subsequent in-depth mechanistic analysis, which
took into
account the presence of three catalysts, Lewis acid activation of
the fluorinating agent by a lithium cation was deduced to be crucial
to the efficient reaction (Figure 20).
Figure 20
Summary of
the key mechanistic aspects involved in enabling the
enantioselective fluorination of alkyl-substituted acid chlorides.
In 2014, Toste and co-workers
reported the asymmetric fluorination
of α-branched cyclohexanones enabled by a combination of chiral
anion phase-transfer catalysis and enamine catalysis using protected
amino acids.
84
While simple ketones did
not undergo fluorination under conditions that had previously been
successful on other substrate classes, they supposed that inclusion
of a catalytic amount of a primary amine may form an enamine that
should be reactive to electrophilic fluorination and also act as a
hydrogen-bond donor for interaction with the chiral phosphate catalyst
(Figure 21, cycle 1). Independently, the lipophilic
chiral phosphate would undergo anion exchange with the achiral tetrafluoroborate
counteranions of insoluble Selectfluor to catalytically generate a
soluble, chiral electrophilic fluorinating reagent (cycle 2).
Figure 21
Proposed
combination of two catalytic cycles for the enantioselective
fluorination of ketones. Reproduced with permission from ref (84). Copyright 2014
American
Chemical Society.
The authors found that
the addition of achiral primary amine catalysts
did increase the yield of fluorinated product, but without any significant
enantioselective induction (Table 17). Because
glycine methyl ester gave reasonable yield, chiral amino acid methyl
esters were subsequently evaluated. While l-phenylalanine
methyl ester gave −40% ee, the two chiral catalysts were evidently
mismatched in this case, as switching to d-phenylalanine
methyl ester gave the desired fluorinated product in 88% ee. This
was able to be increased to 94% ee by the use of a 1-naphthyl version
of d-phenylalanine as increased steric bulk on the amine
acid side chain was found to play a key role. In the absence of either
catalyst, both yield and ee were found to be <10%, demonstrating
the crucial nature of both chiral catalysts to an effective transformation.
Table 17
Optimization of the Fluorination
of 2-Phenylcyclohexanone Using Two Chiral Catalysts
The scope and
limitations of this transformation were explored
(Scheme 71). A range of para- and meta-substituted aryl groups are well tolerated
in the α-position of the cyclohexanone, as well as heteroatom-substituted
cyclohexanones (198l, m). They also demonstrated
that several 2-alkenyl and 2-alkynyl cyclohexanones (198o–r) are also viable substrates
in their fluorination,
delivering good to high enantioselectivities (77–86% ee). Complete
regioselectivity was observed for fluorination at the α-position
and no undesired fluorination of either the alkene or the alkyne was
observed, although 2-alkyl-substituted cyclohexanones were not reactive
toward fluorination.
Scheme 71
2.1.5
One-Pot
and Tandem Processes
A
tandem process can be broadly thought of as one in which the substrate
undergoes multiple distinct reactions in a single synthetic operation.
If such a process involves the fluorine stereocenter being formed
after an initial step, the stereoselectivity of the fluorination step
could be either controlled by the first stereocenter or be under catalyst
control. While this Review is not intended to cover strictly diastereoselective
fluorination reactions, tandem processes will be given an overview
in this section due to the potential ambiguities of determining whether
or not the fluorination step is under catalyst control.
In 2009,
Zhao and co-workers reported an organocatalyzed intramolecular oxa-Michael
addition/electrophilic fluorination tandem reaction for the synthesis
of a series of chiral fluorinated flavanones.
85
They envisioned that by using a bifunctional catalyst, an organocatalytic,
asymmetric, intramolecular oxa-Michael addition of 201 would produce enantioenriched
enolate, which could be subjected
to electrophilic fluorination to provide chiral, fluorinated flavanone
derivatives 202. After careful optimization, it was found
that catalyst 203 efficiently catalyzed the oxa-Michael
addition of the substrate in toluene, after which NFSI and Na2CO3 were added into
the mixture to afford the desired
fluorinated flavanone product 202 in high enantioselectivities
(Scheme 72). The authors found that the rate
of fluorination varied with the catalyst used and proposed that the
catalyst may be involved in the fluorination step.
Scheme 72
In 2011, Ma and co-workers reported the diastereo- and
enantioselective
tandem 1,4-addition/fluorination catalyzed by chiral monodentate phosphoramidite
ligands 206 and copper (Scheme 73).
86
In this reaction, conjugate addition
of organometallic reagents to alkylidene β-ketoesters 204 was followed by fluorination
with NFSI to afford the fluorinated
products 205. In screening of the ligands, the authors
developed a superior catalyst 206 bearing bulky substituents
at the 3- and 3′-positions of the binaphthol unit, with the
aim of narrowing the space around the P-ligated metal
center as well as relaying the axial chirality to the reaction site.
The scope of this reaction was broad, with aryl substituents on both
the ketone and the alkylidene moieties. Alkyl-substituted substrates
(204u, v) also provide the tandem products
in good yield and enantioselectivities, and various dialkylzinc reagents
also participated to give the corresponding products (205m–p).
Scheme 73
The authors also demonstrated
that if the reaction was conducted
in a stepwise manner, the diastereoselectivity of the product was
significantly reduced, suggesting that the one-pot operation is important
to achieve stereoselectivity (Scheme 74).
Scheme 74
In 2012, Ma and co-workers reported an organocatalytic,
asymmetric,
one-pot sequential 1,4-addition/dearomative-fluorination transformation
using pyrazolones (208a, X = NR) as the aromatic partners,
thus leading to optically active fluorine-containing products with
two adjacent stereogenic centers (Scheme 75).
87
This transformation is catalyzed
by a chiral tertiary-amine-thiourea compound 211a in
combination with benzoic acid. Other nucleophilic donors were also
tested for the sequential reaction, and similarly high levels of reactivity
and stereocontrol were observed.
Scheme 75
The Michael-addition product 212 could be isolated
in almost quantitative yield with excellent levels of enantioselectivity
by omission of the fluorination step (Scheme 76). Subjecting this intermediate to
triethylamine and NFSI afforded
dearomatization-fluorination product 213 in high yield
although with slightly lower diastereoselectivity. The authors suggested
that the thiourea catalyst also plays some role in stereocontrol for
the fluorination step.
Scheme 76
A possible mechanism for this
transformation was proposed by the
authors. In the first cycle, the nitroalkene binds with the thiourea
of the catalyst through hydrogen bonding while the pyrazolone in its
enol form hydrogen bonds with the ammonium center of the chiral catalyst,
providing enantiocontrol in the 1,4 addition. The product then reassociates
with the catalyst, again through multiple hydrogen-bond interactions
for the subsequent fluorination step (Figure 22).
Figure 22
Proposed mechanism for the tandem 1,4-addition/fluorination. Reproduced
with permission from ref (87). Copyright 2012 John Wiley and Sons.
Ma and co-workers also reported a variation of this reaction
using
isoxazol-5(4H)-ones (208b, X= O) as
substrates (Scheme 75).
88
In this reaction, 1.2 equiv of Na2CO3 as an additive was found to be essential for
the high yield of the
fluorination step. A variety of different substitutions on the aryl
ring of the nitroalkenes were tolerated, although alkyl substituents
were unsuitable. The authors observed that the catalyst made some
contribution to the control of diastereoselectivity in the fluorination
step.
2.2
Nucleophilic Fluorination
Nucleophilic
sources of fluorine have only been harnessed in the context of asymmetric
catalysis much more recently when compared to their electrophilic
counterparts. This could be partially attributed to difficulties in
dealing with the high basicity of fluoride, relative to its often
fairly low nucleophilicity. In an important early report in 2000,
Bruns and Haufe described the moderately enantioselective ring opening
of meso-epoxides using a stoichiometric amount of
Jacobsen’s (S,S)-(+)-salen)chromium
chloride complex and KHF2/18-crown-6 as fluoride source.
89a
One year later, they reported opening of meso-epoxides 214 with AgF mediated by similar
chiral complexes with moderate enantioselectivities (Scheme 77).
89b,89c
Scheme 77
In 2010, Kalow and Doyle achieved a catalytic version
of the same
reaction (Scheme 78).
90
Their strategy involved amine-catalyzed slow generation of HF from
benzoyl fluoride and an alcohol, which was hypothesized to permit
the mild conditions and efficient catalysis. They envisaged that the
use of a chiral amine as catalyst may provide a route into enantioselective
catalysis. However, neither achiral nor chiral bases [e.g., (−)-tetramisole
(219)] induced measurable reactivity for epoxide opening
at room temperature until a Lewis acid was added as a cocatalyst.
A combination of (−)-tetramisole (219) and Co(salen)
catalyst (R,R)-218 gave
excellent enantioselectivities. A pronounced matched/mismatched effect
for the combination of the two catalysts was observed in this dual-catalysis
approach. Various meso-epoxides were well tolerated
in these reactions, including five-, six-, seven-, and eight-membered
cyclic epoxides with alkene, ester, and protected amine functionalities,
affording fluorohydrins 217 in 85–95% ee.
Scheme 78
Kinetic resolution of terminal epoxides 220 was also
investigated under these conditions (Scheme 79). Regioselective opening at the terminal
position afforded the corresponding
fluorohydrins 221 with up to 99% ee. The tolerance of
a silyl protecting group in one substrate (220c) under
these conditions demonstrated the relative mildness of the catalytically
generated fluoride.
Scheme 79
In 2011, the same authors described
elegant and detailed mechanistic
studies of their Co(salen) and amine cocatalyzed reaction.
91
Substituent effects in the opening of para-substituted styrene oxides established
that ring opening
is the rate-limiting step. The authors deduced that this step proceeds
via a bimetallic mechanism, based on a combination of nonlinear effects
studies with monomeric catalysts and further experiments using linked,
dimeric catalysts and that the active nucleophilic fluorine source
is a metal bifluoride (Scheme 80). With these
insights, they improved the reaction protocol by developing linked
Co(salen) catalyst 222, which afforded significantly
elevated rates, expanded substrate scope, and high enantioselectivity
for the desymmetrization of meso-epoxides 216 in conjunction with cocatalyst DBN (conditions
A, Scheme 81). Notably, an acyclic meso epoxide (216c), which underwent slow ring
opening previously (conditions B), provided
product 217c in good yield and enantioselectivity.
Scheme 80
Scheme 81
Using the dimeric catalyst, rate enhancement was also
observed
for resolution of terminal epoxides, allowing reactions to be carried
out in shorter times with impressively low catalyst loadings (Scheme 82).
Scheme 82
In 2013, Kalow and Doyle reported
the enantioselective ring opening
of aziridines.
92
While Co(salen) complexes
proved effective Lewis acids for activation of epoxides, they were
unsuitable for protected aziridines, so the authors employed a distinct
Lewis acid to achieve this. Ultimately, a combination of two Lewis
acids, the chiral Co(salen) (218) and an achiral Ti(IV)
cocatalyst, provided optimal reactivity and enantioselectivity to
deliver the trans-β-fluoroamine product 224. The use of a chelating protecting group
was crucial to
reactivity, with picolinamide being optimal. Acyclic and cyclic meso N-picolinamide
aziridines 223 underwent fluoride
ring opening in up to 84% ee (Scheme 83).
Scheme 83
Mechanistic studies supported the proposal that the chiral
(salen)Co
catalyst delivers the fluoride nucleophile while the Ti(IV) cocatalyst
activates the aziridine (Figure 23). Unlike
the previous work, dimeric Co(salen) catalysts provided no rate acceleration,
consistent with the proposed mechanistic scenario.
Figure 23
Proposed mechanistic
scenario for ring opening of aziridines. Reproduced
with permission from ref (92). Copyright 2013 Elsevier.
Doyle and co-workers have very recently applied this methodology
in a stoichiometric sense to 18F radiolabeling for PET
imaging.
93
In this procedure, the Co(salen)F
reagent is readily accessed by anion metathesis from the corresponding
tosylate with [18F]fluoride. This can then be reacted with
a range of epoxides directly to obtain the “hot” enantioenriched
fluorohydrins.
In 2010, Katcher and Doyle reported a palladium–bisphosphine
complex-catalyzed enantioselective fluorination of allylic chlorides
with AgF (Scheme 84).
94
They hypothesized that efficient and mild allylic C–F bond
formation proceeded by the nucleophilic attack of fluoride on an electrophilic
Pd(II)-allyl intermediate of the type that is well-established for
Pd-catalyzed allylic alkylation. Investigations of nucleophilic attack
on a stoichiometric Pd(II)–allyl complex showed that more basic
alkali metal fluoride sources mostly resulted in elimination; only
AgF produced a good yield of the desired substitution. Substrates
possessing traditional leaving groups for Pd-catalyzed allylic alkylation
were unreactive with the Pd(0) catalyst and AgF; however, the authors
discovered that allylic chlorides gave good yields and high enantioselectivities
of the desired products. They proposed that precipitation of AgCl
provides a strong driving force for C–F bond formation. The
commercially available Trost ligand 227 imparted high
levels of enantioinduction in the production of the allylic fluoride
products 226 (85–96% ee). The scope of the reaction
includes six-membered cyclic allylic chlorides with various functional
groups.
Scheme 84
In 2011, Doyle and co-workers successfully extended
their methodology
to encompass the regio- and enantioselective fluorination of acyclic
allylic halides (Scheme 85).
95
Interestingly, with triphenylphosphine as ligand their
reaction exhibited good selectivity for the branched allyl fluoride,
rather than the linear. They found that bidentate phosphines with
larger bite angles gave higher regioselectivity, with the commercial
Trost naphthyl ligand 230 giving >20:1 selectivity
for
the branched product. They hypothesized that one reason for this preference
could be that the small size of fluorine favors attack at the more
hindered terminus of the Pd π-allyl complex. They also suggest
the possibility of hydrogen bonding of the fluoride with the ligand,
effectively directing the nucleophile to give the branched product.
Support for this latter hypothesis is provided by the superior regioselectivity
in nonpolar solvents. Despite the high regioselectivity, moderate
to low enantioselectivities were attained with linear substrates 228a–d. The authors
found that substrates
possessing allylic substitution performed well, 228e–i bearing α-branching or heteroatom
substituents undergo
fluorination with 90–97% ee. However, fluorination of cinnamyl
chloride (228j) as substrate gave the minor branched
isomer with 0% ee, revealing the current limitations of the method.
Scheme 85
In 2012, Lautens and co-workers reported the rhodium-catalyzed
asymmetric ring opening of oxabicyclic alkenes (Scheme 86).
96
Triethylamine trihydrofluoride
(Et3N·3HF) was the optimal source of fluoride; sources
such as TBAF, KF, or Doyle’s conditions resulted in no product
formation. Use of [Rh(cod)Cl]2 and chiral Josiphos ligand
(R,S)-ppf-PtBu2 (233) afforded the ring-opened products with
high enantioselectivity. The reaction is suggested to follow a pathway
proceeding by SN2′ nucleophilic displacement, giving
the 1,2-trans product. Various substituents on the
aryl ring were well tolerated. However, with electron-donating groups
on the aryl ring (231d, f), the ring-opened
products would decompose to 1-naphthol by elimination of HF on silica
gel, a problem that was solved by mild hydrogenation after aqueous
workup, affording the stable alkyl fluorohydrin products. For unsymmetrical
oxabicyclic alkenes (231g, h), both regioisomers
were formed with high enantioselectivities, although some isomers
proved to be unstable (233g′, h′).
Scheme 86
Recently, Shibata and co-workers reported a catalytic
fluorination
system consisting of catalytic iodoarene together with HF and mCPBA.
97
This system was applicable
to two classes of substrates including the fluorination of β-dicarbonyl
compounds 47 and the intramolecular aminofluorination
of ω-amino-alkenes 234 (Scheme 87). Mechanistically, the catalyst/reagent combination
was proposed
to generate ArIF2 in situ. In the case of substrates 47, enolate attack onto the active
reagent is followed by
displacement by fluoride to give 48. In the case of aminoalkenes 234, oxidation of
the nitrogen by ArIF2 is thought
to lead to aziridinium intermediates, which are subsequently opened
by attack of fluoride to give piperidines 235. Preliminary
trials of catalytic asymmetric variants were also conducted, and promising
enantioselectivities for the desired products were obtained when (R)-binaphthyldiiodide
(236) was used as catalyst.
Scheme 87
3
Catalytic Enantioselective Trifluoromethylation
and Perfluoroalkylation
Among perfluoroalkyl groups, the
trifluoromethyl group is the most
widespread, appearing in numerous pharmaceuticals and agrochemical
compounds and being often employed in materials science. The increased
lipophilicity and often superior metabolic stability as compared to
methyl analogues often accounts for an improved activity profile in
a medicinal chemistry context.
2f
Hence,
the development of approaches for the straightforward introduction
of trifluoromethyl groups into small molecules has received much recent
attention, including methods for their asymmetric introduction.
3a
Methods for their incorporation can be broadly
classed as nucleophilic, electrophilic, or free radical processes.
However, when compared to nonfluorinated alkyl halides, the reactivity
of perfluoroalkyl halides diverges somewhat, due to the strong negative
inductive effect of the perfluoroalkyl portion. As result of this
strong stabilization of negative charge, perfluoroalkyl iodides in
some instances can be used as electrophilic iodination reagents, resulting
in the relatively stable perfluoroalkyl anion. In other cases, their
behavior can be analogous with that of electrophilic reactivity, although
closer analysis often reveals that the mechanism involves a series
of single electron transfer steps with radical intermediates.
1c
Perhaps due to this element of mechanistic unpredictability,
it is only relatively recently that advances in catalytic asymmetric
electrophilic trifluoromethylation have been made.
3c,3l,3n
3.1
Asymmetric Nucleophilic
Trifluoromethylation
3.1.1
Overview of Nucleophilic
Trifluoromethylation
Asymmetric nucleophilic trifluoromethylation
of carbonyl compounds
is an effective and direct strategy to obtain optically active trifluoromethylated
alcohols.
98
These fluorine-containing chiral
structures have been involved in the design and modification of molecules
such as Befloxatone,
99
Efavirenz,
100
and ferroelectric liquid crystals,
101b,101c
among other applications.
Me3SiCF3 (238) (commonly known as the Ruppert–Prakash reagent)
is the most commonly employed synthetic equivalent of the trifluoromethyl
anion.
102
The first example of nucleophilic
trifluoromethylation of carbonyl compounds using this reagent was
reported in 1989 by Prakash (Scheme 88).
103
Commonly, nucleophilic activators such as tetrabutyl
ammonium fluoride (TBAF) are used in a catalytic manner to activate
the Ruppert–Prakash reagent, initially generating alkoxide 239. The generally accepted
mechanism for this transformation
is depicted in Scheme 88. The fluoride anion
acts as an initiator and reacts at the silicon of 238 to form Me3SiF and tertbutylammonium
alkoxide intermediate 239. Following this initiation step, the catalytic cycle can
commence as 239 reacts with an equivalent of Me3SiCF3 to form a pentavalent silicate
that delivers a trifluoromethyl
anion equivalent to an equivalent of ketone. In this step, alkoxide 239 is regenerated
to restart the catalytic cycle.
Scheme 88
3.1.2
Asymmetric Trifluoromethylation Using Chiral
Ammonium Fluorides
Given the anionic nature of the species
proposed to attack the prochiral carbonyl compound in the generally
accepted mechanism for trifluoromethylation using the Ruppert–Prakash
reagent, an early approach to induce asymmetry aimed to take advantage
of a chiral cation strategy. Iseki, Kobayashi, and co-worker reported
the first example of such an approach catalyzed by chiral quaternary
ammonium fluorides in 1994.
104
The chiral
catalysts 241 derived from cinchona alkaloids were able
to convert the carbonyl compounds to corresponding trifluoromethylated
alcohols in high yield with moderate but encouraging enantioselectivity
(Table 18). It has been described that Prakash
showed in unpublished results in 1993 that 9-anthraldehyde was converted
to the corresponding trifluoromethylated alcohol with 95% ee.
3a
This employed a similar strategy but by using N-benzylquinidinium fluoride 241c; however,
further details such as the amount of chiral fluoride employed and
the yield of 242 are not readily available, making comparison
difficult (Scheme 89).
Table 18
Asymmetric
Trifluoromethylation of
Aldehydes and Ketones Using Chiral Ammonium Fluoride Catalysts 241
entry
R1
R2
cat. (mol %)
yield (%)
ee
1
Ph
H
241a (20)
>99
46 (R)
2
nC7H15
H
241b (20)
>99
15
3
9-anthraldehyde
241b (10)
98
45 (R)
4
Ph
Me
241b (20)
91
48
5
Ph
iPr
241b (20)
87
51
Scheme 89
In related work not involving
ammonium fluorides, Kuroki and Iseki
designed and synthesized novel chiral triaminosulfonium salts 243 in an attempt to
improve their previous results. Unfortunately,
the observed enantioselectivities and substrate scope were not significantly
improved, and only the case of benzaldehyde gave improvement, with
52% ee (Table 19).
105
Table 19
Asymmetric Trifluoromethylation of
Aldehydes Using Chiral Triaminosulfonium Catalysts 243
entry
RCHO
yield (%)
ee (%)
a
R = Ph
96
52 (S)
b
R = 4-MeOC6H4
97
37
c
R = 4-CF3C6H4
90
24
d
R = 4-ClC6H4
93
30
e
R = (E)-PhCH=CH
99
18
f
R = C6H11
88
10
Caron and co-workers at Pfizer successfully
optimized the trifluoromethylation
of a particular aryl ketone of interest to them, to give up to 92%
ee after a thorough screening of conditions and chiral ammonium fluorides
(Scheme 90).
106
Ammonium
fluoride 241d was found to be very effective for the
trifluoromethylation of the target substrate; however, the catalyst
did not prove to be generally applicable. While demonstrating what
is possible, their report suggests that enantioselective induction
using this approach may be rather substrate dependent, and extensive
catalyst screening may be required to tackle new substrates.
Scheme 90
3.1.3
Trifluoromethylation Catalyzed by Chiral
Ammonium Bromides Combined with a Fluoride Source
In the
methodology described in the previous section, the practicalities
associated with handling the highly hygroscopic chiral ammonium fluorides
somewhat limited their application. Shibata and Toru sought to address
this deficiency, and in 2007 developed the highly enantioselective
trifluoromethylation of aryl ketones with Me3SiCF3 catalyzed by a combination of the
less hygroscopic ammonium bromide
of cinchona alkaloids and tetramethylammonium fluoride (TMAF) (Scheme 91).
107
Their approach
is especially impressive because of unprecedented high enantoselectivities
of up to 94% ee for acyclic and cyclic aryl ketones. However, aryl
aldehydes and aliphatic ketones gave much poorer results. They proposed
that the obtained high enantioselectivity may result from π-stacking
interactions between the aromatic ring in the substrates and those
present in the catalysts.
Scheme 91
In 2010, Shibata and co-workers
extended the scope of this transformation
to encompass propargyl ketones, delivering trifluoromethyl-propargyl
alcohols with up to 96% ee (Scheme 92).
108
The products were transformed into aryl heteroaryl
trifluoromethyl carbinols without any loss of enantiomeric purity
of 249a (Scheme 93).
Scheme 92
Scheme 93
The same authors also demonstrated the asymmetric synthesis
of
Efavirenz in five steps from a commercially available precursor through
the enantioselective trifluoromethylation of an alkynyl ketone (Scheme 94).
109
Subsequent catalyst
modifications have allowed for improvement in selectivity in the key
step.
110
Scheme 94
The authors then turned their attention to expanding the
substrate
of the transformation to aryl aldehydes, which failed under the previous
conditions. However, the reaction of 2-naphthaldehyde (261) with Me3SiCF3 gave poor
results despite extensive
optimization attempts using various chiral ammonium bromides combined
with TMAF or KF (Scheme 95).
111
Scheme 95
In 2013, the same authors obtained some improvement
for these aryl
aldehyde substrates by incorporating sterically bulky groups into
the catalyst (Scheme 96).
112
After screening of a range of catalysts, 241i gave enantiomeric excesses of 50–70%
for a range of aldehydes.
Scheme 96
The authors detailed a mechanistic proposal that involved
TMAF
first reacting with Me3SiCF3 to provide trifluoromethyl
tetramethylammonium 265 and releasing stable Me3SiF (Scheme 97). The ammonium 265
reacts with 241i to generate chiral trifluoromethylammonium 266 and tetramethylammonium
bromide (TMAB). This chiral salt 266 then performs the trifluoromethylation of aldehyde
263 in an asymmetric manner. However, achiral ammonium 265 is also reactive to aldehyde
263 furnishing
racemic products, perhaps explaining why very high enantioselectivities
are still elusive using this approach.
Scheme 97
3.1.4
Trifluoromethylation Catalyzed by Chiral
Ammonium Phenoxides
In 2007, Mukaiyama reported the asymmetric
trifluoromethylation of ketones with TMSCF3 catalyzed by
cinchonidine-derived quaternary ammonium phenoxides, which proceeded
smoothly to afford the trifluoromethylated compounds in high yields
with moderate to high enantioselectivities (Table 20).
113
Further studies found that
α-ketoesters were also converted to the corresponding trifluoromethylated
adducts smoothly under similar conditions, potentially offering a
route to Mosher’s acid derivatives (Scheme 98).
114
Table 20
Asymmetric
Trifluoromethylation of
Ketones Catalyzed by Quaternary Ammonium Phenoxides
entry
R1
R2
yield (%)
ee (%)
1
2-(NO2)C6H4
Me
93
71
2
4-(NO2)C6H4
Me
97
73
3
3-(CN)C6H4
Me
96
71
4
3-BrC6H4
Me
97
61
5
3-(MeO)C6H4
Me
90
59
6
1-naphthyl
Me
91
51
7
2-naphthyl
Me
95
77
8
3-pyridyl
Me
90
46
9
4-pyridyl
Me
93
60
10
3-(NO2)C6H4
Et
99
64
Scheme 98
3.1.5
Chiral Ammonium Bromide
and (IPr)CuF-Catalyzed
Trifluoromethylation
A general enantioselective trifluoromethylation
of aldehydes was developed by Chen and co-workers in 2012 using (IPr)CuF
and quinidine-derived quaternary ammonium salt as the two cooperative
catalysts to activate the Ruppert–Prakash reagent. A wide range
of aromatic aldehydes participate, giving up to 92% yield and 81%
ee at only 2 mol % of catalyst loading (Table 21).
115
To gain insight into the operation
of the two catalysts, control experiments were carried out (Table 22). Neither (IPr)CuF
nor quaternary ammonium salt 241j alone produced any product, and both yields and
ee values
decreased to zero by changing the anion of the copper salt from F– to t-BuO– or Cl–.
Use of the fluoride salt of the chiral ammonium catalyst
(241c) without the copper catalyst still gave good yield
but with somewhat reduced ee (57%). This is in contrast to use of
the ammonium bromide salt that gave no conversion without the copper
fluoride. By use of (IPr)CuCl rather than (IPr)CuF with the latter
ammonium salt, the enantioselectivity increased to 67% ee. The authors
propose that (IPr)CuF is of key importance for high catalytic performance,
and the enhanced activity and enantioselectivity result from the rapid
generation of active [(IPr)CuCF3] upon reaction with TMSCF3 (Scheme 99).
116
Table 21
Asymmetric Trifluoromethylation of
Aldehydes Using (IPr)CuF and Quinidine-Derived Quaternary Ammonium
Salts
Ar
yield (ee)
2-naphthyl
90%, 75%
1-naphthyl
88%, 60%
Ph
80%, 60%
4-BrC6H4
81%, 57%
3-BrC6H4
82%, 51%
4-MeC6H4
88%, 68%
4-PhC6H4
90%, 66%
3-MeOC6H4
89%,
74%
3,4-O(CH2)C6H3
92%, 81%
3,4-O(CH2)2C6H3
92%, 79%
4-EtSC6H4
85%, 74%
Table 22
Control
Experiments
entry
ammonium
salt (mol %)
(IPr)CuX (mol %)
yield
(%)
ee (%)
1
241j (2)
NR
2
(IPr)CuF (2)
trace
nd
3
241j (2)
(IPr)CuF (2)
90
75
4
241j (2)
(IPr)CuOt-Bu (2)
57
45
5
241j (2)
(IPr)CuCl (2)
NR
6
241c (5)
87
57
7
241c (5)
(IPr)CuCl (5)
84
67
Scheme 99
3.1.6
Trifluoromethylation
Catalyzed by Chiral
Ammonium Bromide and Sodium Phenoxide
In 2007, Feng and co-workers
developed a new dual catalyst system comprising a disodium binaphtholate
salt prepared in situ and a chiral quaternary ammonium salt. This
allowed enantioselective trifluoromethylation of aromatic aldehydes
in up to 71% ee (Table 23).
117
The authors observed that the monosodium binaphtholate
salt was not an effective catalyst and speculated that the role of
disodium binaphtholate might be as a Lewis base to activate the TMSCF3 and form the
hexavalent intermediate 271 (Scheme 100).
Table 23
Asymmetric Trifluoromethylation
Using
a Dual Catalyst Approach Comprising a Disodium Binaphtholate Salt
and a Chiral Quaternary Ammonium Salt
entry
aldehyde
yield (%)
ee (%)
1
2-naphthaldehyde
85
71
2
benzaldehyde
72
56
3
4-methylbenzaldehyde
87
60
4
3-methylbenzaldehyde
88
58
5
4-chlorobenzaldehyde
72
50
6
3-chlorobenzaldehyde
95
56
7
4-phenylbenzaldehyde
73
56
8
4-methoxybenzaldehyde
87
41
9
piperonal
95
46
10
4-fluorobenzaldehyde
86
57
11
3-thiophenecarboxaldehyde
68
45
Scheme 100
Later, the same authors combined cinchonine-derived
quaternary
ammonium salt 241f with NaH to establish an effective
and fluoride-free system for the catalytic asymmetric trifluoromethylation
of methyl ketones in moderate to good ee (up to 82%) and yield (up
to 98%) (Table 24).
118
Control experiments showed that when the hydroxyl in 241f was methylated, the reaction
still proceeded efficiently, affording
the product in 63% ee and 90% yield; however, in the absence of NaH, 274 could not
catalyze the reaction. On the basis of these
observations, the authors speculated that the hydride ion might serve
as the Lewis base to activate TMSCF3 (Scheme 101).
Table 24
Asymmetric Trifluoromethylation
Using
Chiral Quaternary Ammonium Salt 241f with NaH
entry
R
time (h)
yield (%)
ee (%)
1
2-naphthyl
6
96
81 (R)
2
1-naphthyl
6
98
82
3
2-FC6H4
19
47
68
4
3-ClC6H4
19
96
68
5
4-ClC6H4
19
83
61
6
4-BrC6H4
48
43
60
7
3-NO2C6H4
48
30
68 (R)
8
4-NO2C6H4
24
64
50
9
3-MeOC6H4
96
38
58
10
4-MeC6H4
3
70
67
11
(E)PhCH=CH
22
31
59
Scheme 101
3.1.7
Asymmetric Trifluoromethylation
Using Phase-Transfer
Catalysis
The catalytic asymmetric addition of nucleophilic
trifluoromethyl to imines would constitute a concise approach to trifluoromethyl
amines. Despite the extensive studies on addition to carbonyl compounds,
this was not reported until 2009 when Shibata and co-workers reported
the enantioselective trifluoromethylation of azomethine imines 275 with Me3SiCF3 (Scheme
102).
119
The authors found
that conventional imines, such as N-tosylimines,
were poor in terms of both reactivity and selectivity. They attributed
the poor conversion to the poor nucleophilicity of the sulfonamide
intermediates that are generated upon CF3
– addition, recalling that this species must be sufficiently nucleophilic
to attack Me3SiCF3 in the autocatalytic process
previously described (Scheme 88). On this basis,
the authors selected azomethine imines 275 as likely
being superior, because the resulting anion following addition is
delocalized onto the carbonyl oxygen. They also envisaged that the
more sterically demanding and rigid nature may increase stereoselectivity
of addition. By employing bromide salts of cinchona alkaloids and
KOH as catalysts, trifluoromethylated amines 276 were
obtained with very high enantiomeric excess. Additionally, it was
demonstrated that the trifluoromethylated adduct 276a can be readily transformed to
amine 278 (Scheme 103).
Scheme 102
Scheme 103
The authors proposed a transition
state model for the reaction
in which the catalyst hydroxyl group hydrogen bonds with the carbonyl
oxygen, the sterically demanding portion of the imine is located in
the less sterically congested space, and interactions between the
aromatic rings stabilize the transition state (Figure 24).
Figure 24
Proposed transition state model to account for observed
enantioselectivity.
Reproduced with permission from ref (119). Copyright 2009 John Wiley and Sons.
Later, the same authors demonstrated
that this transformation is
viable in the environmental benign solvent-Solkane365mfc with simple
cinchona alkaloid ammonium salt 241l. Improved chemical
yields and enantioselectivities could be obtained (Scheme 104).
120
Scheme 104
In 2012, Bernardi and co-workers disclosed a racemic protocol
for
trifluoromethylation of imines employing phase-transfer catalysis
using a stoichiometric amount of an insoluble metal phenoxide as promoter.
This modification was found to overcome the aforementioned difficulties
of standard imines inhibiting the autocatalytic cycle. They disclosed
a single example of an enantioselective variant using imine equivalent 279 and cinchona
alkaloid derivative 241m as
catalyst, that proceeded with moderate yield and enantiomeric excess
(Scheme 105).
121
Scheme 105
Shibata and co-workers disclosed a novel and interesting
approach
using a cation-binding C2-symmetric chiral crown ether 281 for the enantioselective
trifluoromethylation of aldehydes
and ketones, effectively creating a chiral cation (Scheme 106).
122
Unfortunately,
the enantioselectivities induced in the desired trifluoromethylated
adducts were low to moderate.
Scheme 106
In 2013, Obijalska and co-workers
disclosed enantioselective addition
of TMSCF3 to α-imino ketones 283 derived
from aryl glyoxals 282 using a chiral ammonium bromide
catalyst and catalytic K2CO3, to form O-silyated β-imino alcohols 284. These
products were reduced to determine the ee values, which ranged from
30% to 71% (Scheme 107).
123
Scheme 107
3.1.8
Allylic Trifluoromethylation
of Morita–Baylis–Hillman
Adducts
In 2010, Shibata and co-workers published a highly
enantioselective allylic trifluoromethylation of Morita–Baylis–Hillman
adducts 286 using the Ruppert–Prakash reagent
and commercially available bis-cinchona alkaloid catalyst, (DHQD)2PHAL.
124
This process takes advantage
of the framework of the MBH adducts, which contain an allylic leaving
group to enable an SN2′ displacement if the alcohol
is suitably derivatized (R1 = Ac or Boc). The authors found
initially that if a suitably nucleophilic amine, such as DABCO, was
employed, the cationic intermediate generated (287) was
liable to a second SN2′ reaction from TMSCF3, the latter reagent presumably activated
by acetate (Scheme 108).
Scheme 108
Once the reaction was optimized
in a racemic sense, they then employed
the chiral tertiary amine catalyst (DHQD)2PHAL in place
of DABCO and were able to realize excellent enantioselectivities when
R1 = Boc, in what they describe as a successive SN2′/SN2′ process (Table 25).
Table 25
Enantioselective Allylic Trifluoromethylation
of Morita–Baylis–Hillman Adducts
entry
R1
R2
Ar
yield (%)
ee (%)
a
Ac
tBu
C6H5
trace
nd
b
Boc
tBu
C6H5
39
92
c
Boc
tBu
2-BrC6H4
37
88
d
Boc
tBu
4-BrC6H4
60
90
e
Boc
Me
C6H5
52
94
The authors showed that the β-trifluoromethyl
esters 290 obtained can be efficiently converted into
potentially
interesting carbocyclic and heterocyclic compounds without loss of
enantiomeric purity (Scheme 109).
Scheme 109
Very shortly after this report, Jiang and co-workers described
an almost identical transformation (Scheme 110).
125
They found that they were able to
perform the reaction at room temperature using a mixed solvent system
and demonstrated a broader scope of aryl and heteroaryl substitution
on the starting material, although in some cases enantioselectivity
was moderate.
Scheme 110
Building on their previous insights, Shibata
and co-workers recently
disclosed a remarkable transformation involving the kinetic resolution
of allyl fluorides 294 by enantioselective allylic trifluoromethylation,
relying on enantioselective, silicon-assisted C–F bond cleavage.
126
When the transformation is halted at 50% conversion,
enantioenriched starting material 294 and enantioenriched
trifluoromethylated compound 290 could both be isolated
with extremely high enantiomeric excesses, using (DHQD)2PHAL as catalyst (Table 26).
Similarly, excellent
results could also be obtained for pentafluorethylation and pentafluorophenylation.
Table 26
Kinetic Resolution of Racemic Allyl
Fluorides by Enantioselective C–F Bond Cleavage/Allylic Trifluoromethylation
entry
R1
R2
conv. (%)
(S)-290 ee (yield)
(%)
recovered 294 ee (yield) (%)
1
Me
Ph
54
95 (51)
97 (41)
2
Me
4-MeC6H4
53
95 (48)
96 (40)
3
Me
3-MeOC6H4
55
94 (50)
97 (40)
4
tBu
Ph
50
94 (48)
93 (42)
The proposed mechanism commences with activation of
the C–F
bond by coordination to the silicon atom of TMSCF3. This
is followed by the kinetic resolution step whereby the chiral catalyst
selectively participates in SN2′ reaction with only
one enantiomer of the activated starting material. Finally, the resulting
cationic intermediate is attacked by CF3
– in an enantioselective trifluoromethylation (Scheme 111). This mechanistic picture
is supported by the observation
that the ee of 290 is consistently high throughout the
course of the reaction, while that of 294 increases steadily
from 0 to ∼95% as the reaction progresses.
Scheme 111
3.1.9
Miscellaneous
In 1997, Kobayashi
and co-workers investigated the use of nonquaternized cinchona alkaloids
as catalysts for the trifluoromethylation of aldehydes. However, only
modest yields and low enantioselectivities were observed (Table 27).
127
Table 27
Investigation of Nonquaternized Cinchona
Alkaloids To Catalyze Trifluoromethylation of Aldehydes
entry
silane
yield (%)
ee (%)
1
Me3Si–CF3
238
49
9
2
PhMe2Si–CF3 296
39
6
3
Ph2MeSi–CF3 297
35
12
4
Et3Si–CF3
298
24
21
5
i-Pr3Si–CF3
299
trace
3.2
Electrophilic Trifluoromethylation
In contrast to their
nucleophilic counterparts, enantioselective
electrophilic trifluoromethylation reactions remain far less developed.
A number of electrophilic reagents have been developed for generation
of what can be regarded as a “synthetic equivalent”
of a trifluoromethyl cation, the development of which has been reviewed
elsewhere (Scheme 112).
3s,128
It should be considered that apparent electrophilic reactivity may
well proceed via electron-transfer induced radical mechanisms, as
perfluoroalkyl radicals tend to be of an electrophilic nature.
1c
This behavior is particularly likely with readily
oxidizable substrates such as enols and enamines.
Scheme 112
Among the first reported investigations of a catalytic
enantioselective
electrophilic trifluoromethylation reaction was that of Ma and Cahard
in 2007. They examined the use of chiral ammonium salts acting as
phase-transfer catalysts for trifluoromethylation of β-ketoester 47a, although the
highest enantiomeric excess recorded was
only 19% (Scheme 113).
129
Scheme 113
Enantioselective trifluoromethylation of this
substrate class was
not improved until Gade’s disclosure in 2012 of the trifluoromethylation
of similar β-ketoesters 309 using commercial electrophilic
trifluoromethylating agents 305a and 302a and employing chiral “boxmi” pincer ligands
52b in conjunction with copper catalysts (Scheme 114).
130
Both five- and
six-membered ring β-ketoesters were converted to the corresponding
products 310 and 311 in high yields with
up to 99% ee. To highlight the utility of the enantioselective trifluoromethylation
developed in this work, the authors demonstrated highly diastereoselective
transformations of the trifluoromethylated products (Scheme 115).
Scheme 114
Scheme 115
In a distinct approach utilizing
both enamine organocatalysis and
transition metal catalysis, Allen and MacMillan in 2010 reported the
highly asymmetric α-trifluoromethylation of aldehydes using
Togni’s reagent 305a (Scheme 116).
131
Scheme 116
They envisioned that 305a should undergo
Lewis acid
promoted bond cleavage to generate the highly electrophilic iodonium
salt 315. Condensation of the amine catalyst with an
aldehyde would generate a chiral enamine that will participate in
an enantioselective C–I bond formation via a closed-shell pathway.
The resulting λ3-iodane species 316 was
envisaged to undergo reductive elimination with stereoretentive alkyl
transfer, thus forming the new C–CF3 bond, although
a single electron transfer mechanism cannot be completely ruled out
for this latter sequence. Hydrolysis would then release the amine
catalyst and the desired enantioenriched α-trifluoromethyl aldehyde
product 313 (Scheme 117). A range
of Lewis acidic metals provided conversion and enantioenrichment,
but CuCl proved to be the most effective and a range of substituted
aldehydes were demonstrated to be effective.
Scheme 117
The authors demonstrated that the α-trifluoromethylated
aldehydes 313 were easily converted to a variety of valuable
synthons
containing the trifluoromethyl group (Scheme 118).
Scheme 118
3.3
Radical Trifluoromethylation
Trifluoromethyl
radicals can be purposely generated by a number of methods and under
the right conditions can be long-lived enough to be exploited in useful
reactivity. In contrast with alkyl radicals that are generally nucleophilic,
trifluoromethyl radicals have electrophilic character.
1c
Nevertheless, it is only recently that the first
progress has been made to harness these species in the context of
catalytic asymmetric synthesis. In 2009, Macmillan and co-workers
described a conceptually novel approach to the asymmetric α-trifluoromethylation
of aldehydes via the merger of enamine catalysis and photoredox catalysis
(Scheme 119).
132
In their proposed cataytic cycle, photoredox catalyst Ir(ppy)2(dtbbpy)+
321 accepts a photon from
a visible spectrum light source to populate the excited-state complex 326 that would
then accept a single electron from a sacrificial
equivalent of enamine to form a strong reductant Ir(ppy)2(dtbbpy) (327). At this stage,
327 engages
in single electron transfer with trifluoromethyl iodide, resulting
in an electrophilic trifluoromethyl radical 323, at the
same time regenerating the photoredox catalyst 321. In
concert with this trifluoromethyl radical formation pathway, the separate
organocatalytic cycle would be initiated by condensation of amine
catalyst 328 with aldehyde substrate 99 to
form the reactive enamine 322. These two cycles would
intersect in the trifluoromethylation step via addition of the trifluoromethyl
radical 323 to the nucleophilic enamine to form the α-amino
radical 324. A second electron transfer event with the
excited-state photocatalyst 326 would close the photoredox
cycle and deliver the iminium ion 325, prior to hydrolysis.
Scheme 119
They demonstrated the effectiveness of this approach to
trifluoromethylation
of a range of aldehydes, using a 26 W household light bulb as the
visible light source (Scheme 120). The authors
noted that at room temperature the products racemized rapidly, but
this was prevented by running the reactions at −20 °C.
Scheme 120
They nicely demonstrated that their strategy could also
be extended
to perfluoroalkylation of aldehydes using the same conditions (Scheme 121).
Scheme 121
4
Catalytic
Enantioselective Monofluoromethylation
Recognizing that despite
the increasing methods for the introduction
of the fluorine atom and perfluoroalkyl groups, there was little attention
given to reactions resulting in monofluoromethylation, Shibata, Toru,
and co-workers introduced 1-fluorobis(phenylsulfonyl)methane (FBSM)
as a fluoromethide equivalent (Scheme 122).
133a,134
This reagent was found to participate in Tsuji–Trost allylic
substitution reactions on substrates 333 to afford the
enantioenriched fluoromethylated products 334 (Scheme 123) using PHOX ligand 335a.
A similar
process has also been reported using the chiral imidazoline-phosphine
ligand 335b.
133b
Importantly,
it was demonstrated after further transformation of the aldehyde the
two sulfone groups could be reductively cleaved using activated magnesium
to leave a monofluoromethyl group.
Scheme 122
Scheme 123
The versatility of this reagent was demonstrated
by the same authors
in the catalytic enantioselective Mannich-type monofluoromethylation
of imines.
135
The prochiral imine substrates
were generated in situ in the presence of a chiral phase-transfer
catalyst and CsOH·H2O as base. With regard to the
scope of the reaction, both alkyl and aryl imines gave high enantioselectivities,
and for all substrates the reductive desulfonylation using magnesium
was demonstrated to occur without loss of enantioenrichment (Scheme 124).
Scheme 124
The same authors also reported
the catalytic, asymmetric conjugate
addition of FBSM to α,β-unsaturated ketones 339 (Scheme 125).
136
The ammonium salts of cinchona alkaloids possessing sterically demanding
substituents catalyzed the conjugate addition reaction efficiently
to give Michael adducts 340 in high yield with excellent
enantioselectivity.
Scheme 125
In 2009, several groups demonstrated
that FBSM is effective as
a nucleophile for asymmetric conjugate addition to α,β-unsaturated
aldehydes, using established amine catalyst 121 and a
catalytic cycle proceeding via iminium activation (Scheme 126).
137−139
Scheme 126
Building on Shibata and Toru’s earlier work, Zhao,
You,
and co-workers in 2009 demonstrated that [Ir(COD)Cl]2 in
combination with phosphoramidite ligand 343 is an efficient
catalytic system to allow allylic alkylation of FBSM with
1,3-unsymmetrical allylic substrates, leading to chiral products bearing
a terminal alkene with good to excellent regioselectivity (Scheme 127).
140
Scheme 127
In 2011, Shibata and co-workers disclosed another application
of
FBSM, the enantioselective allylic monofluoromethylation of Morita–Baylis–Hillman
carbonates using cooperative catalysis (biscinchona alkaloid and FeCl2). This provided
chiral α-methylene β-monofluoromethyl
esters with high ee’s of up to 97% (Scheme 128).
141
The transformation is envisaged
to proceed by SN2′ attack of the quinuclidine nitrogen
atom of the cinchona alkaloid catalyst to afford a cationic intermediate,
which is then attacked by the FBSM anion in another SN2′
reaction to give the product (see section 3.1.8 for further discussion of this approach).
The authors proposed a
transition state model to account for the observed absoute stereochemistry
and suggested that the conformation shown may be stabilized by π–π
interations in the U-shaped cleft of the (DHQD)2AQN (Figure 25). As the Si face would
be blocked
by the left half of the catalyst (as depicted), FBSM would attack
from the Re face. Supporting this scenario is the
observation that low enantioselectivities were observed in nonaromatic
substrates. The addition of a metal cocatalyst provided a modest rather
than a dramatic increase in efficiency, and the authors speculated
that this could be due to bidentate chelation to FBSM increasing its
reactivity.
Scheme 128
Figure 25
Proposed transition-state model for addition of FBSM to
catalyst-bound
intermediate. Reproduced with permission from ref (141). Copyright 2011 John
Wiley and Sons.
Recently, enantioselective
addition of FBSM to vinylogous
imines generated in situ from 2-aryl-3-(1-arylsulfonylmethyl)indoles 348 was achieved
by the same authors using a phase-transfer
catalysis strategy (Scheme 129).
142
Reductive desulfonylation gave the monofluoromethyl
adduct 350. A one-pot reaction starting from simple indoles 351 was also viable, the
first step in the procedure being
indium-promoted Friedel–Crafts alkylation with α-amido
sulfones 352.
Scheme 129
In 2010, Shibata and co-workers
introduced a variant of FBSM, 2-fluoro-1,3-benzodithiole-1,1,3,3-tetraoxide
(FBDT) 353, as a further fluoromethide equivalent.
143
They developed this reagent due to the inability
of FBSM to undergo nucleophilic addition to aldehydes. The authors
hypothesized that this is due to the instability of the resulting
addition product due to steric hindrance of the two phenylsulfonyl
groups; hence, the reverse reaction is overwhelmingly favored. FBDT
was developed as a less sterically demanding reagent to address this
problem. The authors initially demonstrated that this was successful
in a racemic sense. In 2013, they reported a catalytic enantioselective
variant using a bifunctional cinchona alkaloid-derived thiourea catalyst
and stoichiometric titanium complex. While some excellent selectivities
were obtained, this was rather dependent on the precise structure
of the substrate. The authors showed that samarium diiodide is effective
in unmasking the monofluoromethyl motif without loss of enantioenrichment
(Scheme 130).
144
Scheme 130
5
Catalytic Enantioselective Difluoromethylation
Shibata and co-workers extended their protocol previously employed
for enantioselective trifluoromethylation to reagents acting as nucleophilic
difluoromethyl equivalents. Using Me3SiCF2SePh3 as nucleophile, enantioselective difluoromethylation
of 2-naphthylmethyl
ketone (272a) was investigated (Scheme 131). The corresponding difluoromethylated
compound 358 was obtained in 41% yield with 44% ee. Some improvements in the
enantioselectivity were made by employing the more sterically hindered
cinchona alkaloid 241i, although the result was still
moderate.
145
Scheme 131
In 2008, Hu and co-workers described the catalytic enantioselective
difluoroalkylation of aromatic aldehydes with Me3SiCF2SO2Ph (361) and PhSO2CF2H (362)
employing a chiral quaternary ammonium
salt as catalyst and KOH as base (Scheme 132).
146
They found that the enantioselectivity
was substrate-dependent; for 2-chloro benzaldehyde, ee of up to 64%
was obtained.
Scheme 132
Shibata and co-workers also attempted catalytic
electrophilic or
radical enantioselective difluorobromination of β-ketoesters,
using CBr2F2 and 360b as catalyst,
with excess CsOH as base. The α-bromodifluoromethyl tetralone
carboxylate 364 was obtained in 73% yield with 37% ee
(Scheme 133).
145
Scheme 133
6
Catalytic Enantioselective Trifluoromethylthiolation
The trifluoromethanesulfenyl group (SCF3) is of special
interest due to its extremely high lipophilicity, with a Hansch parameter
of 1.44 as compared to 0.88 for CF3.
147
This could be particularly beneficial in tuning compound
pharmacokinetic properties. While this group has been utilized for
some time, it is only very recently that methods for its direct introduction
have come into the mainstream.
148
With
the development of new reagents such as 365(149) and 366,
150
new possibilities have been realized for catalytic asymmetric introduction
of the SCF3 group (Scheme 134).
Scheme 134
Very recently, Lu, Shen, and co-workers reported the preparation
of trifluoromethylthiolated hypervalent iodine reagent 366. They went on to demonstrate
that this could be employed for the
asymmetric trifluoromethylsulfenylation of β-ketoesters in the
presence of cinchona alkaloid-based phase-transfer catalysts.
151
They found that several of the catalysts in
this family were effective and that the hydroxyl group of the cinchona
alkaloids was important for the high reactivity of the transformation
(Scheme 135).
Scheme 135
While these conditions were suitable for indanone-derived
substrates,
tetralone- or benzosuberone-derived β-ketoesters required phase-transfer
catalysts in combination with a base to be used to get conversion,
and under these conditions the enantioselectivities were generally
not as high (Scheme 136).
Scheme 136
The authors considered two plausible pathways. In the
first, the
catalyst 369a reacts with 366 to give a
quaternary ammonium ion 373 that is attacked by the β-ketoester.
However, they could not observe formation of this putative species
in a stoichiometric reaction (Scheme 137).
In the second, quinine deprotonates the β-ketoester and forms
a hydrogen-bonded intermediate that is activated to attack 366. The authors favor
the latter and propose that the hydroxyl group
of the catalyst activates 366; thus, the catalyst acts
as a bifunctional catalyst to generate a highly organized transition
state that is sensitive to the exact steric nature of the substrate
(Scheme 138). They go on to present an argument
to explain the difference in reactivity between the tetralone and
indanone systems.
Scheme 137
Scheme 138
In a back-to-back publication with Shen’s
report, Rueping
and co-workers reported trifluoromethylsulfenylation of similar substrates
using the electrophilic SCF3 source 365 in
combination with cinchona alkaloid catalysts (Scheme 139).
152
The majority of their scope
was comprised of indanones, although a tetralone-derived substrate
also gave high enantioselectivity but with only moderate yield.
Scheme 139
The same authors subsequently found that 3-aryl oxindoles 50 also underwent enantioselective
trifluoromethylthiolation
using (DHQD)2Pyr as organocatalyst (Scheme 140).
153
Scheme 140
Transition metals catalysis has also very recently been
utilized
for enantioselective SCF3-transfer reactions. In 2014,
Gade and co-workers reported the highly enantioselective Cu-catalyzed
trifluoromethylthiolation of β-ketoesters by using 366 as SCF3-transfer reagent (Scheme
141).
154
They deployed their copper “boxmi”
pincer complexes, which they had previously shown to be effective
for trifluoromethylation of β-ketoesters, to great effect.
130
Scheme 141
They propose that the copper(II)
catalyst acts as a Lewis acid,
to stabilize and orientate the ester-enolate form of the substrate,
giving rise to the observed high enantioselectivity according to the
model shown (Scheme 142). They propose that
in the complexed intermediate, the Si face of the
substrate is blocked by the phenyl group of the oxazolinyl unit and
the trifluoromethylthiolation reagent 366 therefore preferentially
approaches from the Re face of the substrate, consistent
with the absolute configuration of product 376a.
155
Scheme 142
7
Summary
and Outlook
There has been remarkable progress in the past
decade toward catalytic
asymmetric methods for the introduction of fluorine and a range of
fluorine-containing groups into small molecules. No doubt this has
been spurred by the demand particularly in the pharmaceutical and
agrochemicals sectors, but these advances have necessarily gone hand-in-hand
with the development of stable and easily handled reagents. Some of
the earlier material included in this comprehensive review has been
also covered by previous reviews, but even in the five years since
2009 there have been a remarkable number of advances, which have in
some cases arisen as a result of the introduction of new reactivity
concepts, for example, photoredox catalysis and chiral anion phase-transfer
catalysis. The rapid growth and development of this field makes it
a particularly stimulating area of study with new advances being made
almost on a daily basis. Yet despite this, there are still limitations;
efforts need to be made to expand the substrate classes that can be
addressed, so that new methods truly cover new ground. To this end,
new fundamental approaches to asymmetric catalysis must be sought
as it is in this way that the variety of substrates can be significantly
expanded into new and previously unimaginable areas. One certainty
is that the introduction of fluorine and fluorine-containing groups
is now high on the list of ways to test any new asymmetric method,
which makes this an exciting arena in which to witness the latest
developments in the field of enantioselective catalysis.