(Topics in Organometallic Chemistry Vol.2) Shu Kobayashi - Lanthanides - Chemistry and Use in Organic Synthesis - Springer (1999)
(Topics in Organometallic Chemistry Vol.2) Shu Kobayashi - Lanthanides - Chemistry and Use in Organic Synthesis - Springer (1999)
(Topics in Organometallic Chemistry Vol.2) Shu Kobayashi - Lanthanides - Chemistry and Use in Organic Synthesis - Springer (1999)
During the last decade, the rare earth elements have given enormous stimulus to the field
of organic synthesis including stereoselective catalysis. This article outlines both the basic
and advanced principles of their organometallic chemistry. The intrinsic electronic features
of this 17-element series are reviewed in order to better understand the structural chemis-
try of their complexes and the resulting structure–activity relationships. Particular empha-
sis is placed on synthetic aspects, i.e. optimization of established procedures and alternative
methods with better access to catalytically relevant species. Accordingly, tailor-made ancil-
lary ligands are reported in detail and the reactivity pattern of lanthanide compounds is ex-
amined with representative examples.
List of Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
4 Ligand Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
6 Perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
7 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
List of Abbreviations
Ar aromatic residue
BINOL binaphthol
CN coordination number
COT cyclooctatetraenyl
Cp η5-cyclopentadienyl
Cp* η5-pentamethylcyclopentadienyl
DME 1,2-dimethoxyethane
HMPA hexamethylphosphoric triamide
HSAB hard soft acid base
L ligand
Ln lanthanide (Sc, Y, La, Ce-Lu)
MMA methylmethacrylate
OTf trifluoromethanesulfonato (“triflate”), CF3SO3
Ph phenyl
PMDETA N, N, N’,N’’,N’’-pentamethyldiethylenetriamine
Py pyridine
R residue
salen N,N’-bis(3,5-di-tert-butylsalicylidene)ethylenediamine
Tp tris(pyrazolyl)borate
THF tetrahydrofuran
TMEDA tetramethylethylenediamine
X ligand
Z nuclear charge
Principles in Organolanthanide Chemistry 3
1
Introduction
The rare earth elements constitute an integral part of modern organic synthesis [1].
It was about 30 years ago that the peculiar redox behavior of several inorganic rea-
gents was discovered for selective reductive and oxidative conversions [2]. In the in-
terim period fine chemicals and polymer synthesis have increasingly benefited
from the application of highly efficient organolanthanide precatalysts [3]. Due to
their intrinsic electronic properties expressed in the “lanthanide contraction”, the
rare earth elements comprising the group 3 metals Sc, Y, La and the inner transition
metals Ce-Lu provide new structural and reactivity patterns, emerging in struc-
ture-activity relationships unprecedented in main group and d-transition metal
chemistry. It is also their low toxicity and availability at a moderate price which
makes this “17-element series” attractive for organic synthesis. The spectrum of
rare earth reagents ranges from inorganic to organometallic compounds as sche-
matically redrawn in Fig. 1 with representative examples.
While highly efficient inorganic reagents such as SmI2(thf)2 and Sc(OTf)3 are
already commercially available, the more sophisticated organometallic reagents
are as a rule prepared on a laboratory scale, often under rigorous exclusion of
moisture using inert gas techniques [4]. In particular, the latter class of com-
pounds offers access to tailor-made, well-defined molecular species via ligand
fine-tuning. The consideration of the intrinsic properties of the lanthanide cati-
Cp*2Sm(thf)2 Yb(metal)
Cp*2LnCH(SiMe3)2 SmI2(thf)x
Organometallics Inorganics
"CeCl3/LiCH3" (NH4)2Ce(NO3)6
Ln(fod)3 Ln(NTf2)3(H2O)
Na3[La(S)-BINOL]3(thf)6(H2O) Ln[(-)BNP]3
ons as well as thermodynamic and kinetic factors are crucial in designing and
synthesizing novel molecular compounds. This article also includes reference to
highly reactive metalorganic compounds, pseudo-organometallics, containing
no direct metal carbon linkage; containing, however, otherwise readily hydro-
lyzable Ln-X bonds. For example, lanthanide compounds such as amide and
alkoxide derivatives not only display important synthetic precursors but also ex-
hibit excellent catalytic behavior in organic transformations [5,6]. Macrocyclic
ligands exhibiting Ln–N and Ln-O bonds are not considered in this survey [7].
The last 20 years have witnessed a rapid development in organolanthanide
chemistry and numerous review articles have been published, emphasizing var-
ious aspects including their use in organic transformations. A comprehensive
list of relevant articles has been given recently [8]. The purpose of this article is
not to give a comprehensive survey of organolanthanide compounds but rather
to address the principles of their chemistry.
2
Intrinsic Properties of the Lanthanide Elements
The rare earth elements represent the largest subgroup in the periodic table and
offer a unique, gradual variation of those properties which provide the driving
force for various catalytic processes. Their peculiar electronic configuration and
the concomitant unique physicochemical properties also have to be consulted
for the purpose of synthetic considerations. The highly electropositive character
of the lanthanide metals, which is comparable to that of the alkali and alkaline
earth metals, leads as a rule to the formation of predominantly ionic com-
pounds, Ln(III) being the most stable oxidation state [9]. This and other intrin-
sic properties are outlined in Scheme 1 which will serve as a point of reference
in this section [10–13].
2.1
Electronic Features
The Ln(III) cations of the series Ce–Lu exhibit the extended Xe-core electronic
configuration [Xe]4fn (n=1–14), a symbol which perfectly pictures the limited
radial extension of the f-orbitals: The 4f shell is embedded in the interior of the
ion, well-shielded by the 5s2 and 5p6 orbitals [14]. A plot of the radial charge den-
sities for the 4f, 5s, 5p and 6s electrons for Gd+ visually explains why Ln(III) cat-
ions are commonly thought of as a “triple-positively charged closed shell inert gas
electron cloud” (Fig. 2) [14].
Ionization energies of the elements [15], optical properties [16], and magnet-
ic moments of numerous complexes [17] prove that the f-orbitals are perfectly
shielded from ligand effects. Consequently, only minimal perturbation of the f-
electronic transitions results from the complexation of dipolar molecules. In
contrast to the broad d→d absorption bands of the outer transition elements,
the f→f bands of the lanthanides are almost as narrow in solid and in solution as
Principles in Organolanthanide Chemistry 5
Ln3+ [10] Sc Y La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
Oxophilicity [12] 165 170 190 188 167 – 136 112 170 144 122 95 159
[D0(LnO), (± 5 kcalmol-1)]
Electronegativity [13] 1.3 1.2 1.1 1.1 1.1
(Pauling)
Fig. 2. Plot of the radial charge densities for the 4f-, 5s-, 5p-, and 6s-electrons of Gd+ from
[14]
6 Reiner Anwander
they are for gaseous ions. These transitions are “LaPorte-forbidden” and result
in weak intensities which are responsible for the pale color of the trivalent spe-
cies. General principles of d-transiton metal ligand bonding such as σ-donor/π-
acceptor interaction, the “18-electron rule”, and the formation of classic
carbene, carbyne, or carbon monoxide complexes are not observed in lantha-
nide chemistry, neither do they form Ln=O or Ln≡N multiple bonds. However,
the lack of orbital restrictions, e.g. the necessity to maximize orbital overlap as
in d-transition metal chemistry, allows “orbitally forbidden” reactions. Because
of very small crystal-field splitting and very large spin-orbit coupling (high Z)
the energy states of the 4fn electronic configurations are usually approximated
by the Russel–Saunders coupling scheme [18]. The peculiar electronic proper-
ties of the f-elements have proved attractive for numerous intriguing opto- and
magneto-chemical applications (“probes in life”) [15].
The inert gas-core electronic configuration also implies a conform chemical
behavior of all of the Ln(III) derivatives including Sc(III), Y(III) and La(III). The
contracted nature of the 4f-orbitals and concomitant poor overlap with the lig-
and orbitals contribute to the predominantly ionic character of organolantha-
nide complexes. The existing electrostatic metal ligand interactions are reflected
in molecular structures of irregular geometry and varying coordination num-
bers. According to the HSAB terminology of Pearson [19], lanthanide cations are
considered as hard acids being located between Sr(II) and Ti(IV). As a conse-
quence, “hard ligands” such as alkoxides and amides, and also cyclopentadienyl
ligands show almost constant effective ligand anion radii (alkoxide: 2.21±0.03 Å;
amide: 1.46±0.02; cyclopentadienyl: 1.61±0.03) [20] and therefore fit the evalua-
tion criteria of ionic compounds according to Eigenbroth and Raymond [21].
The ionic bonding contributions in combination with the high Lewis acidity
cause the strong oxophilicity of the lanthanide cations which can be expressed in
terms of the dissociation energy of LnO [12]. The interaction of the oxophilic
metal center with substrate molecules is often an important factor in governing
chemo-, regio- and stereoselectivities in organolanthanide-catalyzed transfor-
mations [22]. Complexation of the “softer” phosphorus and sulfur counterions
is applied to detect extended covalency in these molecular systems [23,24].
Scheme 1 further indicates the tendency of the Ln(III) cations to form the
more unusual oxidation states in solution [25]. Hitherto, organometallic com-
pounds of Ce(IV), Eu(II), Yb(II) and Sm(II) have been described in detail [4].
More sophisticated synthetic approaches involving metal vapor co-condensa-
tion give access to lower oxidation states of other lanthanide elements [26].
Charge dependent properties such as cation radii and Lewis acidity significantly
differ from those of the trivalent species. Ln(II) and Ce(IV) ions show very in-
tense and ligand-dependent colors attributable to “LaPorte-allowed” 4f→5d
transitions [16b]. Complexes of Ce(IV) and Sm(II) have achieved considerable
importance in organic synthesis due to their strongly oxidizing and reducing be-
havior, respectively [1,27]. Catalytic amounts of compounds containing the “hot
oxidation states” also initiate substrate transformations. As a rule this implies a
switch to the more stable, catalytically acting Ln(III) species [28].
Principles in Organolanthanide Chemistry 7
2.2
Steric Features
Structural changes in homologous rare earth compounds arise from the lantha-
nide contraction [29], i.e. the monotonically decreasing ionic radii with increas-
ing atomic number. The 4f-electrons added along the lanthanide series from lan-
thanum to lutetium do not shield each other efficiently from the growing nuclear
charge, resulting in the contraction phenomenon. It is often this varying cation-
ic size which has a tremendous effect on the formation, coordination geometry
(coordination numbers) and reactivity of their complexes. Reports have accu-
mulated where organic substrates seem to discriminate not only between ligand
environments but also between single lanthanide elements [22]. Successful ex-
planations of these phenomena are based on the systematic theoretical investi-
gation and structural characterization of organolanthanide compounds [4].
Scheme 1 gives the trend of ionic radii of these “large” cations which prefer
formal coordination numbers in the range of 8–12 [30]. For example, consider-
ing the effective Ln(III) radii for 6-coordination, a discrepancy of 0.171 Å be-
tween Lu(III) and La(III) allows the steric fine-tuning of the metal center [11].
The structural implications of the lanthanide contraction are illustrated in Fig. 3
with the well-examined homoleptic cyclopentadienyl derivatives [31]. Three
structure types are observed depending on the size of the central metal atom: A,
[(η5-Cp)2Ln(µ-η5:ηx-Cp)]∞ 1≤x≤2; B, Ln(η5-Cp)3; C, [(η5-Cp)2Ln(µ-η1:η1-
Cp)]∞; these exhibit coordination numbers of 11 (10), 9, and 8, respectively. In
Ln Ln Ln
CN = 11 CN = 9
decreasing
La Pr Nd Pm Sm Gd Tb Ho Y Er Tm Yb Lu Sc ionic radius
a a a b b,c b b b d d d
CN = 10-11 CN = 8
Ln Ln Ln Ln
Fig. 3. Coordination modes in homoleptic, ionic LnCp3 derivatives (a belong to space group
P21; b indication from powder diffraction pattern; c show additional modifications Pbcm
and P21/n (contact dimer – effect of crystallization conditions [31b]); d belong to space
group Pna21 and exhibit lengthened intermolecular Ln-C contacts)
8 Reiner Anwander
accord with ionic bonding, small changes in ligand substitution lead to changed
coordination behavior and number (CN=10), as found in the tetranuclear ring
structure of the “MeCp” derivative. Monomeric type B is preferentially formed
with ligands bearing bulky substituents.
High coordination numbers can usually be accomplished in oligomeric struc-
tures or highly solvated complexes. However, both forms are undesirable for
synthesizing highly reactive compounds. The reactivity and stability, respective-
ly, of lanthanide complexes is correlated with the steric situation at the metal
center. Evaluation criteria as the principle of “steric saturation/unsatura-
tion/oversaturation” have been developed to explain the differences in reactivity
[32]. Hence, the main synthetic efforts as in d-metalorganic chemistry are put
into the fine-tuning of the ligand sphere to obtain tractable (volatile, catalytical-
ly reactive, etc.) compounds. Because of the importance of steric factors, ligand
environments have been numerically registrated, e.g. by the “cone-packing
model” [33], which represents a 3-D extension of Tolman’s “cone-angle model”
[34]. In this model, solid angles are calculated from structural data employing
van der Waals radii [35] and considering effects of second order packing. The in-
troduction of steric coordination numbers for various types of ligands based on
solid angle ratios further emphasizes the importance of steric considerations in
organo-f-element chemistry [36].
The Lewis acidity which is affected by the charge density (Z/r) is less distinct
in complexes derived from the large Ln(III) cations. Hence, these systems are of-
ten reported as mild Lewis acidic catalysts in organic synthesis [1]. However,
Sc(III) as by far the smallest Ln(III) cation is located in a “pole position” not only
with respect to Lewis acidity. Its “aluminum/lanthanide/early transition metal
hybrid character” [37] has revealed its superiority in many catalytic applications
[37,38]. Based on their relative preferences for pyridine, Lappert suggested a rel-
ative Lewis acidity scale: Cp2ScMe>AlMe3>Cp2YMe≈Cp2LnMe (here: Ln=large
lanthanide elements) [39]. Maximum electrostatic metal/ligand interaction and
ionic bond strength (enhanced complex stability) is also expected for scandium,
the smallest element. The Ln(III) charge density and the concomitant complex-
ation tendency also prove useful when studying the nature of Ca2+ binding in bi-
ological macromolecules exploiting the lanthanide elements as spectroscopic
and magnetic probes [15].
3
Synthesis of Organolanthanide Compounds
The availability of pure and well-defined starting materials is crucial for
straightforward and high-yield syntheses of organometallic rare earth com-
pounds. The suitability of both synthetic and catalyst precursors can be judged
by the consideration of thermodynamic and kinetic factors. For example, the
knowledge of metal–ligand bond strengths can assist in a better analysis of the
thermodynamics of archetypical ligand exchange reactions and to elaborate the
mechanistic scenarios of catalytic transformations [40].
Principles in Organolanthanide Chemistry 9
3.1
Thermodynamic and Kinetic Guidelines
Sm—CH(SiMe3)2
Sm—η3-C3H5
Sm—SnPr
Sm—NMe2
Sm—OtBu
Sm—PEt2
Sm—Br
Sm—Cl
≡
Sm—H
Sm—I
33(2) 43(5) 45(2) 47(1) 48(2) 52(2) 69(2)73(2) 81(1) 84(2) 93 97(3)
Fig. 4. Bond disruption enthalpies of organolanthanide(III) complexes. The gray area indi-
cates the bond disruption enthalpies of organolanthanide(0) arene species (TTB=η6-
C6H3tBu3-1,3,5)
10 Reiner Anwander
Inorganics Organometallics
Ln–CR3
+50
Ln–H
Ln–NR2
3.2
Inorganic Reagents
Lanthanide halides, nitrates and triflates are not only common reagents in or-
ganic synthesis (Fig. 1) but also represent, in dehydrated form, key precursor
compounds for the more reactive organometallics (Scheme 2). As a rule, in com-
pounds of strong monobasic acids or even superacids, cation solvation competes
with anion complexation, which is revealed by fully or partially separated anions
and solvated cations in their solid state structures. The tendency to form outer
sphere complexation in coordinating solvents [47] is used as a criterion of the
reactivity of inorganic salt precursors in organometallic transformations.
Ln-Halides
Anhydrous lanthanide halides are ionic substances with high melting points
which take up water immediately when exposed to air to form hydrates (I–>Br–
>Cl–) [48]. Straightforward synthetic access and a favorable complexation/solva-
tion behavior make the lanthanide halides the most common precursors in orga-
nolanthanide chemistry. Many important Ln–X bonds (X=C, Si, Ge, Sn, N, P, As,
Sb, Bi, O, S, Se, Te) can be generated via simple salt metathesis reactions [4,8]. The
so-called ammonium chloride route either starting from the lanthanide oxides or
Principles in Organolanthanide Chemistry 11
NH4Cl THF
LnCl3(H2O)x LnCl3 LnCl3(thf)x
≈150 °C/10-4 torr, -H2O soxhlet
x = 6,7 ≈ 350 °C/10-4 torr, - NH4Cl x = 1.33-3.5
SOCl2, THF
HCl(aq.) - SO2, - HCl
Sc2O3 ScCl3(H2O)6 ScCl3(thf)3
80 °C
the hydrated halides is the most popular laboratory procedure (upscaling is pos-
sible) to anhydrous lanthanide(III) chlorides (Scheme 3) [49]. Simple thermal de-
hydration which works well for lanthanide triflates leads to the formation of un-
desired lanthanide oxychlorides. Evans and co-workers have shown that the
standard recipe for dehydrating CeCl3(H2O)7 to make CeCl3/RLi will produce
[CeCl3(H2O)]n [50]. “CeCl3/RLi” is a popular Grignard-type reagent in organic
synthesis [51] which, for example, increasingly tolerates functional groups.
A coordinating solvent such as tetrahydrofuran (THF) is often necessary to
react the otherwise insoluble lanthanide halides via salt metathesis. These reac-
tions proceed via initial formation of the more soluble compounds LnX3(thf)x,
which are obtained via Soxhlet extraction and are popular, well-defined starting
reagents [52]. The extent of THF coordination depends on both the structural
type of the anhydrous lanthanide halide and the prevailing crystallization con-
ditions, and affects its solubility and hence its reactivity [53]. “ScCl3(thf)3” is
best synthesized by an alternative procedure utilizing SOCl2 as a dehydrating
agent [54]. Neutral donor ligands such as caprolactone [53a], 2,6-dimethyl-4-py-
rone [55] or chelating ligands such as DME [56] and crown ethers [57,58] also
reveal unforeseen and intriguing coordination chemistry.
Other small-scale laboratory procedures have been developed for the direct
synthesis of the more reactive THF adducts, avoiding “inconvenient” high tem-
perature treatment [59–62]. For example, the preparation of “LnCl3(thf)x” from
metal powder and hexachloroethane is facilitated by sonication [Eq. (1)] [59].
Additional metal-based synthetic routes include the redox transmetallation
with mercury(II) halides [Eq. (2)] [60] and the reaction with trimethylsilyl chlo-
ride and anhydrous methanol [Eq. (3)] [61]. Ammonia has been employed as an
alternative donating solvent in the synthesis of lanthanide alkoxides starting
from lanthanide chlorides [63].
THF
2 Ln + C2Cl6 2 LnCl3(thf)x + C2Cl4 (1)
)))
12 Reiner Anwander
+ ICH2CH2I, THF
LnI3(thf)3.5 Ln = Nd, Tm
- CH2CH2
THF
2 Ln + 3 HgCl2 2 LnCl3(thf)x + 3 Hg
∆
(2)
Ln = Yb (x = 3); Er (x = 3.5); Sm (x = 2); Nd (x = 1.5)
THF
2 Ln + 6 Me3SiCl + 6 MeOH 2 LnCl3(thf)x +
∆
(3)
+ 6 Me3SiOMe + 3 H2
NH3(liq)
Yb + 2 NH4I YbI2(NH3)x + H2 (5)
DME
TmI3 + Tm 2 TmI2(dme)3 (6)
∆
Principles in Organolanthanide Chemistry 13
Rare earth borohydrides obtained from the chlorides [Eq. (9)] [83] have been
used in salt metathesis reactions and were found to be attractive for the genera-
tion of cationic species [84]. The presence of more weakly coordinated BF4– an-
ions in [Eu(MeCN)3(BF4)3]x which can be synthesized according to Eq. (10) pro-
motes several catalytic transformations of non-heteroatom-substituted organic
substrates, including the polymerization of styrene [85].
THF, - 3 NaCl
NdCl3 + 3.3 NaBH4 Nd(BH4)3 (thf)2 (9)
60 °C, 48 h
CH3CN
Eu + 3 NOBF4 [Eu(CH3CN)3(BF4)3]x + 3 NO (10)
rt, 1d
Metals
Lanthanide metals which are conveniently prepared from the metal halides are
commercially available in the form of ingots, chips (filings), foils and powders and
are also handled as prominent synthetic precursors. For example, alkoxide com-
plexes derived from cheap and low boiling alcohols are alternatively synthesized
from metals under HgCl2 catalysis [89]. Representative examples for transmetal-
lation and transmetallation/ligand exchange reactions are given in Eqs. (11)–(13)
[90]. Ammonia solutions of ytterbium and europium react with a variety of Brøn-
sted acidic reagents according to Eq. (14) [91]. Metal oxidation/ligand transfer oc-
curs in THF in the presence of catalytic [Eq. (15)] and stoichiometric amounts of
iodine [Eq. (16)] [92]. “Lanthanide Grignard” reagents, formulated as “RLnI” are
prepared in situ from the metal and the alkyl(aryl)halide in THF [Eq. (17)] [93].
Utilization of an extremely bulky alkyl ligand allowed the isolation of
{Yb[C(SiMe3)3]I(OEt2)}2 according to a salt metathesis reaction [94].
THF
Ln + 3 TlCp LnCp3(thf) + 3 Tl (11)
80 °C, 20h
THF
Sm + Hg(C6F5)2 + 2 HNR2 Sm(NR2)2(thf)4 +
12h
(12)
+ Hg + 2 HC6F5
THF
Yb + Sn[N(SiMe3)2]2 Yb[N(SiMe3)2]2(thf)2 + Sn (13)
80 °C, 8h
NH3(liq) HR
Eu Eu2+(NH3)6x2e-(NH3)n EuR2(NH3)m (14)
- H2, -NH3
I2(!), THF, Py
Ln + 1.5 ArSSAr Ln(SAr)3(py)3 (15)
50 °C, 48 h
LaI2(thf)2
THF
2 Ln + Ph + 2 I2 (16)
Ph 50 °C, 48h Ph Ph
LaI2(thf)2
THF
Yb + CH3I CH3YbI (17)
-30 °C
P tBu
P tBu
tBu tBu P
tBu P P
tBu tBu P ScI
tBu
tBu tBu
tBu tBu
tBu Gd0 tBu Ho0 tBu ScII
tBu tBu
tBu tBu tBu I
P Sc P P
tBu tBu P tBu
P tBu
P
tBu
Fig. 5. Arene complexes of low-valent lanthanide elements obtained by co-condensation
methods
and triple-decker complexes are thermally stable and exhibit lanthanide centers
in the formal oxidation states Ln(0), Sc(I) and Sc(II) (Fig. 5) [95].
3.3
Metalorganic Reagents
1. THF, -3 KCl, 70 °C
LnCl3 + 3 KOAr Ln(OAr)3
2. sublimation (19)
pure
OAr = OC6H3tBu2-2,6; OC6H2tBu2-2,6-Me-4
Ln – C≡CR
Ln – Cp
Ln – SnR3
Ln – NR2
Ln – N(SiMe3)2 + HL Ln – PR2 + HN(SiMe3)2
Ln – OR
Ln – SR
Ln – SeR
Ln – TeR
Ln – Cl
tBu iPr
iPr
O
S
tBu iPr
O
N O N
N(SiHMe2)2 N(SiHMe2)2
Me2Si Ln N(SiHMe2)2 Ln Ln
thf thf
N N
O
tBu iPr O
S
O
iPr
tBu iPr
Fig. 6. C2-symmetric rare earth complexes according to the extended silylamide route
tBu
proposed as proceeding via THF dissociation, are not obtained with the
Ln[N(SiMe3)2]3 system.
The application of the more basic Ln(NiPr2)3(thf) as a metalorganic precur-
sor compound is controversial [119] because its availability is hampered by ate
complexation [Sect. 5.1, LiLn(NiPr2)4] and enhanced thermal instability (de-
composition at 100 °C/10–4 Torr) [120]. An efficient alkane elimination reaction
utilizing the in situ formed alkyl species Ln(CH2SiMe3)(thf)2 produced com-
plexes with linked amido cyclopentadienyl ligands (Scheme 6) [121]. However,
the thermal instability of Ln(CH2SiMe3)(thf)2 and ate complex formation seem
to be limiting factors [122].
The silylamide route can also be applied to lanthanide(II) chemistry
(Scheme 7). Although the well-characterized complexes Ln[N(SiMe3)2]2(thf)2
exhibit enhanced steric flexibility [123], the scope of exchange reactions is now
limited by the reductive properties of Sm(II). For example, Sm(II) amides tend
to get oxidized by enolizable alcohols [124]. However, aryloxides of type
Sm(OAr)2(thf)x have been isolated and ate complexation as evidenced in
[KSm(OC6H3-2,6-tBu2)3(thf)]n proves to be a stabilizing factor [125]. According
to this latter approach, mixed metallic complexes can be obtained by retention
of the original metal ligand composition. Partially exchanged heteroleptic com-
plexes such as {KSmCp*2[N(SiMe3)2](thf)2}n are available due to steric restric-
tions [126]. Eu(II) and Yb(II) silylamides are accessible to all of the exchange re-
actions listed in Scheme 5 [127].
18 Reiner Anwander
+ 3n HOAr,
n-hexane
[KSm(OC6H2tBu2-2,6-Me-4)3]n
- 3n HN(SiMe3)2
{KSm[N(SiMe3)2]3}n
(thf)2 (thf)2
K K
+ 3n HCp*,
toluene/THF
Sm Sm
- 2n HN(SiMe3)2,
- HCp*
N(SiMe3)2 N(SiMe3)2 n/2
Due to the high solubility of the starting and the resulting rare earth complex-
es, the reaction can be conducted in nonpolar solvents from which the insoluble
alkali aryloxides can easily be separated. However, this type of kinetically con-
trolled metathesis reaction is very sensitive towards the reaction conditions in-
cluding the type of alkoxide (aryloxide) ligand, type of metal, number and type
of co-ligands, stability and solubility of the eliminated alkali metal alkoxide
(aryloxide), solvent, temperature, etc. As a result, incomplete ligand exchange,
exchange of the co-ligand, ate complexation, exchange equilibria and ligand re-
Principles in Organolanthanide Chemistry 19
distribution can occur. Scheme 8 gives an idea of the complexity of these alkox-
ide-derived alkylation reactions [136–139]. Acetylacetonate complexes have
been discussed as alternative alkyl precursors [87].
Aluminum alkyls, in particular trimethylaluminum, produce chelating alkyl
alkoxide moieties, [(µ-OtBu)AlMe2(µ-Me)], via Lewis acid/base-pair formation
[140,141]. In the reaction with Y3(OtBu)7Cl2(thf)2, AlMe3 simultaneously acts as
a powerful denucleation reagent tolerating ethereal solvents such as THF at the
lanthanide center (Scheme 9). Reaction products such as LnCl3(dme)2 arise from
ligand redistributions (Sect. 5.3). The homoleptic complex Ln[(µ-OtBu)AlMe2
SiMe3
Ar Ar
O O
toluene Sm
Sm(OC6H3iPr2-2,6)3(thf)2 + 3 LiCH2SiMe3 (thf)Li Li(thf)
- LiCH2SiMe3 O
Ar
SiMe3
pentane
Cp*Ce(OAr)2 + 2 LiCH(SiMe3)2 Cp*Ce[CH(SiMe3)2]2 + 2 Li(OAr)
hexane
Cp*Ce(OAr)2 + exc. KCH(SiMe3)2 Cp*Ce(OAr)[CH(SiMe3)2] +
- K(OAr)
approx. 10 % Cp*Ce(OAr)2
hexane
Cp*Y(OAr)2 + 2 MCH(SiMe3)2 Cp*Y(OAr)[CH(SiMe3)2] + M(OAr)
M = Li, K (better yield) + MCH(SiMe3)2
+ 1.6 LiMe,
n-hexane,
- LiOAr Li(OAr)(OEt2) / toluene
Cp*Y(OAr)2 [Cp*Y(µ-Me)2]3 [Cp*Y(OAr)(µ-Me)]2
- 40 °C LiMe / hexane
R=H R = SiMe3
R = Ind
OtBu {[µ-η5:η5-C5H4(SiMe3)]Li}n
tBuO OtBu
Y Li(thf) CH2SiMe3
Li
OtBu tBuO OtBu
O Y
tBuO O (thf)2Li Li(thf)2
Li Li H 2C CH2
Li Si structurally characterized products
OtBu O = OC2H4OMe Me2
Scheme 8. Reaction behavior of lanthanide aryloxide and alkoxide complexes towards alkali
metal alkyl reagents (OAr=C6H3tBu2-2,6)
20 Reiner Anwander
Al thf
RO CH3
CH3 OR
+
R Y
CH3 Y O Al
O OR
CH3 Al R
Al OR CH3
Cl Al
OtBu
Y
tBuO OtBu 3 AlMe3, hexane extract
OtBu Cl hexane
products
Y Y rt, 12h
tBuO THF extract
OtBu OtBu (DME)
thf thf thf
Y3(OtBu)7(Cl)2(thf)2 Cl
Cl R O O
O
Y Al + Y
O O
RO O
R Cl
Cl
thf
4 MMe3,
- 3 LiCl
Ln[(µ-Me)2MMe2][(µ-Me)(µ-NMe2)MMe2]2 + 0.5 [Me2NMMe2]2
Ln(NMe2)3(LiCl)3 hexane, rt
M
excess MMe3,
H 3C CH3
- 3LiCl
CH3 Ln CH3 + 1.5 [Me2NMMe2]2
M CH3 M
CH3
Ln[(µ-Me)2MMe2]3
Scheme 10. Alkylation of lanthanide amide complexes with group 13 metal alkyls (M=Al, Ga)
(µ-Me)]3 can be obtained as the sole product from the reaction of AlMe3 with
Ln3(OtBu)9 [142].
Extended alkylation is observed when lanthanide amide complexes are used
as synthetic precursors [107,143]. The formation of a strong group 13 metal–
N(amide) bond promotes this type of alkylation reaction. Depending on the sto-
ichiometry, the reaction of MMe3 (M=Al, Ga) with Ln(NMe2)3(LiCl)3 yields
partially or peralkylated species (Scheme 10). Again, the reactivity is
determined by steric factors. For example, the sterically encumbered
La[N(SiMe3)2]3 does not show any tendency to form a Lewis acid/base adduct
with group 13 metal alkyls, a prerequisite for subsequent alkylation under these
Principles in Organolanthanide Chemistry 21
TMEDA - AlH3(tmeda)
[(C5H4tBu2-1,3)2SmCl]2 + 2 LiAlH4 [(C5H4tBu2-1,3)2SmAlH4]2(tmeda)
- 2 LiCl
+ TMEDA, - AlH3(tmeda)
[(C5H4tBu2-1,3)2Sm]2H[AlH4(tmeda)] [(C5H4tBu2-1,3)2Sm(µ-H)] 2
Scheme 11. Generation of lanthanide hydride bonds via a salt metathesis reaction
90%
0.5 [Ce(OCHtBu2)3]2 + 3 i-C4H8
Ce(OCtBu3)3
150 °C, vacuum 10%
(1/n) [Ce(OCHtBu2)2H]n + 3 i-C4H8 + tBu2CO
Et2O,
- 2 Al, - 2 H2
2 (C5H4tBu2-1,3)2Sm(thf) + 2 AlH3(NEt3) (24)
[(C5H4tBu2-1,3)2Sm(µ-H)] 2
THF
Cp*2Sm(thf)2 + AgBPh4 [(Cp*2Sm(thf)2][BPh4] + Ag (26)
toluene
Cp*2Sm + AgBPh4 [Cp*2Sm][BPh4] + “black solids” (27)
toluene,
- CH2(SiMe3)2
Cp*La[CH(SiMe3)2]2 + [NHPhMe2]BPh4
(28)
[Cp*LaCH(SiMe3)2][BPh4]
THF, - NMe3
- HN(SiHMe2)2
(salen)Y[N(SiHMe2)2](thf) + [NHMe3]BPh4
rt (29)
[(salen)Y(thf)3][BPh4]
THF
(COT)Nd(BH4)(thf)2 + [NHEt3]BPh4 [(COT)Nd(thf)4][BPh4] +
rt
(30)
+ "[NHEt3]BH4"
1. KL/MeOH, - 2 KCl
SmCl3(H2O)6 [SmL2][BPh4]
2. NaBPh4/H2O, - NaCl (31)
L = tris[3-(2-pyridyl)-pyrazol-1-yl]hydroborate
Principles in Organolanthanide Chemistry 23
3.4
Thermal Stability
Despite the kinetic lability of the Ln–X σ-bonds (even the thermodynamically
very stable Ln–OR bond undergoes rapid ligand exchange reactions [158]), or-
ganolanthanide compounds are thermally robust over a wide range of tempera-
ture [99,100,102,104,159–165]. Thermal stability is important for conducting
ligand exchange reactions and catalytic transformations at elevated tempera-
tures [1,22]. The sublimation behavior is a criterion of thermal stability, and is
frequently consulted to judge the suitability of volatile molecular precursors for
chemical vapor deposition techniques (Fig. 7).
Bulky ligands affect the ionic nature of the polarized Ln–X bond by minimiz-
ing polar interactions (intra- and intermolecular) and optimizing volatility by
the concept of steric shielding. The detection of isolated molecules instead of
salt-like arrangements in the solid state confirms this trend. The polarizing ef-
fect can also be reduced by introduction of donor-functionlized ligands which
can bring about charge transfer to the metal cation. Decomposition pathways
can be sterically blocked by filling the coordination sphere of the metal with
large ligands. However, sterically overcrowded ligands may degradate at elevat-
ed temperature as illustrated for the Ln(OCtBu3)3 system [109].
4
Ligand Concepts
Ligand design occupies a pivotal role in organolanthanide chemistry. The nature
of the ligand, including its size, basicity, and functionalization, promptly affects
complex features such as (mono)nuclearity, cation size and electrophilicity. Pro-
lific metal cation/ligand synergisms impart novel reactivity patterns which can
[Y(OCMe2CH2CH2NMe2)3]2
Nd(C5H4CH2CH2NMe2)2Cl
Sm(OC6H3tBu2-2,6)3
Ce(C5H4SiMe3)3
[Nd(OCHtBu2)3]2
[Nd(C5H4Me)3]4
[CeCp*2(µ-Cl)]2
(OEP)Sc(acac)
Ln[N(SiMe3)2]3
"La(OtBu)3"
YbCp2
Fig. 7. Sublimation behavior of various lanthanide complexes at 10–3 mbar (OEP 2,3,7,8,12,
13,17,18-octa(ethyl)porphyrin; acac acetylacetonate)
24 Reiner Anwander
R R Do
Do R (L)xMI MI(L)x Do R R
Ln R Ln R Do R Ln R Ln MI
R R R
R R
R R
Do Do Do
MI(L)x
e-homoleptic d-homoleptic ate-e-homoleptic ate-d-homoleptic
4.1
Steric Bulk and Donor Functionalization
[(η5-C5H5)2Sm(µ-η5:η2-Cp)]∞ Sm[η5-C5(CH3)5]3
[Ln(CH3)3]a Ln{CH[Si(CH3)3]2}3
[Yb(CH3)2]a Yb{C[Si(CH3)3]3}2
[Ln(NMe2)3]a Ln{N[Si(CH3)3]2}3
[Ln(OC6H5)3]b Ln{OC6H2[C(CH3)3]2-2,6-CH3-4}3
[Ln(SC6H5)3]b Ln{SC6H2[C(CH3)3]2-2,4,6}3
Scheme 13. Tractable, mononuclear species via steric cally bulky ligands (acompound not
known/ not stable; bdegree of agglomerization not determined)
Me2N
PMe2
SiMe3
Ln–C–R Ln SiMe3 Ln SiMe3
Ln
Si CH3
PMe2
tBuO Me2N
CH3
NMe2
– –
Ln–O–R Ln O Ln O Ln O
Si
MeO Me2N
Me2N
PPh2
NMe2
OCH2Ph
H3C
N Si Si
H3C
N
NMe2
MeO
4.2
Ancillary Ligands
(thf)2
Na F3CF2CF2C
SiMe3
O O O
Ln N
Me2Si O
SiMe3 O La Eu
H O Na(thf)2 O
(thf)2Na 2 O
O
3
R R R
N O
Ln Ln Ln
Monovalent Ligands
Numerous mono-charged cyclopentadienyl substitutes have been discussed in rel-
evance to the catalysis topic. The attachment of one or two of such ligands is easily
perfomed and may render two or one reactive sites, respectively. The resulting lig-
and sphere offers enhanced steric flexibility, and kinetic inertness of the monova-
lent ancillary ligand is often achieved via chelation through charge delocalization
or donor functionalization. The stability of the resulting complexes, however, can
be affected by ligand redistribution (disproportionation) and formation of the ho-
moleptic system (Sect. 5.3). A representative selection of this ligand type is shown
(Fig. 13), comprising N,N’-bis(tert-butyl)glyoxaldiimine [(dad)Li]– [211], N-iso-
propyl-2-(isopropylamino)troponimine [(iPr)2ati] [212], substituted benzamidi-
nates [213], substituted tris(pyrazolyl)borates Tp-Rx [214], (N,O-bis(tert-butyl)
(alkoxydimethylsilyl)amide [215], tri(tert-butyl)methoxide (tritox) [102], substi-
tuted aryloxides [137], tri(tert-butyl)siloxide (silox) [102,216] and functionalized
siloxide ligands [217].
Divalent Ligands
Heteroleptic complexes derived from doubly charged, “linked” ligands consti-
tute a class of well-defined metallocene-analogous precatalyst species. Various
C1- and C2-symmetric members of the bis(cyclopentadienyl) fragment have
been reported [218], including linked amido-cyclopentadienyl [219] and dihy-
droanthracene-cyclopentadienyl ligands [220] (Fig. 14). Such divalent ligands
when coordinated in a chelating fashion provide a strongly bonded, rigid back-
bone which not only imparts kinetic stability but is also a prerequisite for asym-
metric induction at the metal center. By nature, the synthesis of these ancillary
ligands is more costly/lengthy and subsequent complex preparations may re-
Principles in Organolanthanide Chemistry 29
SiMe3
N N N
N N
LnR2 Li R' N N
H B
N N N N N
SiMe3
Me2N
Me O
Si C O O Si O Si O
Me N
NMe2
SiMe3
B
C B
Me3Si C
B
B
Si Ln R C Si (THF)Li
N B
B
Me3Si C
C B
B
SiMe3
N
N N O O
N
N N
Trivalent Ligands
Complexes derived from triply charged ligands formally correspond to the ho-
moleptic tris(cyclopentadienyl) complexes LnCp3. These compounds can be of
relevance in Lewis acid catalysis where their activity is directed by the formation
of Lewis acid (catalyst)/base (substrate) pairs [229]. Hence, metal–ligand bond
disruption and formation processes are pushed into the background. Linked cy-
clopentadienyl-carborane ligands form mononuclear commo (sandwiched)
metallacarborane complexes [230]. Highly functionalized podate ligands such
as triamidoamine (“azatrane”) [231] and tribenzyltrifluoroacetoacetate [232]
produce formally 4- and 6-coordinated complexes, respectively (Fig. 15). A di-
nuclear composition has been proven for Ln(III) complexes of the trivalent oli-
gosilsesquioxane ligand, T7(OH)3 [233]. These incompletely condensed ligands
SiMe2tBu
tBuMe2Si
N
Si N
Ln
C B N
B C
B B B B
B N
B
B B
SiMe2tBu
O
Si
O O
O O
Si O Si
O Si O
Si O
O O O
O O O
O
Si Si F3C O
O CF3
CF3
4.3
Immobilization – “Supported Ligands”
C3F7
OiPr
C
O Me2HSi Nd
O Eu1/3
O O O
O O O
CH3 Si Si Si
O Si O Si
MCM-41
CH3 CH3
m n
5
Reactivity Pattern of Organolanthanide Complexes
The intrinsic properties of the rare earth cations as revealed by their oxophilici-
ty, “hardness”, and large size govern the reactivity of organolanthanide com-
plexes. Hence, parallels to the chemistry of aluminum, the group 2 and group 4
elements, and the actinides are often detected. In this section the most impor-
tant reaction pathways of these highly reactive complexes are surveyed with rep-
resentative examples. The stoichiometric reactions outlined also represent the
elementary steps in catalytic reaction sequences.
5.1
Donor–Acceptor Interactions
The interaction of the hard, Lewis acidic Ln(III) centers with neutral electron-
donating moieties is a ubiquitous feature of organolanthanide complexes. In ad-
dition to the electron deficiency of the lanthanide center, steric unsaturation di-
rects this stabilization of the complexes via adduct formation.
dination of soft Lewis bases was observed in the case of terminally bonded PMe3
in a highly electrophilic scandium complex [Eq. (35)] [249]. Even water com-
plexes could be trapped as the initial step of the hydrolysis reaction [250]. The
vast majority of organic transformations mediated by lanthanide centers de-
pend on pre-coordination of a neutral, functionalized substrate and subsequent
formation of an activated species.
hexane (32)
Ln[N(SiMe3)2]3 + O=CPh2 Ln[N(SiMe3)2]3(O=CPh2)
hexane
[Dy(OCHtBu)3]2 + 4 CH3CN 2 Dy(OCHtBu)3(CH3CN)2 (33)
N
THF
(C5Me4Et)2Yb(thf) + : (C5Me4Et)2Yb(carbene) + THF (34)
N
"carbene"
H2, PMe3,
n-hexane
5 1
2 [(η -C5Me4)SiMe2(η -NtBu)]ScCH(SiMe3)2
-CH2(SiMe3)2
(35)
PMe3
H SiMe2
Me2Si Sc Sc
N H PMe3 N
tBu tBu
toluene, - THF
Ce[C5H3(SiMe3)2-1,3]3(thf) Ce[C5H3(SiMe3)2-1,3]3 (36)
∆, vacuum
n-hexane
Ln[N(SiHMe2)2]3(thf)2 + AlMe3 Ln[N(SiHMe2)2]3(thf) + AlMe3(thf) (37)
toluene
Cp*La[CH(SiMe3)2]2(thf) + [Merrifield polymer]-CH2SiMe2I
(38)
Cp*La[CH(SiMe3)2]2 + [Merrifield polymer]-CH2SiMe2O(CH2)4I
34 Reiner Anwander
route A
toluene
Cp*LaI2(thf)3 + excess Me3SiI [Cp*LaI2]n + 3 Me3SiO(CH2)4I
+ 2 K[CH(SiMe3)2], OEt2
Cp*La[CH(SiMe3)2]2
route B
hexane/OEt2
Cp*LaI2(thf)3 + 2 K[CH(SiMe3)2] Cp*La[CH(SiMe3)2]2(thf) + 2 KI
Scheme 14. Solvent-free rare earth alkyl species – THF removal at an early stage
Ate Complexation
The formation of anionic rare earth metal ligand moieties or ate complexation
are commonly observed features of salt metathesis reactions when alkali metal
cyclopentadienyl [147], alkyl [96], amide [97,255] and alkoxide derivatives are
employed [Eqs. (39)-(42)] [256].
1. THF
NdCl3 + 2 LiCp* Cp*2 Nd(µ-Cl)2Li(OEt2)2 + LiCl (39)
2. OEt2
THF, PMDETA,
- 2 LiCl
LaCl3 + 3 LiCH(SiMe3)2 La[CH(SiMe3)2]3(µ-Cl)Li(pmdeta) (40)
rt
THF/TMEDA
SmCl3(thf)3 + 2 LiNiPr2 (NiPr2)2Sm(µ-Cl)3Li2(tmeda) (41)
rt, 3h
Me2N
THF, O
- 2 NaCl
LuCl3 + 3 NaO Cl Lu Na (42)
O
Me2N
O
(tmeda)
Li (tmeda)
CH3 Li S
CH3 N N S
H3C S
Er Li (tmeda) Nd Li(thf) Yb Li (tmeda)
H 3C CH3 S S
N N
CH3
Li S
Li (tmeda)
(tmeda)
SiMe3
Me3Si
La Y
SiMe3
Me3Si
[Li(µ-C4H8O2)1.5]+ {(Me3SiCH2)x(OtBu)1-xY(µ-OtBu)4[Li(thf)]4(µ4-Cl)}+
CH3
Yb Pt(PPh3)2 Y Yb
CH3
iPr
ArO O Sm
ArO
Cl Cl Sm Cl Cl La
Al iPr N
Al Cl iPr La N
Cl Cl Cl OAr
Cl O
Cl OAr
iPr Sm
Al
Cl Cl
Agostic Interactions
The term “agostic” bonding, originally proposed for the formation of two-elec-
tron three-center bonds of the type C–H→M [269], is now often used in lantha-
nide chemistry to describe the interaction of a highly electron-deficient, sterical-
ly unsaturated metal center with “CH”, “SiMe”, and “SiH” ligand fragments.
These intramolecular, chelate-type interactions are of predominantly electro-
static nature (Fig. 19) [270]. Although the agostic bonding is weak and usually
not observed in solution [271], it can have significant implications for the mo-
lecular and electronic structure and hence reactivity of the molecule. The solid
state structures often reveal quite remarkable angle distortions within the agos-
tically interacting fragment. Detailed studies performed on complexes which
are active in olefin polymerization propose that α-C-H agostic interactions as-
sist chain propagation (transition state) [272], while β-C-H interactions retard
ethylene insertion (ground state) [273]. Strong β-Si-H diagostic interactions can
be formed even in the presence of a coordinating solvent. Such potentially tri-
dentate chelating arrays direct the rac/meso ratio in ansa-lanthanidocene com-
plexes [117]. The intramolecular agostic approach is routinely observed in elec-
tron-deficient complexes of the bulky ligand CH(SiMe3)2 [274]. The presence of a
Principles in Organolanthanide Chemistry 37
H H tBuH2C
HH C AlMe2 RO OR
Me2Al C O
H H La La
H O
H R
CH2 Nd C tBuH2C
C
Sc H H H O RO O CH2tBu
CH2 H OR
H H R
H C HH C O
H H La La
RO O OR
AlMe2 CH2tBu
Me3Si SiMe3
H
N N SiMe2 SiMe2
H H Me2Si
Me3Si SiMe3
C La C H3C Yb CH3 Me2Si Y N
5.2
Complex Agglomerization
Steric and electronic factors often force the stabilization of monometallic species
via agglomerization. As a rule, the formation of di- and multinuclear species is
achieved by intermolecular bridging of the smallest, most reactive and labile
Ln–X bond and, hence, leads to decreased reactivity.
Dinuclear Complexes
The formation of dinuclear complexes is routinely observed along with the im-
portant class of lanthanidocene complexes of type Cp2LnR when R is a sterically
less demanding ligand such as H, Me or a small alkoxide group [4]. Depending
on the metal and ligand size, bridging can also occur in an asymmetric fashion
as evidenced in [Cp*2Lu(CH3)]2 [Eq. (43)] [145]. The formation of dinuclear lig-
and bridged species is also observed in organo-Ln(II) chemistry. Two represent-
ative examples are given in Eqs. (44) und (45) [127,278]. Rare earth amide and
38 Reiner Anwander
THF I
2 SmI2(thf)2 + 2 KCp* (thf)2Cp*Sm SmCp*(thf)2 (44)
I
+ 2 HOCtBu3,
hexane, CtBu3
- 2 HN(SiMe3)2 O
{Yb[N(SiMe3)2]2}2 (Me3Si)2N–Yb Yb–N(SiMe3)2 (45)
rt, 16h O
Bu3Ct
n-hexane,
- 6 HN(SiMe3)2
2 Lu[N(SiMe3)2]3 + 6 HOCMe2CH2OMe
rt, 16h (46)
(η2-OR)Lu(µ,η2-OR)3Lu(η1-OR)2
5.3
Ligand Exchange and Redistribution Reactions
The vast majority of ligand exchange reactions such as salt metathesis, amine
elimination, and hydrogenolysis have already been addressed to in the previous
sections. The high ligand exchange ability is a peculiar feature of lanthanide
complexes and of fundamental importance for their catalytic application. Even
the thermodynamically very strong lanthanide alkoxide and amide bonds are
kinetically labile as found in transamination and transalcoholysis reactions
[158]. Donor ligand exchange as a rule occurs via dissociation processes and the
exchange rate of water has been studied in detail in the context of biologically
and medically relevant processes [286]. Counter ligand exchange proceeds via
40 Reiner Anwander
δ+ δ-
+ HSiPh3 Ln C
pentane - CH2(SiMe3)2
Cp*2LnCH(SiMe3)2 Cp*2LnSiPh3
H Si
δ- δ+
H2, c-hexane
[Me2Si(NtBu)(OtBu)]2YCH(SiMe3)2 {[Me2Si(NtBu)(OtBu)]2Y(µ-H)}2
50 °C, 3d
Y[Me2Si(NtBu)(OtBu)]3
[C5H4(SiMe3)]2Y(OtBu)2Li(thf)2 + Y[N(SiMe3)2]3
pentane
[Cp*2Nd(µ-H)]2 + 3
0 °C, 16h
2 Cp*2Nd + CH3CH2CH3 (48)
c-hexane,
- 4 CH2(SiMe3)2
4 Cp*2CeCH(SiMe3)2 {Cp*3Ce2[µ3-η5,η1,η1-C5Me3(CH2)2]}2 (49)
125 °C, 24h
toluene (50)
1.5 {Eu[N(SiMe3)2]2(µ-Cl)(thf)}2 2 Eu[N(SiMe3)2]3 + EuCl3(thf)x
5.4
Insertion Reactions
The formal insertion of double and triple bonds into reactive Ln–H, Ln–C, and
Ln–N bonds is of fundamental significance in a large number of catalytic reac-
tions. Direct insertion of olefins such as propene into a Lu–Me bond has been
detected as part of the “lanthanide model for Ziegler–Natta polymerization”
[Eq. (51)] [291]. Alkene and alkyne insertion is also a key step in the hydroami-
nation/cyclization reaction of N-unprotected aminoolefins [292]. Insertion of
polarized double bonds such as carbon monoxide and acetonitrile often leads to
highly functionalized reaction products. Carbon monoxide insertion into lan-
thanide alkyl bonds can occur in a single and multiple fashion with the forma-
tion of η2-acyl and enedione diolate moieties, respectively (Scheme 18) [293].
The insertion of the metal-bonded carbonyl of CpCo(CO)2 into a Sc–Me bond
led to a heterobimetallic system [Eq. (52)] [294]. A non-classical carbocation
species was isolated from the insertion of CO into one cyclopentadienyl moiety
of the sterically crowded complex SmCp*3 [Eq. (53)] [295]. This is a remarkable
reaction promoted by steric constraints. A “Sc-Si(SiMe3)3” moiety mediates the
insertion and coupling of the isocyanide CN(Xyl) (Scheme 19) [296]. The subse-
quent rapid intramolecular rearrangement reactions probably proceed via a re-
active silene intermediate.
CH3
H
c-hexane CH
Lu CH3 + H2C C Lu (51)
CH3 CH2 CH3
C O
toluene
Cp*2Sc–CH3 + CpCo(CO)2 Co Sc (52)
rt, 12h
C O
H 3C
toluene O
Cp*3Sm + 2 CO Sm (53)
rt, 5min
O
42 Reiner Anwander
Lu
O O
tBu O tBu
CO excess CO
Lu Lu C tBu
toluene toluene tBu
THF O
O
Lu
Scheme 18. Single and multiple carbon monoxide insertion into a lutetium alkyl bond
Xyl
CNXyl, N CNXyl,
benzene benzene
Cp2Sc[Si(SiMe3)3](thf) Cp2Sc C
Si(SiMe3)3
N N SiMe3 N SiMe3
C Cp2Sc
Cp2Sc
Cp2Sc C N Si(SiMe3)2 N
SiMe3 Si(SiMe3)2
N Si
SiMe3
Xyl SiMe3
Scheme 19. Double insertion of CN(Xyl) into a scandium silicon bond (Xyl=C6H3Me2-2,6)
5.5
Elimination Reactions – Ligand Degradation
β-H and β-alkyl elimination have been reported as major decomposition routes
of lanthanide alkyl bonds. These extrusion reactions were initially observed in
Cp*2Ln-R catalysts and can be considered as models for chain termination oc-
curing during propene polymerization [Eqs. (54) and (55)] [297]. The intramo-
lecular ligand metallation and concomitant hydrocarbon extrusion found in
amide templated complexes depend on the metal size and alkyl ligand
(Scheme 20) [298]. Silylamine fragmentation has been observed along with the
synthesis of the alkoxide complexe Tm(OCtBu3)3 according to the silylamide
route [Eq. (56)] [114]. In corresponding exchange reactions with excess of fluor-
inated alcohols HORF, ammonia is the final degradation product of the si-
Principles in Organolanthanide Chemistry 43
THF, toluene,
- MCl -RH
LnCl[N(SiMe2CH2PMe2)2]2 + MR LnR[N(SiMe2CH2PMe2)2]2
rt, 20min ∆
Me2P PMe2
N
Si
Si Me2
Me2
lylamine, trapped as an ammine complex. The bulkiness and the basicity of the
alkoxide ligand, and the type of solvent, effect the ammonia formation [299].
Due to its relevance to the conversion of oxofunctionalized substrates the unex-
pected cleavage of acetylacetone in the presence of an yttrium alkoxide is shown
in Eq. (57) [300].
c-hexane
Cp*2LuCH2CH(CH3)2 Cp*2LuH + CH2=C(CH3)2 (54)
c-hexane
Cp*2LuCH2CH(CH3)2 Cp*2LuCH3 + CH2=CH(CH3) (55)
- HN(SiMe3)2
2 Tm[N(SiMe3)2]3 + 6 HOCtBu3
(tBu3CO)3Tm (56)
SiMe2
H 2N CH2 NH2
SiMe2
Tm(OCtBu3)3
toluene
Y5O(OiPr)13 + excess Hacac
20 °C, 24 h
CH3
H 3C CH3
O OH2
O O CH3
H3C (57)
O O O
Y Y
O O O
H3C O O O CH3
H2O
H 3C CH3
H 3C
44 Reiner Anwander
5.6
Redox Chemistry
Kagan and co-workers pioneered the work on the reductive behavior of the low
oxidation states of the lanthanide elements in organic synthesis [2b]. Ln metals
and Ln(II) derivatives were subsequently found to promote a number of impor-
tant individual reactions [301]. “The combination of one- and two-electron
chemistry sets SmI2 apart from virtually every other reductive coupling agent
currently available” and exhibits exceptional properties for sequential conver-
sions tolerating unprotected functional groups [1b].
Evans and co-workers have worked out the peculiar reducing ability of Sm(II)
with the corresponding organometallic reagents, in particular Cp*2Sm(thf)x,
and characterized many metal-bonded products by X-ray crystallography [32].
Their standard reagent is readily oxidized by oxygen to form an oxo-bridged
dimer [Eq. (58)] [302]. In contrast, the tris(3,5-dimethylpyrazole)hydroborate
ligand produces a superoxo complex under an atmosphere of oxygen [Eq. (59)]
[303]. Again, steric and electronic constraints at the metal center, induced by the
ligand environment, seem to force this different reaction behavior. Examples of
the unique reductive potential of Cp*2Sm include the functionalization of un-
saturated hydrocarbon substrates with carbon monoxide (Scheme 21) [304, 305].
toluene
Cp*2Sm(thf)2 + O2(g) Cp*2Sm O SmCp*2 (58)
rt
toluene O
Sm(TpMe2-3,5)2 + O2(g) (TpMe2-3,5)Sm (59)
-78 °C → rt
O
Ph Cp*2SmO H
CH CH
Ph
CO (90 psi), hexane, 24 h
thf H OSmCp*2
Sm
thf Ph
Ph
N N O N
Ph C
Cp*2Sm
CO (90 psi), THF, 24 h SmCp*2
C
N O
Ph
The reaction of one-electron reducing agents with ketones yields radical ani-
ons (ketyls) which are key intermediates in a variety of carbonyl group transfor-
mations. Reduction of aromatic ketones by low-valent samarium and ytterbium
compounds [Ln metal and Ln(II)] allowed the first trapping of these radicals in
the coordination sphere of a metal [306]. The isolation of an Yb(II) benzophe-
none dianion [Eq. (60)] [307], a lanthanoid-imine azametallacyclopropane
complex [Eq. (61)] [308], and a heteroleptic fluorenone ketyl organosamar-
ium(III) complex [Eq. (62)] [309] by Hou and co-workers emphasize the useful-
ness of hexamethylphosphoramide (HMPA) and sterically demanding groups as
stabilizing ligands. The reaction of La[η5-C5H3(SiMe3)2-1,3] with a potassium
mirror in dimethoxyethane produced a lanthanocene(III) methoxide complex
via persistent paramagnetic La(II) intermediates (Scheme 22) [310].
Ph
Ph
O O
THF/HMPA
2 Yb + 2 C (hmpa)2Yb Yb(hmpa)2 (60)
rt, 1h
Ph Ph O
Ph Ph
Ph
Ph
N N
THF/HMPA
Yb + C Yb(hmpa)3 (61)
rt, 4h
Ph Ph
Ph Ph
THF O (62)
Cp*2Sm(thf)2 + Sm
rt, 2h
thf
O
K(mirror), DME
LaCp''3 [K(dme)x][LaCp''3] [LaCp''2(dme)y] + [KCp''(dme)z]
1. -40 °C, 3h
2. rt, 45h
[LaCp''2(µ-OMe)]2 + ?
- KCp''
+ 6 NaNO3
+ other products
(63)
+ Ce
OtBu
THF,
- Ag, - KI
K{Ce[C8H5(SiMe3)3-1,3,6]2} + excess AgI
rt, 1h
SiMe3
Me3Si
(64)
SiMe3
Ce
Me3Si
SiMe3
Me3Si
benzene
2 Ce(OCtBu3)3 + C6H4O2
rt, 30min
Ce(OCtBu3)3 (65)
O O
(Bu3CtO)3Ce
5.7
Reaction Sequences – Catalytic Cycles
The vast majority of the reaction pathways outlined in the preceding sections
can be rediscovered as basic steps in many organolanthanide-mediated organic
transformations. The well-examined mechanistic scenarios shown as follows
Principles in Organolanthanide Chemistry 47
Cp'2Ln N(SiMe3)2
R R
H2N
HN(SiMe3)2
R
R
H2
N
H R R
N
* Cp'2Ln
H 2N
HN
R R
R R
R R
NH2
R R
R R
H2 H2
N N
H H
Cp'2Ln N Cp'2Ln N R
*
R
R
R
with the hydroamination (Scheme 23) [218a,292], Michael addition (Scheme 24)
[313], and MMA polymerization reactions (Schemes 25 and 26) [314,315] sum-
marize this section.
5.8
Side Reactions
O
O (thf)2
Na
O
O O +
H OMe O O
O Ln O
OMe MeO OMe
(thf)2Na H2O O Na(thf)2
O O
MeO
(thf)2Na (thf)2Na
O
* O
* O
O O
O Na(thf)2 O Na
O OMe
* Ln O
* Ln O
O OH O OH
O O
(thf)2Na H (thf)2Na
* O O
*
MeO OMe
OMe OMe
O C O C Me
MMA
Sm Me Sm Me CH2R
R CH2
C CH2
O C
R = H, Me OMe
OMe OMe
Me O C Me
O C
CH2R
MMA
Sm CH2 Sm
CH2
Me Me
O C O C
COOMe
C C C
Me
H2 CH2R
OMe OMe
O
2 MMA
2 Sm(II) 2 OMe Cp*2Sm(III)
MeO Sm
O
2n-2 MMA
O
Sm OMe
CO2Me Sm
MeO O
n n
O MeO2C OMe
Sm
"link"-functionalized polymers
benzene,
- HC(SiMe3)3, - C2H4
Yb[C(SiMe3)3]2 + 2 Et2O 0.5 [C(SiMe3)3Yb(µ-OEt)(OEt2)]2 (66)
20 °C
toluene, - THF
Cp2Yb(CH3)(thf) / LiCl(thf)x [Cp2Yb(µ-OCH=CH2)]2 (67)
reflux , 72 h
O
Me2Si SiMe2 THF
[Cp*2Sm(µ-H)]2 +
O O Me2 Me2
Si O Si Si O (68)
Me2 O
Cp*2Sm SmCp*2
thf thf
crystallization
hexane / H2O (traces)
2 (C5H4tBu)3Ln [(C5H4tBu)2Ln(µ-OH)]2 + 2 HCptBu (70)
-35 °C
toluene, -THF
3 Ce(OtBu)4(thf)2 Ce3(OtBu)10O (71)
rt, 2-3d
6
Perspectives
Examples of the exceptional and intriguing potential of rare earth reagents in or-
ganic synthesis will be treated in comprehensive form in the following chapters
of this volume. The text of this chapter on “Principles in Organolanthanide
Chemistry” has been directed towards a basic understanding of the chemistry of
the most reactive members of this family. Improved and alternative synthetic
procedures ensure the availability of both inorganic and organometallic rea-
gents in pure and well-defined form. Each of the important areas of reactivity of
organolanthanide compounds which have been addressed to in this survey
should prove fertile for further development. Examination of such highly reac-
tive species will provide important details to explain the reaction pathways of
the inorganic reagents in organic synthesis by means of spectroscopy and struc-
ture determination. So far, the vast majority of active components and reaction in-
termediates is under-determined due to the application of in situ reaction se-
quences. The operating system is often a “black-box” and process optimization
is achieved by empirical methods. This dearth of data of the active components
should further stimulate the interaction between organic synthesis and organo-
metallic chemistry. On the other hand, chiral organometallic and pseudo-orga-
nometallic reagents challenge the field of enantioselective catalysis. For this rel-
atively young branch of lanthanide chemistry, ligand design has become indis-
pensable. Since rare earth ligand interactions are ruled by simple principles such
as ionic bonding and the HSAB theory, combinatorial chemistry could prove a
valuable tool for ligand fine-tuning. In addition to the evaluation of novel ancil-
lary ligand sets, detailed studies on supramolecular rare earth chemistry are to
be expected, tackling host-guest interactions [322], the topic of immobilization
[235], and dendrimer chemistry [323]. Clearly, organolanthanide chemistry has
much to offer to the field of organic synthesis.
7
References
1. For recent reviews, see: (a) Imamoto T (1994) Lanthanides in organic synthesis. Aca-
demic Press, London; (b) Molander GA, Harris CR (1996) Chem Rev 96:307; (c) Nair
V, Mathew J, Prabhakaran J (1997) Chem Soc Rev 127
Principles in Organolanthanide Chemistry 51
2. (a) Evans DF, Fazakerley GV, Phillips RF (1970) Chem Commun 244; (b) Kagan HB,
Namy JL (1984) In: Gschneidner KA Jr, Eyring L (eds) Handbook on the physics and
chemistry of the rare earths. North-Holland Publishing Company, Amsterdam, chap 50
3. For recent examples, see (a) Li Y, Marks TJ (1998) J Am Chem Soc 120:1757; (b) Gröger
H, Saida Y, Sasai H, Yamaguchi K, Martens J, Shibasaki M (1998) J Am Chem Soc
120:3089
4. (a) Schumann H, Meese-Marktscheffel JA, Esser L (1995) Chem Rev 95:865; (b) Edel-
mann FT (1995) In: Abel EW, Stone FGA, Wilkinson G (eds) Comprehensive organo-
metallic chemistry II, vol 4. Pergamon, Oxford, chap 2
5. Anwander R (1996) Top Curr Chem 179:33
6. Anwander R (1996) Top Curr Chem 179:149
7. For recent reviews, see: (a) Guerriero P, Tamburini S, Vigato PA (1995) Coord Chem
Rev 139:17; (b) Alexander V (1995) Chem Rev 95:273
8. Anwander R, Herrmann WA (1996) Top Curr Chem 179:1
9. Klemm (1929) Z Anorg Chem 184:352
10. (a) Morss LR (1976) Chem Rev 76:827; (b) Morss LR (1994) In: Gschneidner KA Jr, Ey-
ring L, Choppin GR, Lander GH (eds) Handbook on the physics and chemistry of rare
earths. Elsevier, Amsterdam, chap 122
11. Shannon RD (1976) Acta Cryst A32:751
12. (a) Murad E, Hildenbrand DL (1980) J Chem Phys 73:4005; (b) Liu MB, Wahlbeck PG
(1974) High Temp Sci 6:179; (c) Samsonov GV (1982) The oxide handbook, 2nd ed.
IFI/Plenum, New York, pp 86–105
13. Husain M, Batra A, Srivastava KS (1989) Polyhedron 8:1233
14. Freeman AJ, Watson RE (1962) Phys Rev 127:2058
15. Bünzli J-CG, Choppin CR (eds) (1989) Lanthanide probes in life, chemical and earth
sciences, theory and practice. Elsevier, Amsterdam
16. (a) Peacock RD (1975) Struct Bonding (Berl.) 22:83; (b) Carnall WT (1979) In:
Gschneidner KA Jr, Eyring L (eds) Handbook on the physics and chemistry of the rare
earths. North-Holland Publishing Company, Amsterdam, chap 24
17. Moeller T (1980) In: Moeller T, Schleitzer-Rust E (eds) Gmelin handbook of inorganic
chemistry, Sc, Y, La-Lu rare earth elements, part D1, 8th edn, Springer, Berlin Heidel-
berg New York, chap 1
18. Moeller T (1973) In: Bailar JC, Emeleus HJ, Nyholm RS, Trotman-Dickenson AF (eds)
Comprehensive inorganic chemistry, vol 4. Pergamon, Oxford, chap 44
19. (a) Pearson RG (1963) J Am Chem Soc 85:3533; (b) Pearson RG (ed) (1973) Hard and
soft acids and bases. Dowden, Hutchinson, and Ross, Stroudsburg, PA
20. Herrmann WA, Anwander R, Scherer W (1993) Chem Ber 126:1533
21. Raymond KN, Eigenbroth CW Jr (1980) Acc Chem Res 13:276
22. (a) Anwander R (1996) In: Cornils B, Herrmann WA (eds) Applied homogeneous ca-
talysis by organometallic complexes. VCH, Weinheim, p 866; (b) Edelmann FT (1996)
Top Curr Chem 179:247
23. Bradley DC, Hursthouse MB, Aspinall HC, Sales KD, Walker NPC (1985) J Chem Soc
Chem Commun 1585
24. For further literature dealing with the ionic/covalent bonding behavior in lanthanide
complexes, see: (a) Jørgensen CK, Pappalardo R, Schmidtke H-H (1963) J Chem Phys
39:1422; (b) Katzin LI, Barnett ML (1964) J Phys Chem 68:3779; (c) Burns CJ, Bursten
BE (1989) Comments Inorg Chem 9:61
25. Johnson DA (1977) Adv Inorg Chem Radiochem 20:1
26. Cloke FGN (1993) Chem Soc Rev 17
27. (a) Kagan HB (1987) Inorg Chim Acta 140:3; (b) Evans WJ (1987) Inorg Chim Acta
139:169; (c) Soderquist JA (1991) Aldrichim Acta 24:15; (d) Sasaki M, Collin J, Kagan
HB (1992) New J Chem 16:89
52 Reiner Anwander
28. For examples, see: (a) Evans DA, Hoveyda AH (1990) J Am Chem Soc 112:6447; (b)
Yokoo K, Mine N, Taniguchi H, Fujiwara Y (1985) J Organomet Chem 279:C19
29. For a recent theoretical treatment, see: Seth M, Dolg M, Fulde P, Schwerdtfeger P (1995)
J Am Chem Soc 117:6597
30. Thompson LC (1979) In: Gschneidner KA Jr, Eyring L (eds) Handbook on the physics
and chemistry of the rare earths. North-Holland Publishing Company, Amsterdam,
chap 25
31. (a) Herrmann WA, Anwander R, Scherer W, Munck FC (1993) J Organomet Chem
462:163 and references therein; (b) Bel’skii VK, Gun’ko YK, Soloveichik GD, Bulychev
BM (1991) Organomet Chem USSR 4:281
32. Evans WJ (1987) Polyhedron 6:803
33. (a) Bagnell K, Xing-Fu L (1982) J Chem Soc Dalton Trans 1365; (b) Fischer RD, Li X-F
(1985) J Less-Common Met 112:303; (c) Li X-F, Guo A-L (1986) Inorg Chim Acta
134:143
34. Tolman CA (1977) Chem Rev 77:313
35. Bondi A (1964) J Phys Chem 68:441
36. Marcalo J, DeMatos AP (1989) Polyhedron 8:2431
37. Piers WE, Shapiro PJ, Bunel EE, Bercaw JE (1990) Synlett 74
38. Kobayashi S, Moriwaki M, Hachiya I (1995) J Chem Soc Chem Commun 1527 and ref-
erences therein
39. Holton J, Lappert MF, Ballard DGH, Pearce R, Atwood JL, Hunter WE (1979) J Chem
Soc Dalton Trans 54
40. Marks TJ, Gagne MR, Nolan SP, Schock LE, Seyam AM, Stern D (1989) Pure Appl Chem
61:1665
41. (a) Nolan SP, Stern D, Marks TJ (1989) J Am Chem Soc 111:7844; (b) King WA, Marks
TJ, Anderson DM, Duncalf DJ, Cloke FGN (1992) J Am Chem Soc 114:9221; (c) King
WA, Marks TJ (1995) Inorg Chim Acta 229:343
42. Schumann H (1979) Nachr Chem Tech Lab 27:393
43. Sasai H, Suzuki T, Itoh N, Tanaka K, Date T, Okamura K, Shibasaki M (1993) J Am
Chem Soc 115:10372
44. Reger DL, Lindeman JA, Lebioda L (1987) Inorg Chim Acta 139:71
45. Sessler JL, Hemmi G, Mody TD, Murai T, Burrell A, Young SW (1994) Acc Chem Res 27:43
46. Shriver DF, Drezdzon MA (1986) In: The manipulation of air-sensitive compounds,
2nd edn. Wiley, New York
47. Choppin GR, Rizkalla EN (1994) In: Gschneidner KA Jr, Eyring L, Choppin GR, Lander
GH (eds) Handbook on the physics and chemistry of the rare earths. Elsevier, Amster-
dam, chap 128
48. (a) Burgess J, Kijowski J (1981) Adv Inorg Chem Radiochem 24:57; (b) Eick HA (1994)
In: Gschneidner KA Jr, Eyring L, Choppin GR, Lander GH (eds) Handbook on the
physics and chemistry of the rare earths. Elsevier, Amsterdam, chap 124 and referenc-
es therein
49. (a) Reed JB, Hopkins BS, Audrieth LF (1939) Inorg Synth 1:28; (b) Taylor MD (1962)
Chem Rev 62:503; (c) Meyer, G (1989) Inorg Synth 25:146
50. Evans WJ, Feldman JD, Ziller JW (1996) J Am Chem Soc 118:4581
51. Imamoto T (1991) In: Trost BM, Fleming I, Schreiber SL (eds) Comprehensive organic
synthesis. Pergamon, London, chap 1.8
52. Rossmanith K, Auer-Welsbach C (1965) Mh Chem 96:602
53. (a) Evans WJ, Shreeve JL, Ziller JW, Doedens RJ (1995) Inorg Chem 34:576; (b) Willey
GR, Woodman TJ, Drew MGB (1997) Polyhedron 16:3385; (c) Deacon GB, Feng T, Junk
PC, Skelton BW, Sobolev AN, White AH (1998) Aust J Chem 51:75 and references there-
in
54. (a) Freeman JH, Smith ML (1958) J Inorg Nucl Chem 7:224; (b) Manzer LE (1982) Inorg
Synth 21:135
Principles in Organolanthanide Chemistry 53
86. Evans WJ, Deming TJ, Olofson JM, Ziller JW (1989) Inorg Chem 28:4027
87. Jacob K, Thiele K-H (1986) Z Anorg Allg Chem 543:192
88. Imamoto T, Koide Y, Hiyama S (1990) Chem Lett 1445
89. Brown LM, Mazdiyasni KS (1970) Inorg Chem 9:2783
90. (a) Deacon GB, Pain GN, Tuong TD (1989) Inorg Synth 26:17; (b) Deacon GB, Forsyth
CM (1988) Inorg Chim Acta 154:121; (c) Cetinkaya B, Hitchcock PB, Lappert MF,
Smith RG (1992) J Chem Soc Chem Commun 932
91. For an example, see: Wayda AL, Dye JL, Rogers RD (1984) Organometallics 3:1605
92. (a) Mashima K, Sugiyama H, Nakamura A (1994) J Chem Soc Chem Commun 1581; (b)
Mashima K, Nakayama Y, Fukumoto H, Kanehisa N, Kai Y, Nakamura A (1994) J Chem
Soc Chem Commun 2523
93. (a) Fukagawa T, Fujiwara Y, Yokoo K, Taniguchi H (1981) Chem Lett 1771; (b) Hou Z,
Fujiwara Y, Jintoku T, Mine N, Yokoo K, Taniguchi H (1987) J Org Chem 3524
94. Eaborn C, Hitchcock PB, Izod K, Smith JD (1994) J Am Chem Soc 116:12071
95. (a) Brennan JG, Cloke FGN, Sameh AA, Zalkin A (1987) J Chem Soc Chem Commun
1668; (b) Arnold PL, Cloke FGN, Hitchcock PB, Nixon JF (1996) J Am Chem Soc
118:7630; (c) Arnold PL, Cloke FGN, Hitchcock PB (1997) Chem Commun 481; (d) Ar-
nold PL, Cloke FGN, Nixon JF (1998) Chem Commun 797
96. For an example, see: Atwood JL, Lappert MF, Smith RG, Zhang H (1988) J Chem Soc
Chem Commun 1308
97. Minhas RK, Ma Y, Song J-I, Gambarotta S (1996) Inorg Chem 35:1866
98. Edelmann FT, Steiner A, Stalke D, Gilje JW, Jagner S, Håkansson M (1994) Polyhedron
13:539
99. Bradley DC, Ghotra JS, Hart FA (1973) J Chem Soc Dalton Trans 1021
100. Lappert MF, Singh A, Smith RG (1990) Inorg Synth 27:164
101. Hitchcook PB, Lappert MF, Singh A (1983) J Chem Soc Chem Commun 1499
102. Herrmann WA, Anwander R, Kleine M, Scherer W (1992) Chem Ber 125:1971
103. Bochkarev LN, Shustov SB, Guseva TV, Zhil’tsov SF (1988) J Gen Chem USSR 58:819
104. (a) Booij M, Kiers NH, Heeres HJ, Teuben JH (1989) J Organomet Chem 364:79; (b)
Stults SD, Andersen RA, Zalkin A (1990) Organometallics 9:115
105. Mu Y, Piers WE, MacDonald M-A, Zaworotko MJ (1995) Can J Chem 73:2233
106. Razuvaev GA, Kalinina GS, Fedorova EA (1980) J Organomet Chem 190:157
107. Evans WJ, Ansari MA, Ziller JW, Khan SI (1996) Inorg Chem 35:5435
108. Allen M, Aspinall HC, Moore SR, Hursthouse MB, Karvalov AI (1992) Polyhedron
11:409
109. Stecher HA, Sen A, Rheingold AL (1989) Inorg Chem 28:3280
110. McGeary MJ, Coan PS, Folting K, Streib WE, Caulton KG (1989) Inorg Chem 28:3283
111. Barash EH, Coan PS, Lobkovsky EB, Streib WE, Caulton KG (1993) Inorg Chem 32:497
112. (a) Rad’kov YF, Fedorova EA, Khorshev SY, Kalinina GS, Bochkarev MN, Razuvaev GA
(1986) J Gen Chem USSR 55:1911; (b) Cary DR, Arnold J (1993) J Am Chem Soc
115:2520; (c) Strzelecki AR, Likar CL, Helsel BA, Utz T, Lin MC, Bianconi PA (1994) In-
org Chem 33:5188
113. Fedorova EA, Kalinina GS, Bochkarev MN, Razuvaev GA (1982) J Gen Chem USSR
52:1041
114. Herrmann WA, Anwander R, Munck FC, Scherer W, Dufaud V, Huber NW, Artus GRJ
(1994) Z Naturforsch B49:1789
115. Anwander R, Runte O, Eppinger J, Gerstberger G, Herdtweck E, Spiegler M (1998) J
Chem Soc Dalton Trans 847
116. Runte O, Priermeier T, Anwander R (1996) Chem Commun 1385
117. Herrmann WA, Eppinger J, Spiegler M, Runte O, Anwander R (1997) Organometallics
16:1813
118. Görlitzer HW, Spiegler M, Anwander R (1998) Eur J Inorg Chem 1009
119. Aspinall H, Moore SR, Smith AK (1993) J Chem Soc, Dalton Trans 993
Principles in Organolanthanide Chemistry 55
120. Evans WJ, Anwander R, Ziller JW, Khan SI (1995) Inorg Chem 34:5927
121. Mu Y, Piers WE, MacQuarrie DC, Zaworotko MJ, Young VG Jr (1996) Organometallics
15:2720
122. Schumann H, Müller J (1979) J Organomet Chem 169:C1
123. Evans WJ, Drummond DK, Zhang H, Atwood JL (1988) Inorg Chem 27:575
124. Evans WJ, Anwander R, Berlekamp UH, Ziller JW (1995) Inorg Chem 34:3583
125. Evans WJ, Anwander R, Ansari MA, Ziller JW (1995) Inorg Chem 34:5
126. Evans WJ, Fries G, Anwander R, Ziller JW (unpublished results)
127. For an example, see: van den Hende JR, Hitchcock PB, Lappert MF (1994) J Chem Soc
Chem Commun 1413
128. (a) Fischer EO, Fischer H (1964) Angew Chem 76:52; (b) Fischer EO, Fischer H (1965)
J Organomet Chem 3:181
129. Hayes RG, Thomas JL (1969) J Am Chem Soc 91:6876
130. Wayda AL, Mukerji I, Dye JL, Rogers RD (1987) Organometallics 6:1328
131. Murphy E, Toogood GE (1971) Inorg Nucl Chem Lett 7:755
132. Pytlewski LL, Howell JK (1967) Chem Commun 1280
133. Howell JK, Pytlewski LL (1969) J Less-Common Met 18:437
134. (a) White JP III, Deng H-B, Shore SG (1989) J Am Chem Soc 111:8946; (b) White JP III,
Shore SG (1992) Inorg Chem 31:2756
135. Hitchcock PB, Lappert MF, Smith RG, Bartlett RA, Power PP (1988) J Chem Soc Chem
Commun 1007
136. (a) Heeres HJ, Meetsma A, Teuben JH (1988) J Chem Soc Chem Commun 962; (b)
Teuben JH, Rogers RD, Heeres HJ, Meetsma A (1989) Organometallics 8:2637; (c) Scha-
verien CJ, Frijns JHG, Herres HJ, Van den Hende JR, Teuben JH, Spek AL (1991) J Chem
Soc Chem Commun 642
137. (a) Schaverien CJ (1992) J Chem Soc Chem Commun 11; (b) Schaverien CJ (1994) Or-
ganometallics 13:69
138. Clark DL, Gordon JC, Huffmann JC, Watkin JG, Zwick BD (1994) Organometallics
13:4266
139. (a) Evans WJ, Boyle TJ, Ziller JW (1993) Organometallics 12:3998; (b) Evans WJ, Boyle
TJ, Ziller JW (1993) J Organomet Chem 462:141
140. Evans WJ, Boyle TJ, Ziller JW (1993) J Am Chem Soc 115:5084
141. Evans WJ, Ansari MA, Ziller JW (1995) Inorg Chem 34:3079
142. Biagini P, Lugli G, Abis L, Millini R (1994) J Organomet Chem 474:C16
143. (a) Evans WJ, Anwander R, Doedens RJ, Ziller JW (1994) Angew Chem 106:1725; An-
gew Chem Int Ed Engl 36:1641; (b) Evans WJ, Anwander R, Ziller JW (1995) Organo-
metallics 14:1107
144. Fu P-F, Brard L, Li Y, Marks TJ (1995) J Am Chem Soc 117:7157
145. Watson PL, Parshall GW (1985) Acc Chem Res 18:51
146. (a) Evans WJ (1983) J Organomet Chem 250:217; (b) Evans WJ, Dominguez R, Hanusa
TP (1986) Organometallics 5:263; (c) Evans WJ, Drummond DK, Hanusa TP, Doedens
RJ (1987) Organometallics 6:2279
147. Jeske G, Schock LE, Swepston PN, Schumann H, Marks T J (1985) J Am Chem Soc
107:8103
148. Soloveichik GL, Knyazhanskii SY, Bulychev BM, Bel’skii VK (1992) Organomet Chem
USSR 5:73
149. Soloveichik GL (1995) New J Chem 19:597
150. Evans WJ, Meadows JH, Hunter WE, Atwood JL (1984) J Am Chem Soc 106:1291
151. Schumann H, Genthe W, Hahn Ekkehardt, Hossain MB, Van der Helm D (1986) J Or-
ganomet Chem 299:67
152. Hazin PN, Bruno JW, Schulte GK (1990) Organometallics 9:416
153. (a) Evans WJ, Ulibarri TA, Chamberlain LR, Ziller JW, Alvarez Jr D (1990) Organome-
tallics 9:2124; (b) Evans WJ, Seibel CA, Ziller JW (1998) J Am Chem Soc 120:6745
56 Reiner Anwander
154. (a) Heeres HJ, Meetsma A, Teuben JH (1991) J Organomet Chem 414:351; (b) Schaver-
ien CJ (1992) Organometallics 11:3476
155. Anwander R, Görlitzer HW, Runte O, Priermeier T (unpublished results)
156. Amoroso AJ, Jeffery JC, Jones PL, McCleverty JA, Rees L, Rheingold AL, Sun Y, Takats
J, Trofimenko S, Ward MD, Yap GPA (1995) J Chem Soc Chem Commun 1881
157. For reviews, see: (a) Wipf P, Xu W (1995) Tetrahedron 51:4551; (b) Hoveyda AH, Mork-
en JP (1996) Angew Chem 108:1378; Angew Chem Int Ed Eng 35:1263
158. (a) Hubert-Pfalzgraf LG (1987) New J Chem 11:663; (b) Hubert-Pfalzgraf LG (1995)
New J Chem 19:727
159. Anwander R, Munck FC, Priermeier T, Scherer W, Runte O, Herrmann WA (1997) In-
org Chem 36:3545
160. Herrmann WA, Anwander R, Munck FC, Scherer W (1993) Chem Ber 126:331
161. Reynolds LT, Wilkinson G (1959) J Inorg Nucl Chem 9:86
162. Buchler JW, Eikelmann G, Puppe L, Rohbock K, Schneehage HH, Weck D (1971) Lie-
bigs Ann Chem 745:135
163. Bradley DC, Faktor MM (1958) Chem Ind (London) 1332
164. Rausch MD, Morlarty KJ, Atwood JL, Weeks JA, Hunter WE, Brittain HG (1986) Orga-
nometallics 5:1281
165. Calderazzo F, Pappalardo R, Losi S (1966) J Inorg Nucl Chem 28:987
166. For reviews, see: (a) Reetz MT (1984) Angew Chem 96:542; Angew Chem Int Ed Engl
23:556; (b) Hoveyda AH, Evans DA, Fu GG (1993) Chem Rev 93:1307; (c) Berrisford DJ,
Bolm C, Sharpless KB (1995) Angew Chem 107:1159; Angew Chem Int Ed Engl 34:1059
167. (a) Davidson PJ, Lappert MF, Pearce R (1974) Acc Chem Res 7:209; (b) Schumann H
(1985) J Less-Commun Met 112:327
168. For examples, see : (a) Hogerheide MP, Jastrzebski JTBH, Boersma J, Smeets WJJ, Spek
AL, van Koten G (1994) Inorg Chem 33:4431; (b) Kessler VG, Hubert-Pfalzgraf LG, Ha-
lut S, Daran J-C (1994) J Chem Soc Chem Commun 705
169. For average bond lengths, see: Orpen AG, Brammer L, Allen FH, Kennard O, Watson
DG, Taylor R (1989) J Chem Soc Dalton Trans S1
170. Birmingham JM, Wilkinson G (1956) J Am Chem Soc 78:42
171. (a) Ernst RD, Cymbaluk TH (1982) Organometallics 1:708; (b) Schumann H, Dietrich
A (1991) J Organomet Chem 401:C33; (c) Baudry D, Nief F, Ricard L (1994) J Organom-
et Chem 482:125
172. (a) Evans WJ, Hughes LA, Hanusa TP (1986) Organometallics 5:1285; (b) Evans WJ,
Forrestal KJ, Leman JT, Ziller JW (1996) Organometallics 15:527
173. Atwood JL, Burns JH, Laubereau PG (1973) J Am Chem Soc 95:1830
174. (a) Huang Z, Chen M, Qiu W, Wu W (1987) Inorg Chim Acta 139:203; (b) Taube R,
Windisch H, Görlitz FH, Schumann H (1993) J Organomet Chem 445:85
175. Ghotra JS, Hursthouse MB, Welch AJ (1973) J Chem Soc Chem Commun 669
176. Wedler M, Knösel F, Pieper U, Stalke D, Edelmann FT, Amberger H-D (1992) Chem Ber
125:2171
177. Hitchcock PB, Holmes SA, Lappert MF, Tian S (1994) J Chem Soc Chem Commun 2691
178. Buchler JW, De Cian A, Fischer J, Kihn-Botulinski M, Paulus H, Weiss R (1986) J Am
Chem Soc 108:3652
179. Stecher H, Sen A, Rheingold AL (1988) Inorg Chem 27:1130
180. Hitchcook PB, Lappert MF, Mackinnon IA (1988) J Chem Soc Chem Commun 1557
181. De Villiers JPR, Boeyens JCA (1972) Acta Cryst B28:2335
182. Terzis A, Mentzafos D, Tajmir-Riahi H-A (1984) Inorg Chim Acta 84:187
183. (a) Rabe GW, Riede J, Schier A (1996) Inorg Chem 35:40; (b) Rabe GW, Riede J, Schier
A (1996) Inorg Chem 35:2680
184. Tatsumi K, Amemiya T, Kawaguchi H, Tani K (1993) J Chem Soc Chem Commun 773
185. For a review, see: Hogerheide MP, Boersma J, van Koten G (1996) Coord Chem Rev
155:87
Principles in Organolanthanide Chemistry 57
186. Haaland A (1989) Angew Chem 101:1017; Angew Chem Int Ed Engl 28:992
187. For a review, see: Bader A, Lindner E (1991) coord Chem Rev 108:27
188. Gruter G-JM, van Klink GPM, Akkerman OS, Bickelhaupt F (1995) Chem Rev 95:2405
189. Schumann H, Meese-Marktscheffel JA, Dietrich A, Pickardt J (1992) J Organomet
Chem 433:241
190. Herbrich T, Thiele K-H, Thewalt U (1996) Z Anorg Allg Chem 622:1609
191. Clegg W, Eaborn C, Izod K, O’Shaughnessy P, Smith JD (1997) Angew Chem 109:2925;
Angew Chem Int Ed Eng 36:2815
192. Wayda AL, Rogers RD (1985) Organometallics 4:1440
193. (a) Hogerheide MP, Grove DM, Boersma J, Jastrzebski JTBH, Kooijman H, Spek AL,
van Koten G (1995) Chem Eur J 1:343; (b) Hogerheide MP, Boersma J, Spek AL, van Ko-
ten G (1996) Organometallics 15:1505
194. Schumann H, Lee PR, Loebel J (1989) Chem Ber 122:1897
195. Fryzuk MD (1992) Can J Chem 70:2839
196. Deacon GB, Forsyth CM, Junk PC, Skelton BW, White AH (1998) J Chem Soc Dalton
Trans 1381
197. Herrmann WA, Anwander R, Denk M (1992) Chem Ber 125:2399
198. Shao P, Berg DJ, Bushnell GW (1994) Inorg Chem 33:3452
199. Karsch HH, Ferazin G, Steigelmann O, Kooijman H, Hiller W (1993) Angew Chem
105:1814; Angew Chem Int Ed Engl 32:1739
200. (a) Edelmann FT (1994) J Alloys Compounds 207/208:182; (b) Edelmann FT (1996)
Top Curr Chem 179:113
201. (a) Wang B, Deng D, Qian C (1995) New J Chem 19:515 and references therein; (b)
Zhang S, Zhung X, Wei G, Chen W, Liu J (1994) Polyhedron 13:2867; (c) Van den Hende
JR, Hitchcock PB, Lappert MF, Nile TA (1994) J Organomet Chem 472:79; (d) Jutzi P,
Dahlhaus J, Kristen MO (1993) J Organomet Chem 450:C1; (e) Schumann H, Erbstein
F, Weimann R, Demtschuk J (1997) J Organomet Chem 536:541
202. (a) Schumann H, Görlitz FH, Hahn FE, Pickardt J, Qian C, Xie Z (1992) Z Anorg Allg
Chem 609:131; (b) Qian C, Zhu D (1994) J Chem Soc Dalton Trans 1599; (c) Gräper J,
Fischer RD, Paolucci G (1994) J Organomet Chem 471:87; (d) Qian C, Zhu D (1993) J
Organomet Chem 445:79; (e) Paolucci G, D’Ippolito R, Ye C, Qian C, Gräper J, Fischer
RD (1994) J Organomet Chem 471:97
203. Roesky PW, Stern CL, Marks TJ (1997) Organometallics 16:4705
204. (a) Van de Weghe P, Bied C, Collin J, Marcalo J, Santos I (1994) J Organomet Chem
475:121; (b) Molander GA, Schumann H, Rosenthal ECE, Demtschuk J (1996) Organo-
metallics 15:3817; (c) Trifonov AA, Van de Weghe P, Collin J, Domingos A, Santos I
(1997) J Organomet Chem 527:225
205. (a) Schumann H, Meese-Markstscheffel JA, Gorella B, Görlitz FH (1992) J Organomet
Chem 428:C27; (b) Deacon GB, Fallon GD, Forsyth CM (1993) J Organomet Chem
462:183; (c) Broussier R, Delmas G, Perron P, Gautheron B, Petersen JL (1996) J Orga-
nomet Chem 511:185; (d) Lin G, Wong W-T (1996) J Organomet Chem 523:93
206. Xia J, Zhuang X, Jin Z, Chen W (1996) Polyhedron 15:3399
207. For an example, see: Giardello MA, Conticello VP, Brard L, Sabat M, Rheingold AL,
Stern CL, Marks TJ (1994) J Am Chem Soc 116:10212
208. For an example, see: Tsukube H, Shiba H, Uenishi J-I (1995) J Chem Soc Dalton Trans 181
209. For recent examples, see: (a) Tsuie B, Swenson DC, Jordan RF, Petersen JL (1997) Organome-
tallics 16:1392; (b) Scollard JD, McConville DH, Vittal JJ (1997) Organometallics 16:4415; (c)
Gibson VC, Kimberley BS, White AJP, Williams DJ, Howard P (1998) Chem Commun 313; (d)
Schrock RR, Schattenmann F, Aizenberg M, Davis WM (1998) Chem Commun 199
210. For recent examples, see: (a) Fokken S, Spaniol TP, Okuda J, Sernetz FG, Mülhaupt R
(1997) Organometallics 16:4240; (b) Repo T, Klinga M, Pietikäinen P, Leskelä M, Uusi-
talo A-M, Pakkanen T, Hakala K, Aaltonen P, Löfgren B (1997) Macromolecules 30:171
58 Reiner Anwander
211. Görls H, Neumüller B, Scholz A, Scholz J (1995) Angew Chem 107:732; Angew Chem
Int Ed Eng 34:673
212. Bürgstein MR, Berberich H, Roesky PW (1998) Organometallics 17:1452
213. Duchateau R, van Wee CT, Meetsma A, Teuben JH (1993) J Am Chem Soc 115:4931
214. (a) Long DP, Bianconi PA (1996) J Am Chem Soc 118:12453; (b) Santos I, Marques N
(1995) New J Chem 19:551
215. (a) Duchateau R, van Wee CT, Teuben JH (1996) Organometallics 15:2291; (b) Ducha-
teau R, Tuinstra T, Brussee EAC, Meetsma A, van Duijnen PT, Teuben JH (1996) Orga-
nometallics 16:3511
216. Covert KJ, Neithamer DR, Zonnevylle MC, LaPointe RE, Schaller CP, Wolczanski PT
(1991) Inorg Chem 30:2494
217. Shao P, Berg DJ, Bushnell GW (1994) Inorg Chem 33:6334
218. (a) Giardello MA, Conticelli VP, Brard L, Gagne MR, Marks TJ (1994) J Am Chem Soc
116:10241; (b) Mitchell JP, Hajela S, Brookhart SK, Hardcastle KI, Henling LM, Bercaw
JE (1996) J Am Chem Soc 118:1045
219. Hultzsch KC, Spaniol TP, Okuda J (1997) Organometallics 16:4845
220. Chauvin Y, Heyworth S, Olivier H, Robert F, Saussine L (1993) J Organomet Chem
455:89
221. Shah SAA, Dorn H, Roesky HW, Lubini P, Schmidt H-G (1997) Inorg Chem 36:1102
222. Lee L, Berg DJ, Einstein FW, Batchelor RJ (1997) Organometallics 16:1819
223. Schaverien CJ, Meijboom N, Orpen AG (1992) J Chem Soc Chem Commun 124
224. Sasai H, Arai T, Satow Y, Houk KN, Shibasaki M (1995) J Am Chem Soc 117:6194
225. Edelmann FT (1995) New J Chem 19:535
226. Arnold J, Hoffman CG, Dawson DY, Hollander FJ (1993) Organometallics 12:3645
227. (a) Oki AR, Zhang H, Hosmane NS (1991) Organometallics 10:3964; (b) Hosmane NS,
Wang Y, Zhang H, Maguire JA, McInnis M, Gray TG, Collins JD, Kremer RK, Bider H,
Waldhör E, Kaim W (1996) Organometallics 15:1006
228. (a) Manning MJ, Knobler CB, Hawthorne MF (1988) J Am Chem Soc 110:4458; (b) Ba-
zan GC, Schaefer WP, Bercaw JE (1993) Organometallics 12:2126; (c) Xie Z, Liu Z, Chiu
K, Xue F, Mak TCW (1997) Organometallics 16:2460
229. For an example, see: Bednarski M, Danishefsky (1986) J Am Chem Soc 108:7060
230. Xie Z, Wang S, Zhou Z-Y, Mak TCW (1998) Organometallics 17:1907
231. (a) Aspinall HC, Tillotson MR (1996) Inorg Chem 35:2163; (b) Roussel P, Alcock NW,
Scott P (1998) Chem Commun 801
232. Spino C, Clouston L, Berg DJ (1996) Can J Chem 74:1762
233. Herrmann WA, Anwander R, Dufaud V, Scherer W (1994) Angew Chem 106:1338; An-
gew Chem Int Ed Engl 33:1285
234. Feher FJ, Budzichowski TA (1995) Polyhedron 14:3239
235. (a) Robinson AL (1976) Science 194:1261; (b) Sachtler WMH, Zhang Z (1993) Adv Catal
39:129; (c) Marks TJ (1992) Acc Chem Res 25:57
236. (a) Bergbreiter DE, Chen L-B, Chandran R (1985) Macromolecules 18:1055; (b) Yu G,
Li Y, Qu Y, Li X (1993) Macromolecules 26:6702
237. (a) Kobayashi S, Nagayama S (1996) J Org Chem 61:2256; (b) Kobayashi S, Nagayama
S (1998) J Am Chem Soc 120:2985
238. Kobayashi S, Nagayama S (1996) J Am Chem Soc 118:8977
239. Keller F, Weinmann H, Schurig V (1997) Chem Ber/Recueil 130:879
240. Anwander R, Roesky R (1997) J Chem Soc Dalton Trans 137
241. Anwander R, Palm C (1998) Stud Surf Sci Catal 117:413
242. (a) Imamura H, Konishi T, Sakata Y, Tsuchiya S (1991) J Chem Soc Chem Commun
1527; (b) Imamura H, Konishi T, Sakata Y, Tsuchiya S (1992) J Chem Soc Faraday Trans
88:2251; (c) Imamura H, Konishi T, Sakata Y, Tsuchiya S (1993) J Chem Soc Chem Com-
mun 1852; (d) Imamura H, Suda E, Konishi T, Sakata Y, Tsuchiya S (1995) Chem Lett
215 and references therein
Principles in Organolanthanide Chemistry 59
243. (a) Baba T, Koide R, Ono Y (1991) J Chem Soc Chem Commun 691; (b) Baba T, Kim GJ,
Ono Y (1992) J Chem Soc Faraday Trans 88:891; (c) Baba T, Hikita S, Koide R, Ono Y,
Hanada T, Tanaka T, Yoshida S (1993) I Chem Soc Faraday Trans 89:3177
244. Takaki K, Fujiwara Y (1990) Appl Organomet Chem 4:311
245. Bradley DC, Ghotra JS, Hart FA, Hursthouse MB, Raithby PR (1977) J Chem Soc Dalton
Trans 1166
246. Aspinall HC, Moore SR, Smith AK (1992) J Chem Soc Dalton Trans 153
247. Barnhart DM, Clark DL, Gordon JC, Huffman JC, Vincent RL, Watkin JG, Zwick BD
(1994) Inorg Chem 33:3487
248. (a) Schumann H, Glanz M, Winterfeld J, Hemling H, Kuhn N, Kratz T (1994) Angew
Chem 106:1829; Angew Chem Int Ed Engl 33:1733; (b) Arduengo AJ III, Tamm M,
McLAin SJ, Calabrese LC, Davidson F, Marshall WJ (1994) J Am Chem Soc 116:7927; (c)
Herrmann WA, Munck FC, Artus GRJ, Runte O, Anwander R (1997) Organometallics
16:682
249. (a) Shapiro PJ, Bunel E, Schaefer WP, Bercaw JE (1990) Organometallics 9:867; (b) Sha-
piro PJ, Cotter WD, Schaefer WP, Labinger JA, Bercaw JE (1994) J Am Chem Soc
116:4623
250. Daniele S, Hubert-Pfalzgraf LG, Vaissermann J (1995) Polyhedron 14:327
251. Burns CJ, Andersen RA (1987) J Organomet Chem 325:31
252. Stults SD, Andersen RA, Zalkin A (1990) Organometallics 9:115
253. Anwander R, Palm C, Groeger O, Engelhardt G (1998) Organometallics 17:2027
254. van der Heijden H, Schaverien CJ, Orpen AG (1989) Organometallics 8:255
255. Guan J, Jin S, Lin Y, Shen Q (1992) Organometallics 11:2483
256. Hogerheide MP, Jastrzebski JTBH, Boersma J, Smeets WJJ, Spek AL, van Koten G
(1994) Inorg Chem 33:4431
257. For an example, see: Haar MD, Stern CL, Marks TJ (1996) Organometallics 15:1765
258. Herrmann WA, Anwander R, Riepl H, Scherer W, Withaker CR (1993) Organometallics
12:4342
259. Evans WJ, Shreeve JL, Broomhall-Dillard RNR, Ziller JW (1995) J Organomet Chem 501:7
260. Schumann H, Genthe W, Bruncks N, Pickardt J (1982) Organometallics 1:1194
261. Deacon GB, Shen Q (1996) J Organomet Chem 511:1
262. Burns CJ, Andersen RA (1987) J Am Chem Soc 109:915
263. Casey CP, Hallenbeck SL, Pollock DW, Landis CR (1995) J Am Chem Soc 117:9770
264. Burns CJ, Andersen RA (1987) J Am Chem Soc 109:941
265. (a) Cotton FA, Schwotzer W (1986) J Am Chem Soc 108:4657; (b) Fan B, Shen Q, Lin Y
(1989) J Organomet Chem 376:61
266. Nolan SP, Marks TJ (1989) J Am Chem Soc 111:8538
267. Evans WJ, Ulibarri TA, Ziller JW (1988) J Am Chem Soc 110:6877
268. (a) Evans WJ, Kociok-Köhn G, Ziller JW (1992) Angew Chem 104:1114; Angew Chem
Int Ed Engl 31:1081; (b) Evans WJ, Kociok-Köhn G, Leong VS, Ziller JW (1992) Inorg
Chem 31:3592
269. (a) Brookhart M, Green MLH (1983) J Organomet Chem 250:395; (b) Brookhart M,
Green MLH, Wong L-L (1988) Prog Inorg Chem 36:1
270. Ziegler T, Folga E, Berces A (1993) J Am Chem Soc 115:636
271. Schaverien CJ, Nesbitt GJ (1992) J Chem Soc Dalton Trans 157
272. Piers WE, Bercaw JE (1990) J Am Chem Soc 112:9406
273. Burger BJ, Thompson ME, Cotter WD, Bercaw JE (1990) J Am Chem Soc 112:1566
274. For examples, see: (a) Jeske G, Lauke H, Mauermann H, Swepston PN, Schumann H,
Marks TJ (1985) J Am Chem Soc 107:8091; (b) van der Heijden H, Pasman P, de Boer
EJM, Schaverien CJ, Orpen AG (1989) Organometallics 8:1459
275. Klooster WT, Lu RS, Anwander R, Evans WJ, Koetzle TF, Bau R (1998) Angew Chem
110:1326; Angew Chem Int Ed Engl 37:1268
60 Reiner Anwander
276. Barnhart DM, Clark DL, Gordon JC, Huffmann JC, Watkin JG, Zwick BD (1993) J Am
Chem Soc 115:8461
277. Tilley TD, Andersen RA, Zalkin A (1982) J Am Chem Soc 104:3725
278. Evans WJ, Grate JW, Choi HW, Bloom I, Hunter WE, Atwood JL (1985) J Am Chem Soc
107:941
279. Anwander R (1998) Angew Chem 110:619; Angew Chem Int Ed Engl 37:599
280. Obora Y, Ohta T, Stern CL, Marks TJ (1997) J Am Chem Soc 119:3745
281. Evans WJ, Drummond DK (1988) Organometallics 7:797
282. Watson PL, Tulip TH, Williams I (1990) Organometallics 9:1999
283. Kretschmer WP, Teuben JH, Troyanov SI (1998) Angew Chem 110:92; Angew Chem Int
Ed Eng 37:88
284. Poncelet O, Hubert-Pfalzgraf LG, Daran J-C, Astier R (1989) J Chem Soc Chem Com-
mun 1846
285. Mehrotra RC, Singh A (1996) Chem Soc Rev 1
286. (a) Aime S, Crich SG, Gianolio E, Terreno E, Beltrami A, Uggeri F (1998) Eur J Inorg
Chem 1283; (b) Aime S, Botta M, Fasano M, Terreno E (1998) Chem Soc Rev 27:19
287. Watson PL (1983) J Am Chem Soc 105:6491
288. Thompson ME, Baxter SM, Bulls AR, Burger BJ, Nolan MC, Santarsiero BD, Schaefer
WP, Bercaw JE (1987) J Am Chem Soc 109:203
289. Booij M, Meetsma A, Teuben JH (1991) Organometallics 10:3246
290. Bradley DC, Hursthouse MB, Aspinall HC, Sales KD, Walker NPC, Hussain B (1989) J
Chem Soc Dalton Trans 623
291. Watson PL (1982) J Am Chem Soc 104:337
292. (a) Cagne MR, Stern CL, Marks TJ (1992) J Am Chem Soc 114:275; (b) Li Y, Fu P-F,
Marks TJ (1994) Organometallics 13:439
293. Evans WJ, Wayda AL, Hunter WE, Atwood JL (1981) J Chem Soc Chem Commun 706
294. Clair MA St, Santarsiero BD, Bercaw JE (1989) Organometallics 8:17
295. Evans WJ, Forrestal KJ, Ziller JW (1995) J Am Chem Soc 117:12635
296. Campion BK, Heyn RH, Tilley TD (1990) J Am Chem Soc 112:2011
297. Watson PL, Roe DC (1982) J Am Chem Soc 104:6471
298. Fryzuk MD, Haddad TS, Rettig SJ (1991) Organometallics 10:2026
299. Bradley DC, Chudzynska H, Hammond ME, Hursthouse MB, Motevalli M, Wu R
(1992) Polyhedron 11:375
300. Poncelet O, Hubert-Pfalzgraf LG, Daran J-C (1990) Polyhedron 9:1305
301. Kagan HB, Sasaki M, Collin J (1988) Pure Appl Chem 60:1725
302. Evans WJ, Grate JW, Bloom I, Hunter WE, Atwood JL (1985) J Am Chem Soc 107:405
303. Zhang X, Loppnow GR, McDonald R, Takats J (1995) J Am Chem Soc 117:7828
304. Evans WJ (1987) In: Suslick KS (ed) High Energy Processes in Organometallic Chem-
istry, ACS Symposium Series 333, American Chemical Society, Washington, DC, p 278
305. (a) Evans WJ, Hughes LA, Drummond DK, Zhang H, Atwood JL (1986) J Am Chem Soc
108:1722; (b) Evans WJ, Drummond DK (1986) J Am Chem Soc 108:7440
306. Hou Z, Wakatsuki Y (1997) Chem Eur J 3:1005
307. Hou Z, Yamazaki H, Kobayashi K, Fujiwara Y, Taniguchi H (1992) Organometallics 11:2711
308. Makioka Y, Taniguchi Y, Fujiwara Y, Takaki K, Hou Z, Wakatsuki Y (1996) Organome-
tallics 15:5476
309. Hou Z, Fujita A, Zhang Y, Miyano T, Yamazaki H, Wakatsuki Y (1998) J Am Chem Soc
120:754
310. Cassani MC, Lappert MF, Laschi F (1997) Chem Commun 1563
311. Evans WJ, Deming TJ, Ziller JW (1989) Organometallics 8:1581
312. Sen A, Stecher HA, Rheingold AL (1992) Inorg Chem 31:473
313. Sasai H, Arai T, Shibasaki M (1994) J Am Chem Soc 116:1571
314. (a) Yasuda H, Tamai H (1993) Prog Polym Sci 18:1097; (b) Yasuda H, Yamamoto H,
Yokota K, Miyake S, Nakamura A (1992) J Am Chem Soc 114:4908
Principles in Organolanthanide Chemistry 61
Although the polymerization prowess of organolanthanide complexes has been known for
some time, efforts to apply these catalysts to small molecule synthesis have only recently be-
gun. The selectivity of these metallocenes is predominantly steric in nature, and they are
compatible with a wide variety of organic functional groups. A review of their use in olefin
hydrogenation, hydrosilylation, and polyene cyclization with emphasis on chemoselectivity
and diastereoselectivity is presented here. The various ways in which the catalysts and rea-
gents can be tuned to produce the desired products is also discussed.
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
7 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
120 Gary A. Molander, Eric D. Dowdy
1
Introduction
Carbon-carbon bond-forming reactions constitute the heart and soul of syn-
thetic organic chemistry. Nowhere are these reactions more prolific than in com-
mercial polymerization reactions, where worldwide production of polyethylene
and polypropylene is carried out on enormous scale. The polymerization of α-
olefins employing Ziegler-Natta catalysts [1], organolanthanides and group 3 or-
ganometallics [2], palladium and nickel cationic complexes [3], and related cat-
alysts [4] have revolutionized the controlled synthesis of polymers derived from
terminal alkenes. Group 3 organometallics and organolanthanides in particular
are among the most active known catalysts in olefin polymerization chemistry
[5]. However, in spite of the spectacular ability of these complexes to generate
new carbon-carbon bonds in polymerization reactions, relatively little effort has
been made in applying these catalytic systems to small molecule synthesis
through cyclization reactions of dienes, enynes, and related substrates
LnCp*2 FG
cat. "Cp*2LnH"
n( ) (1)
n( ) n( )
2
General Features of the Catalytic Systems and the Olefin Insertion Reaction
The key step in nearly all of the catalytic processes to be discussed is olefin in-
sertion into a metal hydride [Eq. (2)] or organometallic species [Eq. (3)]. These
hydrometallation and carbometallation processes also form the basis for the po-
lymerization of alkenes. Olefin insertions generally occur with the same regiose-
lectivity as hydroboration reactions [9], with the bulky metal and associated lig-
ands residing at the least hindered site of the two carbon reactive unit.
In most instances these reactions are extremely exothermic and effectively ir-
reversible under reasonable reaction conditions [10]. Even tert-butyl-substitut-
ed organolanthanides (prepared by other methods) exhibit significant kinetic
stability [11].
The organolanthanide- and group 3 organometallic catalysts are highly elec-
trophilic species. Although there is a significant electronic driving force in the
olefin insertion process, for the most part steric factors predominate. Much of
this can be attributed to the effects of coordinative unsaturation at the metal
center. High reactivity is associated with free coordination sites and terminal
(non-bridging) ligands [12]. The requirement for free coordination sites dictates
that noncoordinating solvents be utilized for the catalytic reactions. Thus com-
mon ether solvents lead to low catalytic turnover rates [10] and can even deacti-
vate catalysts via ether cleavage reactions [7, 13]. Consequently, hydrocarbon
solvents are used exclusively in the catalytic reactions.
Although it would seem logical to utilize the most sterically unhindered lig-
ands about the metal to achieve maximum reactivity with hindered alkenes, in
fact there is a delicate balance that must be achieved between the “openness” of
the metal center and the tendency for the organometallic hydrides to undergo a
deactivating ligand redistribution [14] or to dimerize, forming hydride-bridged
dimers [Eq. (4)] [10].
H
2 Cp2LnH Cp2Ln LnCp2 (4)
H
122 Gary A. Molander, Eric D. Dowdy
TMS TMS
Me
Ln Ln = [(CpTMS)2LnMe]2
Me
TMS TMS
Ln = Sm, Y, Lu
Me2Si SmCH(TMS)2
provides more facile approach of substrate to the metal. Figure 2 depicts one
such precatalyst [20].
With a wide range of metals and ligands available [2a, 21], organolanthanide
and group 3 organometallic catalysts are readily “tuned” to provide the desired
reactivity and selectivity patterns in reactions of interest.
3
Catalytic Hydrogenation Reactions
Organolanthanides and group 3 organometallics are extraordinarily reactive hy-
drogenation catalysts [22]. They are also relatively challenging to prepare and
handle. This, combined with their lack of tolerance to reactive functional
groups, makes them less attractive than many transition metal-based catalysts
in standard hydrogenation reactions. It is instructive, however, to examine the
catalytic cycle of hydrogenation reactions because it serves as a useful model for
the other catalytic reactions to be discussed subsequently. Additionally, useful
information concerning selectivity has been derived from studies of the hydro-
genation reaction.
Transition metal-based hydrogenation reactions most often operate on a cat-
alytic cycle that involves oxidative addition, olefin insertion, and reductive elim-
ination. The mechanistic basis for organolanthanide hydrogenation is quite dif-
ferent, and involves olefin insertion and σ-bond metathesis (Fig. 3).
A σ-bond metathesis initiates the process, converting an organometallic pre-
catalyst into the catalytic hydride species through a four-centered exchange of
124 Gary A. Molander, Eric D. Dowdy
Cp*2LnR'
H2
- R'H
H
"Cp*2LnH"
H
R R
H2 H
LnCp*2
R
Fig. 3. Catalytic cycle for hydrogenation
ligands [Eq. (6)] [23]. Olefin insertion is the product-determining step of the
process, and is generally irreversible under optimal conditions for hydrogena-
tion [10]. A σ-bond metathesis reaction constitutes the final step of the process,
wherein the organometallic reacts with hydrogen to release the hydrocarbon
and regenerate the active catalyst.
Cp*2Ln R
Cp*2LnR + H2 Cp*2LnH + RH (6)
H H
Me Me 2% Cp*2YMe·THF Me Me
R R
1 atm H2, benzene, rt, 1 h (9)
R = OTBS, OMe, OBn; yield = 74-99%
R = OAc, Cl; yield = 0%
10% Cp*2SmCH(TMS)2
1 atm H2 (10)
cyclopentane, rt
84%
3% Cp*2SmCH(TMS)2
R R
1 atm H2, cyclopentane
R = Ph, (CH2)3NMe2, 50 ˚C (11)
76-96%, >10 : 1 ds
R = OMe, 70 ˚C, 0%
3% Cp*2YbCH(TMS)2
1 atm H2 (12)
cyclopentane, -20 ˚C
t-Bu 73%, 3.3 : 1 ds t-Bu
0.7% cat. H
1 atm. H2
heptane, rt, <1 h
95%, 71% ee
(13)
R*
Sm H
Me2Si Sm H
R* Ph
Ph
R*
Lateral trajectory model Frontal trajectory model
4
Catalytic Hydrosilylation Reactions
Metal catalyzed hydrosilylation reactions provide the most efficient and eco-
nomical route to organosilanes [29]. Organosilanes can be readily oxidized to
the corresponding alcohols utilizing a Tamao procedure or related reactions
[30]. The overall process thus constitutes the synthetic equivalent of a catalytic
hydroboration/oxidation sequence [31]. One advantage of the silylation/oxida-
tion protocol is the stability of the intermediate silane which allows the unmask-
ing of the alcohol to be performed at a synthetically convenient time. In spite of
the potential of hydrosilylation reactions in selective organic synthesis, relative-
ly little effort has been made to develop procedures for the selective hydrosilyla-
tion of polyfunctional alkenes and alkynes, especially when compared to the
analogous hydroboration reaction.
Organolanthanide and group 3 organometallic catalysts provide an alterna-
tive to the more traditional platinum-based catalysts for the selective hydrosi-
lylation of alkenes and alkynes. Mechanistically, the transformation is analo-
gous to the catalytic hydrogenation reaction detailed previously [22]. When si-
lane is utilized in place of hydrogen, the σ-bond metathesis occurs to place the
silane moiety on the alkyl unit, and the organometallic hydride is again regen-
erated (Fig. 5). In the overall process a rapid, exothermic, and essentially irre-
versible olefin insertion is followed by the slower, rate-determining σ-bond me-
tathesis. Because the group 3 metallocenes and organolanthanides are highly ef-
fective catalysts for dehydrogenative polysilylation [32], the desired process de-
mands that olefin insertion and σ-bond metathesis occur much more rapidly
than the generation of polysilanes.
Cp*2LnR'
R3SiH
- R'SiR3
H
SiR3 "Cp*2LnH"
R"
R"
R3SiH H
LnCp*2
R"
4.1
Silylation of Alkenes
5% Cp2LuC6H4CH3•THF
PhMeHSi
n-C6H13 n-C6H13 +
MePhSiH2, 90 ˚C, 2 d
75%, 28 : 1
(14)
SiHMePh
+
n-C6H13
4% Cp*2NdCH(TMS)2 SiH2Ph
+ PhH2Si
Ph Ph
PhSiH3, 80 ˚C, 2 d Ph (15)
88%, 4.5 : 1
The regioselectivity of olefin insertion varies with the complex used in the re-
action [27, 34]. In the hydrosilylation of a monosubstituted olefin, the use of
complexes with larger metals and more open ligands provide increased yields of
the product derived from reversed (“2,1”) insertion (Eq. 16). These results re-
veal that a variety of complexes give excellent selectivity for terminal insertion,
but the conditions to elevate the amount of “2,1” insertion remain elusive.
5% catalyst SiH2Ph
PhH2Si
n-C8H17 n-C8H17 +
PhSiH3, rt, 1-24 h n-C8H17
Catalyst Yield (% isolated) ds
Cp*2LuMe•THF 98 100 : 0
Cp*2YbCH(TMS)2 91 100 : 0 (16)
Cp*2YMe•THF 84 100 : 0
Cp*2SmCH(TMS)2 90 11 : 1
Cp*2NdCH(TMS)2 85 3.2 : 1
Cp*2LaCH(TMS) 2 90 1.9 : 1
Me 2SiCp"2YCH(TMS)2 84 31 : 1
Me 2SiCp"2SmCH(TMS) 2 98 1:2
Me 2SiCp"2NdCH(TMS)2 89 1:2
organosilanes (Eq. 17). It was also demonstrated that common organic func-
tional groups (halides, ethers, and acetals) could be tolerated under the reaction
conditions (Eq. 18).
3% Cp*2YCH(TMS)2
SiH2Ph (17)
PhSiH3, benzene
rt, 3.5 h, 85% H
3% Cp*2YCH(TMS)2
SiH2Ph
X X
PhSiH3, benzene (18)
rt, 83-96%
X = Cl, OBn, OTBDMS, OTHP
In the same study, remarkable steric selectivity was demonstrated in the hy-
drosilylation of dienes with varying substitution patterns [35]. Virtually com-
plete selectivity was observed for the reaction of monosubstituted olefins in the
presence of disubstituted alkenes (Eqs. 19, 20) and for the silylation of a 1,1-di-
substituted olefin in preference to a trisubstituted double bond (Eq. 21).
3% Cp*2YCH(TMS)2
3% Cp*2YCH(TMS)2 SiH2Ph
PhSiH3, toluene, rt (20)
18 h, 97%
3% Cp*2YCH(TMS)2 SiH2Ph
PhSiH3, benzene, rt (21)
96 h, 61%, 2:1 ds
The yttrium catalysts are less effective for more sterically hindered olefins,
but the flexibility afforded by being able to alter both the metal and the ligand
system provides a means to adjust reactivity in a manner that allows hydrosilyla-
tion of more highly substituted alkenes. The simple modification of increasing
the ionic radius of the metal permits the hydrosilylation of 1,1-disubstituted
alkenes (Eq. 22) [25]. This effect dominates over slight ligand modifications, as
a complex with more hindered ligands (C5Me4i-Pr) (Eq. 23) shows similar reac-
tivity to the C5Me5-derived yttrium complex (Eq. 21) [36]. A silicon-hinged cat-
alyst further increases turnover frequency over nonbridged systems by a factor
of eight (Eq. 24) [27]. Unfortunately, there is a trade-off with the more open cat-
alysts. Although increased reactivity with more highly substituted alkenes is ob-
served, monosubstituted alkenes react with poor regioselectivity.
130 Gary A. Molander, Eric D. Dowdy
SiH2Ph
5% Cp*2SmCH(TMS)2
R
R
PhSiH3, cyclohexane (22)
70 ˚C, 12 h
63-92%, 9 : 1 ds
R = Me, i-Bu, Bn, (CH2)3NMe2
5% (C5Me4i-Pr)2SmCH(TMS)2 SiH2Ph
PhSiH3, cyclohexane (23)
rt, 96 h, 82%
Styrene derivatives react with “2,1-” regioselectivity [27]. This reversal of se-
lectivity varies considerably with the metal ionic radius and the ligand array
present, with larger metals and bridged ligands giving higher ratios of the “2,1”
product. As with the catalytic hydrogenation of styrene derivatives discussed
previously, the olefin insertion reaction defines the regiochemistry and stereo-
chemistry of the final product. Thus the olefin insertion is essentially irreversi-
ble under the reaction conditions, and the σ-bond metathesis presumably oc-
curs with retention of configuration to provide the observed products with re-
markably high ee’s considering the overall nature of the transformation (Eq. 25).
(25)
R*
4.2
Silylation of Alkynes
the alkyne is observed (Eq. 26). Branching at one of the propargyl positions is
necessary for high regioselectivity in unsymmetrical alkynes. A variety of
branched substituents are suitable for use (Eqs. 27–31) widening the possibili-
ties of subsequent synthetic steps. Placing a tertiary group on the alkyne (Eq. 30)
slows the reaction, allowing the competitive dehydrogenative polymerization of
silane to lower the yield of the desired product.
5% (C5Me4i-Pr)2YCH(TMS)2 n-C6H13
n-C6H13 +
PhSiH3, 50 ˚C
12 h, 78%, 4 : 1 PhH2Si H
(26)
n-C6H13
+
H SiH2Ph
5% Cp*2YMe•THF
PhSiH3, cyclohexane
50 ˚C, 24 h, 80%
(27)
SiH2Ph
O
O 5% Cp*2YMe·THF O (28)
n-C5H11 n-C5H11
O PhSiH3, 50 ˚C
24 h, 85% H SiH2Ph
TBDMSO 5% Cp*2YMe•THF
n-C10H21 PhSiH3, cyclohexane
90 ˚C, 24 h, 89%
(29)
TBDMSO n-C10H21
SiH2Ph
5% Cp*2YMe·THF
(31)
PhSiH3, 90 ˚C
7 d, 50% H SiH2Ph
5% Cp*2YMe•THF
X
PhSiH3, cyclohexane
50 ˚C, 24 h, 73-84%
X = Cl, OTHP, NMe2
(32)
SiH2Ph
The chemoselectivity of the catalyst for alkynes over alkenes is of interest. Ex-
cellent discrimination is achieved in substrates containing an alkyne paired with
a hindered olefin (Eqs. 27, 33) [38]. When offered a monosubstituted olefin
(Eq. 34) the catalyst is less selective, producing mixtures of alkyl- and vinylsi-
lanes. As previously noted for the hydrogenation of dienes, the addition of a
group allylic to the alkene sterically shields the double bond and can electroni-
O S O
n-C8H10 n-C8H10 n-C8H10
O S O
cally deactivate it as well (Eq. 35). This allows virtually complete selectivity for
alkyne insertion.
5% Cp*2YMe·THF
(33)
PhSiH3, 50 ˚C
24 h, 90% H SiH2Ph
5% Cp*2YMe·THF
( ) 0.5 equiv. PhSiH3
6
50 ˚C, 12 h
(34)
SiH2Ph SiH2Ph
( ) ( )
6 + + 6
( )
H SiH2Ph 6 H SiH2Ph
OTBDMS
5% Cp*2YMe·THF
TBDMSO
PhSiH3, rt
(35)
21 h, 93%
H SiH2Ph
5
Catalytic Cyclization Reactions
The propensity for organolanthanides and group 3 organometallics to undergo
olefin insertion reactions leading to the polymerization of α-olefins provides the
possibility of cyclizing dienes and other polyunsaturated substrates. A reasona-
ble catalytic cycle for such a transformation is depicted in Fig. 7. There are sev-
eral requirements for successful cyclization. In an unsymmetrical diene, selec-
tive reaction at a single alkene is necessary to avoid a mixture of regioisomeric
products. As noted previously, this requirement has been met with organoyt-
trium catalysts (Eq. 8) [24]. If the reactions are carried out under a hydrogen at-
mosphere, intramolecular olefin insertion must occur more rapidly than σ-
bond metathesis of the newly formed organometallic with hydrogen. Diastereo-
selectivity is established in the cyclization, and should be predictable based
upon a simple chair transition structure for the cyclization. Finally, in the ab-
sence of hydrogen the catalytic cycle can be completed by β-hydride elimination
to afford the exomethylene-substituted cycloalkane.
134 Gary A. Molander, Eric D. Dowdy
Cp*2LnR'
H2
- R'H
H H
CH3
"Cp*2LnH"
R R
n( ) n( )
H2
H H
LnCp*2
R R LnCp*2
n( ) n( )
- Cp*2LnH
H
R
n( )
5.1
Termination by β-Hydride Elimination
9% [DpScH]2
( )5 rt, 3 d
85%
(36)
t-Bu
t-Bu
Lanthanide- and Group 3 Metallocene Catalysis in Small Molecule Synthesis 135
5% [DpScH]2
80 ˚C, 10 min
(37)
quant
4% Me2SiCp'2Sc(H)(PMe3)
N N
80 ˚C, 10 min
(38)
93%
6% [DpScH]2
80 ˚C, 10 min
+ (39)
99%, 2 : 1
5.2
Termination by Hydrogenation
Organoyttrium catalysts have been utilized to effect the cyclization of dienes un-
der reductive conditions [39]. Excellent selectivity is achieved in these reactions
between two monosubstituted alkenes leading to a single regioisomeric product
(Eq. 40), and the diastereoselectivity is consistent with the simple chair transi-
tion structure model (Fig. 7). Both acetals and thioacetals are tolerated (Eq. 41),
whereas nitriles, esters, and sulfones preclude product formation (Eq. 42).
5% Cp*2YMe•THF
OTr
OTr 1 atm H2 (40)
benzene, rt, <1 h
99%, 21 : 1 ds
X 5% Cp*2YMe•THF X
R 5% Cp*2YMe·THF
no reaction
R 1 atm H2, benzene (42)
R = CN, CO2Me, SO2Ph
state, making cyclization more difficult. Additionally, it seems likely that a sub-
stantial amount of inverse (“2,1”) addition occurs. The failure of the secondary
benzylic organometallic thus formed to cyclize would lead to formation of the
uncyclized hydrogenated product.
5% Cp*2YMe·THF
+
1 atm H2, benzene (43)
rt, 1 h
53% 26%
Me 5% Cp*22YMe·THF Me
Si Si (44)
Me 1 atm H22, benzene Me
rt, 1 h, 64%
0.5% cat.
1 atm H2
neat, rt, 2 h
(45)
TMS
R*
5.3
Termination by Silylation
Cp*2LnR'
R3SiH
- R3SiR'
H H
CH2SiR3
"Cp*2LnH" R
R
n( ) n( )
R3SiH
H H
LnCp*2
R R
LnCp*2
n( ) n( )
5.3.1
Cyclization/Silylation of Terminal Dienes and Trienes
Early studies centered on the utilization of organolutetiums (Eq. 46) [32a], orga-
noneodymiums [33c], and organosamariums [27] for the cyclization of 1,5-hex-
adiene and homologs. The early studies of the cyclization/silylation process in-
cluded only unsubstituted dienes, leaving questions of regioselection, diastere-
oselection, and functional group toleration unanswered. A more thorough study
of this chemistry that focused on the application to small molecule synthesis was
performed utilizing organoyttrium complexes [40]. The organoyttrium-cata-
lyzed process employed on monosubstituted dienes appears to be quite general
for the synthesis of both five- and six-membered rings. For the synthesis of five-
membered rings, phenylsilane is a convenient “chain terminator.” It provides
high yields of cyclized/silylated products with no stereocenters introduced as a
result of the incorporation of the silicon atom. High diastereoselectivities are
achieved in many instances (Eq. 47).
cat. Cp*2LuMe
SiH2Ph
PhSiH3, pentane
(46)
30 min, quant.
OCPh3
5% Cp*2YMe•THF
OCPh3 PhSiH3, cyclohexane
(47)
SiH2Ph
rt, 1 h, 71%, 24 : 1 ds
strates and reagents are incorporated into the desired product and there are no
byproducts produced. Many of these reactions are so clean, in fact, that pouring
the reaction mixture through a short bed of Florisil to remove the catalyst, fol-
lowed by evaporation of the solvent and bulb-to-bulb distillation, leads to essen-
tially quantitative yields of analytically pure product.
The synthetic equivalency of the silane and an alcohol can be easily demon-
strated by subjecting the crude silane product to any of a variety of available ox-
idizing conditions (Eq. 48) [30, 40].
5% Cp*2YMe•THF Ph
PhSiH3
Ph cyclohexane, rt, 1 h SiH2Ph
1. HBF4•OEt2 (48)
CHCl3, 0 ˚C, 1 h Ph
2. KF, KHCO3, H2O2
THF, MeOH OH
heat, 18 h
74% overall yield
9 : 1 ds
3% Cp*2YCH(TMS)2
PhH2Si SiH2Ph
PhSiH3, benzene, rt (49)
26 h, 96%
OTBDMS
OTBDMS 5% Cp*2YMe·THF
(50)
MePhSiH2, cyclohexane SiHMePh
rt, 2 h, 99%, 1.5 : 1 ds
Lanthanide- and Group 3 Metallocene Catalysis in Small Molecule Synthesis 139
t-Bu
5% Cp*2YMe·THF N
MePhSiH2, cyclohexane SiHMePh
t-Bu rt, 2 h, 97%
N (51)
t-Bu
5% Cp*2YMe·THF N
Ph2SiH2, cyclohexane SiHPh2
rt, 20 h, 65%
S S 5% Cp*2YMe•THF S S
PhMeSiH2
(52)
cyclohexane, rt, 1 h SiHMePh
99%
In the reaction of triallylamine (Eq. 53) [40], after the first ring forming event
the organometallic is trapped by σ−bond metathesis with the silane instead of
undergoing an entropically unfavorable intramolecular olefin insertion to yield
a bridged bicyclic structure. The remaining double bond then competes effec-
tively for the catalyst, making the isolation of monocyclic silylated product bear-
ing a free allyl group impossible.
SiHMePh
5% Cp*2YMe·THF
N (53)
N MePhSiH2, cyclohexane
rt, 1 h, 90% SiHMePh
Monosubstituted diene systems have been employed for the synthesis of bicy-
clic systems as well as monocyclics [17]. The simplest way to accomplish this is
to construct a second ring onto an existing structure (Eq. 54). This cyclization
process is initiated at the alkene lacking allylic substitution. The formation of
the six-membered ring necessitates the use of methylphenylsilane as the silylat-
ing reagent. Because the silane itself comprises a new stereocenter, it must be re-
moved by oxidation to assess accurately the diastereoselectivity of cyclization.
Fluxionality of the five-membered ring results in a mixture of diastereomers at
the silylmethyl-substituted stereocenter.
140 Gary A. Molander, Eric D. Dowdy
OTBDMS
OTBDMS
5% Cp*2YCH3·THF
cyclohexane H
PhMeSiH2, rt,1 h
92%, 1.7:1 ds SiHMePh
(54)
OTBDMS
t-BuOOH, KH
DMF, 50 °C, 12 h H
86%, 1:1 ds
OH
SiHMePh
5% Cp*2YMe•THF H
PhMeSiH2
(55)
OTBS cyclohexane, rt, 1 h
86% OTBS
‡
Y
Y H
Olefin
Insertion
OR OR
Favored
‡
Olefin
Y Insertion Y
Disfavored
OR OR
‡
Y
Olefin
Y
Insertion
OR H
Favored OR H
H
H ‡ Olefin
OR Y Insertion
OR
Disfavored Y
PhMeSiH2
(56)
OTMS cyclohexane, rt, 1 h
89% OTMS
cyclohexane (57)
PhMeSiH2, rt, 1 h SiHMePh
98%
142 Gary A. Molander, Eric D. Dowdy
Cp*2YMe·THF
PhMeSiH2
PhMeSiH2
YCp*2
TBDMSO TBDMSO
YCp*2
TBSO
TBSO
5% Cp*2YMe•THF
PhMeSiH2
(58)
cyclohexane, rt, 1 h H
86% SiHMePh
OTMS SiHMePh
5% Cp*2YMe•THF TMSO
PhMeSiH2 (59)
cyclohexane, rt, 6 h
80% H
OTMS OTMS
H ‡ Olefin H
(a)
Insertion
Y Y
Favored
‡
OTMS
Olefin OTMS
Y
Y (b)
Insertion
H
H
Disfavored
OH (60)
OH
H
73%
H
5% Cp*2YMe•THF SiHMePh
PhMeSiH2 (61)
cyclohexane, rt, 1 h H
68%
first OTMS
OTMS OTMS YCp*2
"Cp*2YH" cyclization
YCp*2
H
second
cyclization
σ-bond
PhH2Si H metathesis Cp*2Y H
Scheme 1
"Cp*2YH"
YCp*2
first
cyclization
SiHMePh YCp*2
YCp*2
PhMeSiH2 second
σ-bond cyclization
metathesis
Scheme 2
H
5% Cp*2YCH3·THF
H
5% Cp*2YCH3·THF
In spite of their recognized Lewis acidity and the propensity to complex with
Lewis bases (particularly in an intramolecular chelate) [8], the organoyttrium
complexes can be utilized for the synthesis of nitrogen heterocycles. The proto-
col has been employed in a concise synthesis of (±)-epilupinine (Scheme 3) [43].
Lanthanide- and Group 3 Metallocene Catalysis in Small Molecule Synthesis 145
SiHMePh OH
t-BuOOH, KH, CsF
DMF, 45 ˚C, 12 h
N N
(±) - Epilupinine
51-62%, two steps
Scheme 3
5.3.2
Cyclization/Silylation of Hindered Dienes and Trienes
SiH2Ph
cat. Me2Si(CpTMS)2YCH(TMS)2
(64)
PhSiH3, cyclohexane
90 ˚C, 16 h, 79%
5% Me2Si(CpTMS)2YCH(TMS)2 SiH2Ph
PhSiH3, cyclohexane
(65)
90 ˚C, 16 h, 57%
146 Gary A. Molander, Eric D. Dowdy
5% Me2Si(CpTMS)2YCH(TMS)2
PhSiH3, cyclohexane
90 ˚C, 24 h, 64% (66)
SiH2Ph
Recognizing the need to carry out these reactions under milder conditions, a
second generation catalyst was tested in similar systems [44]. This new system,
lacking the silicon hinge, proved to be extraordinarily reactive, in fact orders of
magnitude more reactive than any other neutral catalyst tested to date in similar
substrates (compare Eqs. 66, 67). As described previously, the precatalyst has
modest air stability, and yet the catalyst itself is remarkably reactive. Catalyst
loadings as low as 0.5 mol% have been employed in reasonably large scale reac-
tions (10–20 mmol), and the catalyst also supports the presence of the standard
functional groups for this general class of catalysts (Eq. 68). One problem with
the catalyst is that it apparently becomes deactivated over time because of hy-
dride dimer formation. Ideal reaction conditions for slow-reacting substrates
thus involve addition of smaller portions of the catalyst at fixed intervals to
maintain an active concentration of the catalyst.
5% [(CpTMS)2YMe]2 SiH2Ph
(67)
PhSiH3, cyclohexane
rt, 1 h, 88%
OTBDPS OTBDPS
5% [(CpTMS)2YMe]2 SiH2Ph
Me (68)
PhSiH3, cyclohexane
rt,16 h, 74%
5% [(CpTMS)2YMe]2 SiH2Ph
(69)
PhSiH3, cyclohexane
rt, 20 h, 50%
Lanthanide- and Group 3 Metallocene Catalysis in Small Molecule Synthesis 147
5% [(CpTMS)2YMe]2
+
PhSiH3, cyclohexane
SiH2Ph
rt, 1 h
75%
(70)
SiH2Ph
+
14%
5% [(CpTMS)2LnMe]2 SiH2Ph
( )n ( )n +
PhSiH3, cyclohexane
rt, 1 h
n = 0, Ln = Lu 0%
n = 2, Ln = Y 84%
(71)
PhH2Si
+ ( )n
84%
0%
5% [(CpTMS)2LuMe]2
SiH2Ph
PhSiH3, cyclohexane (72)
rt,1 h, 93%
5.3.3
Cyclization/Silylation of Enynes and Dienynes
O
O OCPh3 5% Cp*2YMe•THF O
(73)
O PhSiH3
cyclohexane, rt, 2 h
88%, 35 : 1 ds OCPh3
PhH2Si
R 5% Cp*2YMe·THF
+
PhSiH3 R
cyclohexane, rt, 2h
SiH2Ph
+
R
SiH2Ph (74)
R % Yield ds
OTBDMS 93 6.5 : 1
OTIPS 93 12 : 1
OTr 84 24 : 1
Et 88 >50 : 1
CH2OMe (100 ˚C) 80 20 : 1
CH2OTBDMS 76 >50 : 1
N 91 40 : 1
is undoubtedly because of the poor chemoselectivity between the alkyne and the
allylic olefin. Alkyl substitution on the allyl group gives complete selectivity for
initial alkyne insertion, but the hindered olefin prevents bicyclic product forma-
tion (Eq. 77). Geminal dimethyl substitution at the allylic position, however,
provides selectivity in the initial insertion without stopping the second intramo-
lecular insertion from taking place (Eq. 78). This dienyne cyclization allows an-
other entry into the strained trans-bicyclo[3.3.0]octane system, and contains an
additional handle for further functionalization.
5% Cp*2YMe•THF
SiH2Ph (76)
benzene-d6, PhSiH3
rt, 2.5 h, 33%
5% Cp*2YMe•THF
SiH2Ph (77)
benzene-d6, PhSiH3
rt, 24 h, 84%
H
5% Cp*2YMe•THF
O OTr
O 5% Cp*2YMe•THF
(79)
SiH2Ph
benzene-d6, PhSiH3 O H
OTr rt, 90%, >40 : 1 ds
O
150 Gary A. Molander, Eric D. Dowdy
M H
R' R'
H M
R R H
R' H
H R
H H
R
conformation leading to the formation
of the minor product
M H
H H
R' M
R R R'
R' H
R' R
H R'
R
conformation leading to the formation
of the major product
O Bn
O 5% Cp*2YMe•THF
SiH2Ph (80)
Bn benzene-d6, PhSiH3 O H
rt, 89%, 30 : 1 ds
O
If the alkyl chain is shortened by one carbon, a substrate is generated that has
the possibility for bicyclo[2.2.1]heptane generation (Eq. 81) [38]. After the ini-
tial alkyne insertion, the catalyst must choose between cyclobutane and cy-
clopentane formation. The five-membered ring is formed because of the lower
Lanthanide- and Group 3 Metallocene Catalysis in Small Molecule Synthesis 151
strain involved, but the remaining olefin is not inserted because the lowest ener-
gy conformation of the organometallic intermediate places the olefin out of
reach of the carbon-metal bond (Eq. 82).
SiH2Ph
OTBDMS
5% Cp*2YMe•THF
OTBDMS (81)
benzene-d6, PhSiH3
rt, 2h, 70%
SiH2Ph
H OTBDMS H
H
H (82)
R
R OTBDMS
YCp*2
YCp*2
6
Conclusions
The lanthanide- and group 3 metallocenes exhibit a rich chemistry that can be
exploited for the selective synthesis of small molecules. At least one class of cat-
alysts that are easily synthesized in a single pot exhibit reasonable stability in the
air. Consequently, they should be readily accessible to practicing synthetic or-
ganic chemists. Catalysts have been developed that permit reactions of mono-
substituted and 1,1-disubstituted alkenes, as well as internal alkynes. Practical
solutions for the inability of more highly substituted alkenes to insert have yet
to be reduced to practice. Terminal alkynes are rapidly metallated by these cat-
alysts, and are unlikely to be adaptable to many of the processes outlined in this
contribution. A variety of functional groups are tolerated by the catalysts (hali-
des, acetals, thioacetals, ethers, and amines), and thus highly functionalized
substrates of interest in complex molecule synthesis should be amenable to uti-
lization in selected processes. A high degree of chemoselectivity can be achieved
in polyunsaturated systems, and regiochemistry in the olefin insertion reactions
can also be controlled. Both of these conspire to provide an effective means to
control the direction of cyclization in unsymmetrical systems. Excellent diaster-
eoselectivity can often be achieved, taking advantage of both the inherent selec-
tivity of the catalysts themselves as well as the highly ordered transition struc-
tures involved in the intramolecular processes they promote. Finally, elegant
asymmetric catalysts have been synthesized and utilized in selective reactions.
Further developments in this arena are certain to produce a class of readily avail-
able catalysts that provide rapid and efficient entry to a wide range of complex
structures in enantiopure form.
152 Gary A. Molander, Eric D. Dowdy
Acknowledgments
Our work in this area was supported by a research grant from the National Insti-
tutes of Health (GM48580). The authors thank a very talented group of cowork-
ers whose names appear in the references below for their dedicated effort to this
chemistry.
7
References
1. (a) Gavens PD, Bottrill M, Kelland JW, McMeeking J (1982) Ziegler-Natta catalysis. In:
Wilkinson G, Stone, FGA, Abel EW (eds) Comprehensive organometallic chemistry,
vol 3. Pergamon, Oxford, chap 22.5; (b) Ballard DGH (1973) Adv Catal 23:263; (c) Bal-
lard DGH, Burnham DR, Twose DL (1976) J Catal 44:116; (d) Sinn H, Kaminsky W
(1980) Adv Organomet Chem 18:99; (e) Chien JCW (ed) (1979) Coordination polym-
erization. Academic Press, New York; (f) Boor J (1979) Ziegler-Natta catalysts and po-
lymerizations. Academic Press, New York; (g) Keim W, Behr A, Röper M (1982) Alkene
and alkyne oligomerization, cooligomerization and telomerization reactions. In:
Wilkinson G, Stone FGA, Abel EW (eds) Comprehensive organometallic chemistry, vol
8. Pergamon, Oxford, chap 52; (h) Gates BC, Katzer JR, Schuit GCA (1979) Chemistry
of catalytic processes. McGraw-Hill, New York; (i) Parshall G (1980) Homogeneous ca-
talysis. Wiley-Interscience, New York; (j)Pino P, Mülhaupt R (1980) Angew Chem, Int
Ed Engl 19:857; (k) Coates GW, Waymouth RM (1995) Transition metals in polymer
synthesis: Ziegler-Natta reaction. In: Abel EW, Stone FGA, Wilkinson G (eds) Compre-
hensive organometallic chemistry II, vol 12. Pergamon, Oxford, chap 12.1; (l) Zucchini
U, Dall’Occo T, Resconi L (1993) Indian J Technol 31:247; (m) Simonazzi T, Gianni U
(1994) Gazz Chim Ital 124:533; (n) Brintzinger HH, Fischer D, Waymouth RM (1995)
Angew Chem, Int Ed Eng 34:1143; (o) Keii T, Soga K (eds) (1986) Catalytic polymeri-
zation of olefins. Kodenska, Tokyo; (p) Reichert KH (1983) In: Quirk RP (ed) Transi-
tion metal catalyzed polymerizations: alkenes and dienes. Harwood Academic, New
York
2. (a) Marks TJ, Ernst RD (1982) Scandium, yttrium and the lanthanides and actinides.
In: Wilkinson G, Stone FGA, Abel EW (eds) Comprehensive organometallic chemistry,
vol 3. Pergamon, Oxford, chap 21; (b) Watson PL, Parshall GW (1985) Acc Chem Res
18:51; (c) Edelmann FT (1995) Scandium, yttrium, and the lanthanide and actinide el-
ements, excluding their zero oxidation state complexes. In: Abel EW, Stone FGA,
Wilkinson G (eds) Comprehensive organometallic chemistry II, vol 4. Pergamon
Press, Oxford, chap 2; (d) Edelmann FT (1996) Top Curr Chem 179:247
3. (a) Johnson LK, Killian CM, Brookhart M (1995) J Am Chem Soc 117: 6414; (b)
Brookhart M, Wagner MI (1996) J Am Chem Soc 118:7219; (c) Feldman J, McLain SJ,
Parthasarathy A, Marshall WJ, Calabrese JC, Arthur SD (1997) Organometallics
16:1514; (d) Killian CM, Johnson LK, Brookhart M (1997) Organometallics 16: 2005
4. (a) Emrich R, Heinemann O, Jolly PW, Krüger C, Verhovnik GPJ (1997) Organometal-
lics 16:1511; (b) Scollard JD, McConville DH, Rettig SJ (1997) Organometallics 16:1810
5. Jeske G, Lauke H, Mauermann H, Swepston PN, Schumann H, Marks TJ (1985) J Am
Chem Soc 107:8091
6. (a) Borricello A, Busico V, Sudemijer O (1996) Macromolecular Rapid Commun
17:589; (b) Sinclair KB, Wilson RB (1994) Chem Ind 857
7. (a) Watson PL (1983) J Chem Soc, Chem Commun 276; (b) Evans WJ, Chamberlain LR,
Ulibarri TA, Ziller JW (1988) J Am Chem Soc 110:6423
8. Piers WE, Shapiro PJ, Bunel EE, Bercaw JE (1990) Synlett 74
9. Brown HC (1975) Organic syntheses via boranes. Wiley, New York
10. Haar CM, Stern CL, Marks TJ (1996) Organometallics 15:1765
Lanthanide- and Group 3 Metallocene Catalysis in Small Molecule Synthesis 153
11. (a) Schumann H, Genthe W, Bruncks N (1981) Angew Chem, Int Ed Engl 20:119; (b)
Evans WJ, Wayda AL, Hunter WE, Atwood JL (1981) J Chem Soc, Chem Commun 292
12. Evans WJ (1987) Polyhedron 6:803
13. Deelman BJ, Booij M, Meetsma A, Teuben JH, Kooijman H, Spek AL (1995) Organome-
tallics 14:2306
14. Stern D, Sabat M, Marks TJ (1990) J Am Chem Soc 112:9558
15. den Haan KH, Wielstra Y, Teuben JH (1987) Organometallics 6:2053
16. Wang KK, Brown HC (1982) J Am Chem Soc 104:7148
17. (a) Nichols PJ (1997) PhD thesis. University of Colorado at Boulder; (b) Molander GA,
Nichols PJ, Noll BC (1998) J Org Chem 63:2292
18. Markus Keitsch (unpublished results)
19. Shriver DF, Drezdzon MA (1986) The manipulation of air-sensitive compounds, 2nd
edn. Wiley-Interscience, New York
20. Jeske G, Schock LE, Swepston PN, Schumann H, Marks TJ (1985) J Am Chem Soc
107:8103
21. (a) Evans WJ (1982) In: Hartley FR, Patai S (eds) The chemistry of the metal-carbon
bond. Wiley, Chichester; (b) Schumann H, Genthe W (1984) In: Gschneidner KA Jr, Ey-
ring L (eds) Handbook on the chemistry and physics of rare earths. Elsevier, Amster-
dam; (c) Schumann H (1984) Angew Chem, Int Ed Engl 23:474; (d) Evans WJ (1985)
Adv Organomet Chem 24:131; (e) Schumann H, Meese-Marktscheffel JA, Esser L
(1995) Chem Rev 95:865; (f) Schaverien CJ (1994) Adv Organomet Chem 36:283
22. Jeske G, Lauke H, Mauermann H, Schumann H, Marks TJ (1985) J Am Chem Soc
107:8111
23. (a) Zinnen HA, Pluth JJ, Evans WJ (1980) J Chem Soc, Chem Commun 810; (b) Broth-
ers PJ (1981) Prog Inorg Chem 28:1; (c) Evans WJ, Meadows JH, Wayda AL, Hunter WE,
Atwood JL (1982) J Am Chem Soc 104:2008; (d) Schumann H, Genthe WJ (1981) J Or-
ganomet Chem 213:C7; (e) Marks TJ (1987) Inorg Chim Acta 140:1; (f) Thompson ME,
Baxter SM, Bulls AR, Burger BJ, Nolan MC, Santarsiero BD, Schaefer WP, Bercaw JE
(1987) J Am Chem Soc 109:203
24. Molander GA, Hoberg JO (1992) J Org Chem 57:3266
25. Molander GA, Winterfeld J (1996) J Organomet Chem 524:275
26. Giardello MA, Conticello VP, Brard L, Gagné MR, Marks TJ (1994) J Am Chem Soc
116:10,241
27. Fu PF, Brard L, Li Y, Marks TJ (1995) J Am Chem Soc 117:7157
28. Evans WJ, Bloom I, Hunter WE, Atwood JL (1981) J Am Chem Soc 103:6507
29. (a) Speier JL (1979) Adv Organomet Chem 17:407; (b) Ojima I (1989) The hydrosilyla-
tion reaction. In: Patai S, Rappoport Z (eds) The chemistry of organic silicon com-
pounds. Wiley, London; (c) Yamamoto K, Takemae M (1990) Synlett 259; (d) Takahashi
T, Hasegawa M, Suzuki N, Saburi M, Rousset CJ, Fanwick PE, Negishi E (1991) J Am
Chem Soc 113:8564; (e) Boudjouk P, Han BH, Jacobsen JR, Hauck BJ (1991) J Chem
Soc, Chem Commun 1424; (f) Kesti MR, Abdulrahman M, Waymouth RM (1991) J Or-
ganomet Chem 417:C12; (g) Kesti MR, Waymouth RM (1992) Organometallics 11:1095
30. (a) Tamao K, Ishida N, Tanaka T, Kumada M (1983) Organometallics 2:1649; (b) Ber-
gens SH, Nogeda P, Whelan J, Bosnich B (1992) J Am Chem Soc 114:2121; (c) Taber DF,
Yet L, Bhamidipati RS (1995) Tetrahedron Lett 36:351; (d) Fleming I, Hennings R, Plaut
H (1984) J Chem Soc, Chem Commun 29; (e) Fleming I, Sanderson PE (1987) Tetrahe-
dron Lett 28:4229; (f) Fleming I, Winter SBD (1993) Tetrahedron Lett 34:7287; (g) Sm-
itrovich JH, Woerpel KA (1996) J Org Chem 61:6044
31. Burgess K, Ohlmeyer MJ (1991) Chem Rev 91:1179
32. (a) Watson PL, Tebbe FN (1991) Chem Abstr 114:123,331p; (b) Forsyth CM, Nolan SP,
Marks TJ (1991) Organometallics 10:1450; (c) Radu NS, Tilley TD, Rheingold AL
(1992) J Am Chem Soc 114:8293; (d) Radu NS, Tilley TD (1995) J Am Chem Soc
117:5863
154 Gary A. Molander, Eric D. Dowdy Lanthanide- and Group 3 Metallocene Catalysis
33. (a) Beletskaya IP, Voskoboinikov AZ, Parshina IN, Magomedov GKI (1990) Izv Akad
Nauk SSR Ser Khim 693; (b) Sakakura T, Lautenschlager HJ, Tanaka M (1991) J Chem
Soc, Chem Commun 40; (c) Onozawa S, Sakakura T, Tanaka M (1994) Tetrahedron Lett
35:8177
34. Molander GA, Dowdy ED, Noll BC (1998) Organometallics 17:3754
35. Molander GA, Julius M (1992) J Org Chem 57:6347
36. Schumann H, Keitsch MR, Winterfeld J, Mühle S, Molander GA (1998) J Organomet
Chem 559:181
37. (a) Heeres HJ, Heeres A, Teuben JH (1990) Organometallics 9:1508; (b) Heeres HJ,
Teuben JH (1991) Organometallics 10:1980
38. (a) Molander GA, Retsch WH (1995) Organometallics 14:4570; (b) Retsch WH (1997)
PhD thesis, University of Colorado at Boulder
39. Molander GA, Hoberg JO (1992) J Am Chem Soc 114:3123
40. Molander GA, Nichols PJ (1995) J Am Chem Soc 117:4415
41. (a) Trost BM (1995) Angew Chem, Int Ed Engl 34:259; (b) Trost BM (1991) Science
254:1471
42. (a) Bailey WF, Khanolkar AD, Gavaskar KV (1992) J Am Chem Soc 114:8053; (b) Cooke
MP, Gopal D (1994) Tetrahedron Lett 35:2837; (c) Davis JM, Whitby RJ, Jaxa-Chamiec
A (1992) Tetrahedron Lett 33:5655; (d) Rousset CJ, Swanson DR, Lamaty F, Negishi E
(1989) Tetrahedron Lett 30:5105; (e) Doi T, Yanagisawa A, Takahashi T, Yamamoto K
(1996) Synlett 145; (f) Piers E, Kaller AM (1996) Tetrahedron Lett 37:5857
43. Molander GA, Nichols PJ (1996) J Org Chem 61:6040
44. Molander GA, Dowdy ED, Schumann H (1998) J Org Chem 63:3386
45. Molander GA, Retsch WH (1997) J Am Chem Soc 119:8817
Influence of Solvents or Additives on the Organic Chemistry
Mediated by Diiodosamarium
Henri B. Kagan*, Jean-Louis Namy
The reactivity of diiodosamarium in solvents other than THF or in mixtures of solvents are
discussed. The influence of additives able to coordinate to samarium are then considered
(amides, amines or ethers). Proton donors sometimes drastically modify the selectivity of
reactions induced by diiodosamarium; some metal salts [such as Fe(III) or Ni(II)] in cata-
lytic amounts may also strongly accelerate or modify reactions induced by diiodosamar-
ium. Thanks to the above tuning of the diiodosamarium reactivity, rich and diversified or-
ganic transformations have been performed, some examples of which are presented.
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
1
Introduction
Diiodosamarium was prepared for the first time by Matignon and Cazes in 1906
by heating triiodosamarium at high temperature under an atmosphere of hydro-
gen [1]. For a long time, subsequent preparations used high temperature tech-
nology and solid state chemistry (for example, see [2]).
In 1977 we described the mild preparation of diiodosamarium at room tem-
perature by reaction between samarium powder and 1,2-diiodoethane in THF
[3]. The dark-green solution of SmI2 in THF (0.1 M) was used as a convenient
reducing agent. We reported in 1980 the basic organic transformations induced
by this new reagent [4]. Since that time we, as well as many other groups, have
discovered plenty of new reactions, often performed under smooth conditions
and with a high selectivity, through a combination of radical and organometallic
chemistry. Diiodosamarium is becoming a major reagent in organic chemistry,
many reviews on its reactivity are available (for example, see [5–12]). In this ar-
ticle we shall discuss the various ways to tune the reducing properties of diio-
dosamarium by changing THF to another solvent or by introduction of various
additives which can coordinate to samarium or act as catalysts (for a short re-
view, see [13]).
2
Influence of Solvents
2.1
Tetrahydrofuran (THF)
is performed in THF to give stable solutions (under inert atmosphere) of the re-
agent [3,4,14]. During the period 1977–1986, the basic organic transformations
induced by SmI2 involved reactions in THF. In 1987 Inanaga et al. described the
acceleration of SmI2-mediated reactions by addition of some HMPA (usually
5%) in THF [15,16]. This effect is discussed in Sect. 3. It involves the coordinat-
ing properties of HMPA towards Sm(II) and Sm(III) ions. Another way to alter
the redox properties of the Sm(II)/Sm(III) couple is to replace THF by another
solvent. One method is to evaporate THF and to add the desired solvent. This ap-
proach has been used to prepare SmI2 complexed by various nitriles (vide infra).
The difficulty comes from the necessity to fully remove the last THF molecules
which remain bound to SmI2, and which may influence its reactivity [17]. The
best strategy, therefore, is to try to generate SmI2 directly in the desired solvent.
This has been achieved successfully in nitriles, in tetrahydropyran (THP), in tet-
raglyme and even in benzene. There are several incentives for performing SmI2-
induced chemistry in various solvents. One can expect some of the following
consequences:
1. Acceleration of some reactions,
2. An improved selectivity,
3. New reactions, and
4. Elimination of by-products generated from competitive reactions with THF
(H-abstraction, ring opening, etc.).
2.2
Tetrahydropyran (THP)
Scheme 1
158 Henri B. Kagan, Jean-Louis Namy
yields were slightly better than in THF which acts as an intermolecular hydrogen
donor to the aryl radical. The appended orthoiodobenzyl in 1 thus generated or-
ganosamarium 2 which then produced N-benzyl amines 3 by reaction with var-
ious electrophiles. SmI2 was directly prepared in THP. Undheim et al. applied
this reaction to the alkylation of saturated heterocycles α to nitrogen [21].
In 1994 we set up independently a preparation of SmI2 in THP from samar-
ium metal and 1,2-diodoethane [22]. We wanted to explore the possibility of im-
proving the yields of the acid chloride chemistry, where by-products arise by
ring opening of THF. Indeed there was a complete absence of the above by-prod-
ucts. Interestingly, acid chlorides with a branched α-carbon react with SmI2 to
give stable THP solutions of acylsamariums 4 which can then react with an alde-
hyde or ketone to give the mixed α-ketols 5 (Scheme 2). This two-step process
avoids the competitive pinacol formation, especially fast when an aldehyde is in
the presence of SmI2. Some examples are indicated in Scheme 2. The acylsamar-
ium structure 4 is well supported by the formation of deuterated aldehyde 8, by
reduction of acid chloride 7 and by reaction with D2O.
Scheme 2
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 159
Acid chlorides with an α-carbon with at least one hydrogen are prone to give
self-coupling to symmetrical α-ketols 6. However, the Barbier procedure in THP
allows the preparation of mixed α-ketols 5 in good yields even when using alde-
hydes.
Recently we discovered that allylic and benzylic samarium compounds could
be generated in THP [23]. For example, allylic iodide 9 treated at –15 °C with a
THP solution of SmI2 (2 equiv) followed by addition of butanone at the same
temperature gave in 90% isolated yield homoallylic alcohol 10 devoid of the usu-
al branched isomer (Scheme 3). Treatment of the allylic organosamarium with
D2O provided a fair yield of deuterated alkene. Allylsamarium itself is smoothly
prepared at 0 °C and reacts with various types of substrates. Especially interest-
ing is the high endo stereoselectivity of addition to camphor, the allylation of im-
ine 11 and the addition on β-keto ester 12 which is easily enolizable. Diiodoben-
zylsamarium may be prepared at –15 °C from benzyl bromide and reacts with 2-
octanone to give the tertiary alcohol in good yield.
In conclusion, THP may in many cases improve the efficiency of reactions in-
itially performed in THF and stabilize organosamarium species which are gen-
erated and which may subsequently react with various electrophiles.
Scheme 3
160 Henri B. Kagan, Jean-Louis Namy
2.3
Ethers Other Than THF or THP
There are almost no reports of preparation of SmI2 in ethers other than THF or
THP. Inanaga et al. prepared SmI2 in 1,3-dioxolane and then added 10% ace-
tonitrile to produce a clear solution. This solution was used in the masked
formylation of aldehydes or ketones in the presence of iodobenzene [24]. Iodo-
benzene is transformed into the benzene radical which generates the 1,3-diox-
olanyl radical. The latter is presumably reduced to the corresponding organosa-
marium which then adds on the carbonyl. This process is illustrated in Scheme 4
for cyclododecanone. It is essential to avoid the presence of THF which will com-
pete with 1,3-dioxolane as a hydrogen donor to the benzene radical.
Tetraglyme (2,5,8,11,14-pentaoxapentadecane) has been used in the hy-
droxymethylation of aldehydes via addition of benzyl chloromethyl ether in the
presence of SmI2. Addition of tetraglyme suppresses the competitive pinacol for-
mation, presumably by a complex formation with SmI2 as evidenced by a purple
color [25]. Tetraglyme has also been used as a cosolvent in an intramolecular
Barbier reaction involving an iodoaldehyde and a THF solution of diiodosamar-
ium [26].
2.4
Nitriles
Scheme 4
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 161
Scheme 5
could be readily prepared in acetonitrile [28]. This author claimed that mixed α-
ketols 13 were produced by a two-step procedure and in good yields in ace-
tonitrile. We reinvestigated this work and found that it gave a mixture of prod-
ucts where the pinacol derived from the ketone was the major product [29]. Re-
activity of SmI2 in acetonitrile towards acid chlorides is roughly the same as in
THF. Many by-products are formed involving some reactions with acetonitrile.
In order to avoid the C-H acidity in the α of the nitrile group we investigated the
use of pivalonitrile. It is possible to prepare SmI2 (as a slurry) in pivalonitrile
from samarium powder and 1,2-diiodoethane [29]. An X-ray crystal structure of
SmI2-(NCCMe3)2 was performed by Sen et al. and shows a distorted octahedron
around samarium with a bent Sm-N=C structure [30].The cross-coupling be-
tween acid chlorides and ketones gave α-ketols 2 in moderate yields. Barbier re-
actions are slower in acetonitrile than in THF; however, the regioselectivity of
the reactions between allylic halides and ketones is significantly improved (in fa-
vor of the unbranched isomer). Surprisingly, HMPA does not enhance the reac-
tivity of SmI2 in nitriles (Barbier reactions or pinacolization of ketones). Reac-
tions are accelerated by catalytic amounts of some transition-metal salts (see
Sect. 3.3.6). Diiodosamarium may be prepared in propanenitrile or octaneni-
trile, but these solvents do not offer special advantages over pivalonitrile.
It is worth pointing out that acetonitrile was used as a cosolvent by Inanaga et
al. in 1987 [4]. These authors later studied SmI2-promoted aryl radical cycliza-
tion with olefins and obtained quite good yields in acetonitrile [31].
SmI2 in acetonitrile has been used for the preparation in good yields of 1,3-
diketones 15 [32]. The authors first added the α-halo ketone to the SmI2 solution
and subsequently added the desired acid chloride or anhydride. No comparisons
were given between reactions run in THF or acetonitrile.
162 Henri B. Kagan, Jean-Louis Namy
Scheme 6
Ishii et al. prepared an Sm(II) reagent in acetonitrile from NaI, ClSiMe3 and
samarium grain. A soluble species was produced with the deep-green color of
SmI2 but its structure was not established [33]. At –40 °C this Sm(II) reagent is
able to dehalogenate α-chloro or α-bromo ketones or esters, if methanol is
present as the proton source. Acetonitrile was superior to THF or DME for this
reaction. The reagent gave a faster reaction compared to SmI2/THF for the for-
mation of pinacols of acetophenone [34]. Curiously, benzaldehyde did not pro-
duce hydrobenzoin unless HMPA was added. A lactone formation was the result
of the coupling of methyl acrylate and a ketone or imine (Scheme 6). Reformat-
sky-type reactions between ethyl α-bromoacetate and ketones gave satisfactory
yields of β-hydroxy esters, without showing any special improvement with re-
spect to the SmI2/THF system. However, it is interesting to note that some Refor-
matsky or Barbier reactions have been achieved in good yields with octanal,
without perturbation by pinacol formation.
2.5
Benzene
Scheme 7
3
Influence of Additives
3.1
Electron-Rich Additives
3.1.1
Hexamethylphosphoramide (HMPA)
The highly promoted effect of HMPA on reactions of SmI2 was reported for the
first time in 1986 by Inanaga et al. [38] for the reductive cross-coupling of carb-
onyl compounds with α,β-unsaturated esters (Scheme 8). HMPA was used as a
cosolvent with THF (ca. 5%). Enhancement of the coupling rates and yields were
quite remarkable; however, the diastereoselectivity was sometimes diminished
in the presence of HMPA.
Independently, Fukuzawa and co-workers [39] reported intramolecular reac-
tions leading to bicyclic γ-lactones from keto or aldo α,β-unsaturated esters
(Scheme 9). The reaction was mediated by SmI2 in the presence of HMPA
(THF/HMPA=10/1) but the beneficial effect of HMPA was less obvious in that
case.
164 Henri B. Kagan, Jean-Louis Namy
Scheme 8
Scheme 9
Scheme 10
Inanaga et al. have developed the use of HMPA for the highly regioselective
reduction of α,β-epoxy esters and δ,γ-epoxy α,β-unsaturated esters [40]
(Scheme 10).
The best results were obtained when HMPA (5 equiv) and dimethylamino
ethanol (DMAE, 2 equiv) were used as additives. The authors also reported the
reductive cross-coupling of 1,3-dioxolane with carbonyl compounds to yield α-
hydroxy derivatives [24]; the reaction was performed in the presence of HMPA
(ca. 5% of the solvent) and was complete within 5 min at room temperature
(Scheme 11).
A remarkable effect by HMPA was observed in the reduction of organic hali-
des with SmI2 in THF [41] (Scheme 12).
As mentioned in Sect. 2.3, Inanaga and co-workers have demonstrated that
Barbier-type reactions of organic halides with carbonyl compounds are promot-
ed by addition of HMPA [16]. They reported a mild convenient method for the
direct synthesis of lactones from bromo esters and ketones or aldehydes by using
a HMPA-promoted Barbier-type reaction with SmI2 (Scheme 13). They also
found that the SmI2/THF-HMPA system was highly useful for the generation of
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 165
Scheme 11
Scheme 12
Scheme 13
Scheme 14
Scheme 15
purple solution was concentrated and toluene was added. After a few days at
room temperature, black-purple blocks of SmI2(HMPA)4 were deposited in 90%
yield, and they were structurally characterized. It appears that the Sm(II) ion sits
on an inversion center and is bonded by two I– anions and four HMPA ligands
in a distorted octahedron. The central Sm(II) ion and the four HMPA ligands are
exactly coplanar and the two iodide anions are mutually trans (Scheme 15).
In a recent work, electrochemical studies of the reducing power of SmI2 in
THF and the effect of HMPA cosolvent as a function of concentration have been
reported [45]. This effect was studied by recording a linear sweep voltamogram
for each cosolvent addition. The oxidation potential vs. the Ag/AgNO3 reference
electrode for 0 to 6 equiv of HMPA cosolvent was measured. The addition of
3 equiv of HMPA to SmI2 had a drastic effect on the redox potential, increasing
the oxidation potential from –1.33 to –1.95 V. The addition of 4 equiv of HMPA
to SmI2 increased the potential even further to –2.05. Further addition of HMPA
showed no effect on the redox potential.
These observations are consistent with the structure determined by Hou et
al.; the complex SmI2(HMPA)4 should be the reactive intermediate responsible
for the unique reactivity of the SmI2/THF–HMPA system.
Preparation, structural characterization and reactivity studies of SmI3–
HMPA complexes have also been performed [46]. The composition of the dried
complex was found to be SmI3(HMPA)4. The reactivity of this complex was ex-
amined in comparison with that of SmI3 in THF only. In contrast to the latter
complex, the HMPA complex was not reduced to low-valent samarium species
by n-butyl lithium or sec-butyl lithium. These results might be interpreted by
considering that this Sm(III) complex is well stabilized by the coordination of
HMPA which could be one of the driving forces for the facile electron transfer
from SmI2–HMPA complexes.
Molander et al. [47] suggested that other factors could be involved in the ef-
fective role of this additive: disaggregation of SmI2, f-orbital perturbation due to
the ligand field effect in the presence of the strong donor ligand HMPA raising
the energy of the HOMO (electron-donating orbital). A combination of these ef-
fects might also be considered.
The effect of HMPA on the spectroscopic properties of SmI2 has also been ex-
amined [48].
We wish to review here some studies that stress the specific role played by
HMPA in samarium diiodide reactions.
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 167
Scheme 16
Scheme 17
168 Henri B. Kagan, Jean-Louis Namy
Scheme 18
Scheme 19
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 169
Scheme 20
Scheme 21
170 Henri B. Kagan, Jean-Louis Namy
Scheme 22
Scheme 23
Scheme 24
Scheme 25
Scheme 26
Scheme 27
Scheme 28
Scheme 29
Scheme 30
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 173
run in the presence of HMPA resulted in all cases in some depression in the di-
astereoselection [71–73]. Various other reactions have been reported which ex-
emplify the crucial role of the solvent HMPA.
Inanaga et al. achieved the hydrodimerization of conjugated esters and
amides with the SmI2/THF–HMPA system in the presence of a proton source
[74].(Scheme 31). However, inexplicably, Alper et al. achieved reductions (not
reductive dimerization) of the double bond of α,β-unsaturated esters and
amides with the same system [75]. In the latter paper, however, the presence of
a proton donor is not mentioned.
The reductive coupling of α-bromoacetate to succinic acid diesters mediated
by SmI2 in the presence of HMPA has been reported [76] (Scheme 32).
In contrast, reactions performed without HMPA as additive do not lead to
dimerization but to β-keto esters [77] (Scheme 33). Obviously, the intermediate
samarium enolate which is formed behaves in a different way according to the
presence or the absence of HMPA.
The reductive dimerization of cyclopropane-1,1-dicarboxylic esters using
SmI2/THF has also been reported (Scheme 34). In this case, the addition of
Scheme 31
Scheme 32
Scheme 33
Scheme 34
174 Henri B. Kagan, Jean-Louis Namy
HMPA did not improve yields of the dimer. On the contrary, yields were higher
in the absence of HMPA [78].
A new intramolecular reductive carbon–carbon bond formation has been
noted in reactions of 1-(2-formyloxyethyl)-3-formyl-oxycycloalkene with SmI2
giving a spiro hemiacetal (Scheme 35). The absence of HMPA caused a consid-
erable decrease in the yield of the hemiacetal [79].
Coupling reactions of indoles and carbonyl compounds (either intra- or
intermolecularly) have been described. This is a new method for hydroxylation
at the C2 position of indole (Scheme 36). The role of HMPA is crucial to prevent
reduction or pinacolic coupling of aromatic carbonyl compounds [80].
Stereocontrolled decalin ring annulation reactions through the hydroxyl
group directed pinacol coupling using SmI2 have been reported by Matsuda and
et al. Stereocomplementarity was observed depending on the presence or ab-
sence of HMPA [81] (Scheme 37). The authors propose that without HMPA only
the aldehyde of the substrate is reduced to the ketyl radical during the initial re-
Scheme 35
Scheme 36
Scheme 37
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 175
duction by SmI2, whereas in the presence of HMPA a ketyl radical pair is gener-
ated through a single electron transfer from SmI2.
Katritzky et al. have prepared α-amino carbanions from tosyl methyl amines.
The transformation was achieved by treatment with 2 equiv of SmI2 [82] in the
presence of an electrophile (Scheme 38). In the absence of HMPA the use of iso-
butyraldehyde as the electrophile gave only the amine dimer (98%). However,
the addition of 5% of HMPA resulted in a 67% yield of the desired product (62%)
with a minor amount of diamine (17%). The formation of the hydroxy amine
probably involves a carbanion intermediate whereas coupling gives the diamine.
Samarium diiodide has also been used for the intramolecular coupling of al-
dehydes and ketones with O-benzyl formaldoxime [83], for the corresponding
intermolecular coupling with diphenylhydrazone [84,85] and for the intra-
molecular coupling of an α,β-unsaturated ester with an oxime ether [86]
(Scheme 39). In all these cases the addition of HMPA was found to be essential
for a successful reaction.
3.1.2
N,N'-Dimethylpropyleneurea (DMPU)
Scheme 38
Scheme 39
176 Henri B. Kagan, Jean-Louis Namy
Curran et al. were the first to use SmI2/THF–DMPU. This system promoted a
tandem radical cyclization [87] (Scheme 40). Reduction of 23 with SmI2/THF
produced 24, 25 and 26 in low yield; addition of HMPA accelerated the reaction
and gave 24 in 91% yield. DMPU as additive gave similar results to HMPA with
respect to reaction rate and yield but 25 and 26 were also formed albeit in low
yield (9 and 4%). In addition, the cyclization in THF–HMPA required less than
2 equiv of SmI2 (1.3 equiv) but when DMPU was used, a larger excess of SmI2
was needed, perhaps because the reaction mixture was heterogeneous.
Ring scission of cyclic β-halogeno ethers to olefinic alcohols mediated by
SmI2 has been studied. Scission in the tetrahydrofuran series could be accelerat-
ed by addition of HMPA or DMPU with only a small deterioration in diastereo-
selectivity, but in the tetrahydropyran series there was a drastic change in the
stereochemistry of the product when DMPU was used. Obviously, DMPU has a
substantial effect on the formation and conformation of complexes in solution
[88,89].
A mixture of THF and DMPU has been used as a solvent for SmI2 in various
reactions such as cyclization of alkynyl halides [90,91], tandem iodo-enone cy-
clization/samarium enolate aldol reaction [92], coupling of β-silylacrylic esters
[93], deprotection of arenesulfonamides [94] and pyridine-2-sulfonamides [95],
radical ring-opening reactions of cyclopropyl ketones and the trapping of the re-
sulting samarium(III) enolates by a variety of electrophiles [96] (Scheme 41).
From the above studies, the SmI2/THF–DMPU combination emerges as the
most useful, in particular because in this case the problem of second electron de-
livery is avoided.
Scheme 40
Scheme 41
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 177
Scheme 42
3.1.3
Other Nitrogen Ligands
Various other commercially available nitrogen ligands have been screened to in-
crease the reducing power of SmI2 in THF.
Inanaga et al. have studied the selective conjugate reduction of α,β-unsaturat-
ed esters and amides with SmI2. While the reductive dimerization was promoted
by HMPA [74], the use of dimethylformamide (DMF), dimethylacetamide
(DMA), tetramethylurea (TMU) and bidentate ligands such as N,N,N',N'-tetram-
ethylethylenediamine (TMEDA)and N,N,N'N'-tetramethylpropylenediamine
(TMPDA) were effective for the conjugate reduction of the substrate. DMA was
found to be the most promising. In contrast, in the presence of the tridentate
chelating agent N,N,N'N''N''-pentamethyldiethylenetriamine, no reaction took
place. This result might be due to the complete occupancy of the coordination
sphere of samarium [99].
The formation of an aryl radical from the corresponding aryl iodide or aryl
bromide using SmI2 has been carried out in the presence of various other nitro-
gen ligands (Scheme 43). The results obtained with Et3N, DBU or TMG were sat-
isfactory or even superior to the ones obtained with HMPA (27+28: 71%; 4 h;
27/28=81:29). It is worth noting that the ligand/SmI2 molar ratio has practically
no effect on the reaction course if more than 2 equiv of ligand are present [100].
TMEDA has sometimes been used as an additive in the reduction of α,β-
epoxy esters. In this case it was associated with HMPA and isopropanol [40], as,
178 Henri B. Kagan, Jean-Louis Namy
Scheme 43
Scheme 44
Scheme 45
3.1.4
Miscellaneous
Kamochi and Kudo found that the reducing ability of SmI2 is enhanced by addi-
tion of a base (KOH, LiNH2 or LiOMe), allowing the fast reduction of carboxylic
acids to alcohols [104]. Reduction of nitriles to primary amines has also been re-
alized by the system SmI2/50% KOH (1:2) in a few minutes at room temperature
[105].
BINAPO (the bis-oxide of BINAP) can complex samarium and gives some
stereocontrol in the formation of γ-butyrolactones by reaction of ketones with
α,β-unsaturated esters and SmI2 in THF. Enantiomeric excesses (ee's) of up to
90% were obtained [106].
3.2
Proton Sources
All the reactions induced by SmI2 start with a one-electron transfer giving rise
to an anion radical which evolves in various ways: radical chemistry, cleavage
into a carbanion and a radical, further reduction to a dianion, etc. Very often
marked differences in the product distribution are observed when SmI2-in-
duced transformations are performed in aprotic or protic conditions, because in
the latter case in situ protonation of key intermediates may occur. Kinetic stabil-
ity of SmI2/THF solutions in the presence of some water or alcohols have been
reported [15,107]. It is then possible to use these solutions for in situ protonation
of end products or intermediates. For example, it has been firmly established
that the samarium Barbier reaction between an organic halide RI and a ketone
involves a reactive organosamarium intermediate RSmI2 since the reaction per-
formed in the presence of SmI2/THF–t-BuOD or EtOD leads to the formation of
some RD. Under the same conditions, and in the absence of ketones, the deuter-
ation of RI into RD was also achieved in good yield [108,109].
Apart from its usefulness for some mechanistic investigations the presence of
protic additives may disturb the course of a reaction and change the product dis-
tribution. It is this aspect which will be considered here.
3.2.1
Water
In 1980, we established that water was the additive of choice for the SmI2 reduc-
tion of ketones and aldehydes [15]. In fact, there is growing evidence that water
serves not only as a proton source but can also accelerate certain classes of sa-
marium reductions. Addition of water to THF solutions of SmI2 induces a color
change from the original deep blue to a wine-red similar to the one encountered
in a THF–HMPA mixture. The color persists for a few hours if pure and thor-
oughly deaerated water is used. It appears that water acting as a coordinating lig-
and (as well as a proton source) greatly increases the reducing ability of SmI2
180 Henri B. Kagan, Jean-Louis Namy
(see Sect. 3.1.3 and [103]) and THF–H2O can thus be considered to be in some
reactions a good substitute for THF–HMPA.
Curran et al. have studied reductions of 1,3-diphenylpropanone, ethyl cinna-
mate, diphenyl sulfoxide, 1-iodododecane, and o-allyloxyiodobenzene with
SmI2 in the presence of water [98] (Scheme 46). The accelerating role of water in
the reduction of alkyl and aryl iodide has been readily demonstrated [99].
Inanaga et al. reported the use of a 37% aqueous solution of formaldehyde for
the reductive cross-coupling of carbonyl compounds with α,β-unsaturated es-
ters with the SmI2/THF–H2O system [38] (Scheme 47).
Ohgo et al. have achieved a cross-coupling reaction of aldehydes with α-dike-
tones to give the corresponding adducts [110] (Scheme 48); reactions without
HMPA gave better results compared with the reactions using HMPA as additive.
Scheme 46
Scheme 47
Scheme 48
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 181
3.2.2
Alcohols
3.2.2.1
Reduction of Carbonyl and C=N Groups
We have shown that THF–MeOH solutions of SmI2 are excellent reducing agents
of ketones or aldehydes to the corresponding alcohols [15]. A mechanistic study
Scheme 49
Scheme 50
182 Henri B. Kagan, Jean-Louis Namy
with MeOD established that 2-octanone was transformed into 2-D-octanol, pre-
sumably through the intermediate formation of the carbanion obtained by the
reduction of the hydroxy radical issued from the ketyl protonation [114]. Reduc-
tion of a mixture of aldehyde and ketone gave a high preference for the aldehyde
reduction [15].
The combination of SmI2 and trimethylsilyl chloride (in replacement of t-
BuOH) in THF–HMPA was found to accelerate the reduction of sterically hin-
dered or enolizable ketones such as 6-keto- or 20-keto-steroids and 5-cholesten-
one [115].
Isocyanates R–N=C=S have been transformed by SmI2 in good yields into thio-
formamides RNH–(H)C=S in THF–HMPA in the presence of t-BuOH at –78 °C
[116]. The use of t-BuOD gave the C-deuterated thioformamide, establishing
that the reaction occurs by two successive one-electron transfers giving this car-
banion which is trapped by in situ protonation.
3.2.2.2
Pinacol Formation
In 1983 we recognized that aprotic THF solutions of SmI2 gave pinacol forma-
tion from aldehydes or ketones; the reaction being quite fast with aldehydes or
aromatic ketones [117]. However, aromatic carbonyl compounds were coupled
in THF by the samarium/I2/MeOH system, SmI2 being formed in situ [118].
Hanessian et al. discovered that the intramolecular reductive coupling of var-
ious 1,6- and 1,5-dialdehydes gave mainly cis-diols [119]. This useful methodol-
ogy involves the presence of an alcohol (t-BuOH or MeOH) acting as an in situ
protonating agent for the reaction intermediate. For example, the myo-inositol
derivative 29 [120] and 30, an intermediate in the total synthesis of forskolin
[121], have been synthesized (Scheme 51). An intramolecular coupling of α-keto
aldehyde was also successful, giving a bicyclic system 31 which may be a relay in
the total synthesis of taxoids [122].
3.2.2.3
Carbonyl-Ene Couplings
Scheme 51
Scheme 52
184 Henri B. Kagan, Jean-Louis Namy
Scheme 53
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 185
3.2.2.4
Cleavage of Carbon–Heteroatom Bonds
Scheme 54
186 Henri B. Kagan, Jean-Louis Namy
Scheme 55
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 187
Scheme 56
3.2.2.5
Conclusions
Protic additives other than water strongly modify the course of the reactions
mediated by diiodosamarium. As discussed above, the additives protonate some
anionic intermediates to new intermediates, driving the reaction in the desired
direction. In some specific cases it may coordinate to samarium and modify the
redox and Lewis acid properties of some of the samarium complexes. Chiral pro-
tic additives may also act as asymmetric protonating agents of prochiral inter-
mediates. For example, Takeuchi et al. reduced benzil to the corresponding sa-
marium enediolate and treated it with various chiral amino alcohols. The result-
ing benzoin could be obtained with ee's up to 91% by using quinidine as the pro-
ton source [141]. The catalytic use of a C2-symmetric diol for asymmetric pro-
tonation of some samarium enolates was achieved in the presence of trityl alco-
hol (an achiral proton source for regeneration of the catalyst). In this way ee's up
to 93% were observed [142]. Enantioselective protonation of samarium enolates
188 Henri B. Kagan, Jean-Louis Namy
3.2.3
Acids
The combination of SmI2 with a protic or aprotic acid (POCl3, HCl, H2SO4 or
H3PO4) has been investigated by Kamochi and Kudo for the reduction of aro-
matic acids and derivatives [145]. Excellent results were obtained by the addi-
tion of an amount of 85% H3PO4. Thus, at room temperature, acids or esters gave
primary alcohols, while amides ArCONH2 were transformed into aldehydes and
aromatic nitriles into primary amines. All these reactions were performed in
THF in a few minutes. In some cases there were some differences to the corre-
sponding reduction in the presence of a base which was described in Sect. 3.1.4.
A mechanism was proposed by the authors to explain the influence of acids, in-
volving protonation of various intermediates.
Curran and Studer recently used 1 equiv of trifluoroacetic acid or BF3–OEt2
as a useful additive to promote the reductive coupling of aromatic dimethyla-
cetals to 1,2-diaryl-1,2-dimethoxyethanes [112]. If the acids were replaced by
water, a reductive demethoxylation occurred (see Sect. 3.2.1). The added acid
may activate the acetal, giving formation of an oxocarbenium ion which is then
reduced by SmI2 to a radical which couples or is further reduced. It is interesting
to note that a Lewis acid such as BF3–OEt2 is compatible with the use of SmI2.
3.3
Metal Salts
3.3.1
Iron Salts
In 1980 we noticed that a samarium Barbier reaction between n-butyl iodide and
2-octanone which needed 8 h at 65 °C could be run in 3 h at room temperature
in the presence of a catalytic amount (2 mol%) of ferric chloride [15]. This ex-
periment was inspired by the known catalytic effect of the Fe3+ ion on the
Ln2+/Ln3+ conversion [146]. The beneficial influence of the Fe(III) derivative in
reactions mediated by SmI2 has been used by several authors in various process-
es. Thus, Molander et al. described a general procedure for five- and six-mem-
bered ring annulation starting from 2-(ω-iodoalkyl) cycloalkanones [147].
Diiodosamarium induced an intramolecular Barbier reaction, under mild
conditions (room temperature), in the presence of a catalytic amount of iron
tris(dibenzoylomethane [Fe(DBM)3]. In this way many bicyclic tertiary alco-
hols were prepared. These conditions were also convenient for the cyclization
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 189
Scheme 57
3.3.2
Nickel Salts
Scheme 58
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 191
3.3.3
Cobalt Salts
In 1991, Inanaga et al. described the reduction at room temperature of some dis-
ubstituted alkynes by the combination SmI2/proton source in the presence of
some transition-metal catalysts (3% equiv) in THF [155]. By a good choice of
catalyst (CoCl2, 4 PPh3) and proton source (MeOH, i-PrOH or AcOH) it was pos-
sible to orientate the reaction towards the exclusive formation of the Z-alkene.
When the same reaction was performed in the presence of HMPA then the E-
alkene was produced. Iron(III) and Ni(II) catalysts were found to be less effi-
cient. It was assumed that the reactive species were the corresponding transi-
tion-metal hydrides obtained by reduction of the initial complexes.
3.3.4
Copper Salts
3.3.5
Chromium Salts
Scheme 59
192 Henri B. Kagan, Jean-Louis Namy
3.3.6
Palladium Complexes
In 1986 Inanaga et al. found that allylic or propargylic acetates could be reduced
in THF by diiodosamarium in the presence of a catalytic amount of a Pd(0) com-
plex and 1 equiv of 2-propanol [160,161]. Some examples are indicated in
Scheme 60. The mechanism of the reaction likely involves π-allyl (or σ-allenyl)
palladium intermediates which are reduced first to radicals and then to carban-
ions (see Scheme 60). A final protonation by the alcohol generates the products.
Scheme 60
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 193
Some limitations to this reaction have been noticed [162]. The reductive cou-
pling of allylic acetates with carbonyl compounds (which act as electrophiles in
replacement of a proton donor) provided homoallylic alcohols [163]. Similarly,
coupling of propargylic esters to carbonyl compounds gave allenyl alcohols or
acetylenic alcohols according to the structure of the reactants [164,165]. Some
examples can be found in Scheme 60. The intramolecular coupling reaction us-
ing ω-keto-alkynyl esters allowed the synthesis of five- or six-membered ring
systems [166]. A highly regio- and stereoselective synthesis of allylsilanes was
achieved by reductive silylation of allylic phosphates with a SmI2/HMPA/Pd(0)
system and TMSCl [167].
A regiodivergent reduction of allylic esters with SmI2 in the presence or ab-
sence of a Pd(0) catalyst gave either α- or γ-protonated products by tuning the
proton source (alcohols, H2O) and the ester functionality (Scheme 61) [168].
A changeover of regioselectivity was observed by Mikami et al. in the reduc-
tion of secondary propargylic phosphates by the SmI2/Pd(0)/proton source sys-
tem [169]. tert-Butanol and dimethyl (R,R)-tartrate gave allene and acetylene,
respectively (Scheme 61). This process was modified in an asymmetric synthesis
of chiral allenes [170].
2,3-Naphthoquinodimethanes were easily generated from ortho-bis(α-ace-
toxypropargyl) benzene derivatives by the SmI2/Pd(0)/proton donor system
[171]. The reduction occurs by a 6π-electron cyclization of an O-diallenylben-
zene intermediate.
3.3.7
Lithium Salts
The Barbier reaction discussed in Sect. 3.3.2 was not accelerated by addition of
LiCl. In the literature there are only a few examples where lithium salts are ben-
Scheme 61
194 Henri B. Kagan, Jean-Louis Namy
Scheme 62
eficial for reactions mediated by SmI2. However, it is worth mentioning that a re-
ductive dialkylation of isoindigo by cis-1,4-dichloro-2-butene only works in the
presence of an excess of LiCl (KCl is inefficient) as indicated in Scheme 62 [172].
The mechanism was not elucidated; a chelate may be involved.
Recently, Flowers et al. initiated a study on the influence of additives on the
reactivity of SmI2, in parallel with electrochemical investigations (see
Sect. 3.1.1). These authors found that the oxidation potentials of SmI2 in THF
containing 12 or more equivalents of LiBr or LiCl were, respectively, –1.98 and –
2.11 V (instead of –1.33 V). They established that the corresponding molecular
species were not SmBr2 and SmCl2. These new reagents are highly reactive and
facilitate the pinacol coupling of cyclohexanone at room temperature in a few
minutes at room temperature [173].
4
Conclusions
In this chapter we have tried to present the main changes which occur for reac-
tions mediated by diiodosamarium when THF is replaced by other solvents or
when some additives are introduced into the THF solution. Many useful organic
transformations have been achieved by this approach but, due to lack of space,
this area has not been covered extensively. However, we hope to have demon-
strated that introduction into THF of various additives in stoichiometric or cat-
alytic amounts or replacement of THF by another solvent may drastically mod-
ify the behavior of SmI2. The most promising approach involves the use of mo-
lecular catalysts since one may expect a wide range of chemo- or stereoselectiv-
ities by suitable changes in the structure of the catalysts.
5
References
1. Matignon CA, Cazes E (1906) Ann Chim Phys 8:417
2. Jantsch G, Skalla N, Jawurek L (1931) Z Anorg Chem 201:207
3. Namy J-L, Girard P, Kagan HB (1977) New J Chem 1:5
4. Girard P, Namy J-L, Kagan HB (1980) J Am Chem Soc 102:2693
Influence of Solvents or Additives on the Organic Chemistry Mediated by Diiodosamarium 195
5. Kagan HB (1985) In: Marks TJ, Fragala IL (eds) Fundamental aspects of organo-f-ele-
ments chemistry. Reidel, Dordrecht, p 49
6. Kagan HB, Namy J-L (1986) Tetrahedron 42:6573
7. Kagan HB (1990) New J Chem 14.453
8. Soderquist JA (1991) Aldrichimica Acta 24:15
9. Molander GA (1992) Chem Rev 92.15
10. Molander GA, Harris CR (1996) Chem Rev 96:307
11. Skrydstrup T (1998) Angew Chem Int Ed Engl 36:345
12. Molander GA (1994) Organic Reactions 46:211
13. Inanaga J (1990) Rev Heteroatom Chem 3:75
14. Namy J-L, Girard P, Kagan HB, Caro P (1981) New J Chem 5:479
15. Matsukawa M, Inanaga J, Yamaguchi M (1987) Tetrahedron Lett 28:5877
16. Otsubo K, Kawamura M, Inanaga J, Yamaguchi M (1987) Chem Lett 1487
17. The stoichiometry of bonded THF molecules to SmI2 is 5, the removal of coordinated
THF has been discussed [18]
18. Evans W, Gummersheimer TS, Ziller JW (1995) J Am Chem Soc 117:8999
19. Murakami M, Hayashi M, Ito Y (1992) J Org Chem 57:793
20. Murakami M, Hayashi M, Ito Y (1995) Appl Organomet Chem 9:385
21. Booth SE, Benneche T, Undheim K (1995) Tetrahedron 51:3665
22. Namy J-L, Colomb M, Kagan HB (1994) Tetrahedron Lett 35:1723
23. Hamann-Gaudinet B, Namy J-L, Kagan HB (1997) Tetrahedron Lett 38:6585
24. Matsukawa M, Inanaga J, Yamaguchi M (1987) Tetrahedron Lett 28:5877
25. Imamoto J, Takeyama T, Yokoyama M (1984) Tetrahedron Lett 25:3225
26. Zoretic PA, Yu BC, Caspar L (1989) Synthetic Commun 1859
27. Collin J, Namy J-L, Dallemer F, Kagan HB (1991) J Org Chem 56:3118
28. Ruder SM (1992) Tetrahedron Lett 33:2621
29. Hamann B, Namy J-L, Kagan HB (1996) Tetrahedron 52:14225
30. Chebolu A, Whittle RR, Sen A (1985) Inorg Chem 24:3082
31. Inanaga J, Ujikawa O, Yanaguchi M (1991) Tetrahedron Lett 32:1737
32. Ying T, Bao W, Zhang Y, Xu W (1996) Tetrahedron Lett 37:3885
33. Akane N, Kanegawa Y, Nishiyama Y, Ishii Y (1992) Chem Lett 2431
34. Akane N, Hutano T, Kusui H, Nishiyama Y, Ishii Y (1994) J Org Chem 59:7902
35. Kunishima M, Hioki K, Ohara T, Tani S (1992) J Chem Soc Chem Commun 219
36. Kunishima M, Hioki K, Tani S, Kato A (1994) Tetrahedron Lett 35:7253
37. Kunishima M, Hioki K, Kono K, Kato A, Tani S (1997) J Org Chem 62:7542
38. Otsubo K, Inanaga J, Yamaguchi M (1986) Tetrahedron Lett 27:5763
39. Fukuzawa S-I, Ilida M, Nakawishi A, Fiyinami T, Sakai SJ (1987) J Chem Soc Chem
Commun 920
40. Otsubo K, Inanaga J, Yamaguchi M (1987) Tetrahedron Lett 28:4437
41. Inanaga J, Ishikawa M, Yamaguchi M (1987) Chem Lett 1485
42. Ujikawa O, Inanaga J, Yamaguchi M (1989) Tetrahedron Lett 30:2837
43. Handa Y, Inanaga J, Yamaguchi M (1989) J Chem Soc Chem Commun 298
44. Hou Z, Wakatsuki Y (1994) J Chem Soc Chem Commun 1205
45. Shabangi M, Flowers II RA (1997) Tetrahedron Lett 38:137
46. Imamoto T, Yamanoi Y, Tsuruta H, Yamaguchi K, Yamazaki M, Inanaga J (1995) Chem
Lett 949
47. Molander GA, Mc Kie JA (1992) J Org Chem 57:3132
48. Skene WG, Scaiano JC, Cozens FL (1996) J Org Chem 61:7918
49. Hasegawa E, Curran DP (1993) Tetrahedron Lett 34:1717
50. Curran DP, Totleben MJ (1992) J Am Chem Soc 114:6050
51. Totleben MJ, Curran DP, Wipf P (1992) J Org Chem 57:1740
52. Wipf P, Venkatraman S (1993) J Org Chem 58:3455
53. Molander GA, Mckie JA (1994) J Org Chem 59:3186
196 Henri B. Kagan, Jean-Louis Namy
Abbreviations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
Abbreviations
Ln lanthanoid metal ion (III)
M alkali metal ion
L lithium
S sodium
P potassium
B 1,1'-binaphthol (BINOL)
LnMB lanthanoid-alkali metal-binaphthoxide complex of type LnM3tris(bi-
naphthoxide)
LLB lanthanum-lithium-binaphthoxide complex LaLi3tris(binaphthoxide)
1
Introduction
One of the most fascinating aspects in the history of asymmetric catalysis with
its countless successful applications in the stereoselective synthesis of a broad
variety of functional groups is the structural variety of the complexes which are
able to be used as catalysts [1, 2]. Many catalysts have been developed based on
different ideas and concepts of mechanistic effect. However, in spite of the abun-
dance of such catalysts which have been successfully applied in asymmetric ca-
talysis, not a handful of them possess multifunctional abilities catalyzing differ-
ent type of enantioselective reactions. The development of such a type of chiral
catalyst, the catalytic effect of which is not limited to one reaction but to differ-
ent types of asymmetric synthetic organic transformations, remained an attrac-
tive challenge for a long time.
In the following, such a desired new and innovative multifunctional catalytic
system is reviewed: The chiral heterobimetallic lanthanoid complexes, devel-
oped by Shibasaki et al., have recently been shown to catalyze a broad spectra of
organic reactions including many “classical” carbon-carbon bond formations
Chiral Heterobimetallic Lanthanoid Complexes 201
like the nitroaldol reaction, Michael addition, aldol reaction and so on, but also
an oxidation reaction as well as the asymmetric formation of carbon-phospho-
rus bonds with excellent stereoselectivity in all cases (Fig. 1) [3–5]. This new
concept for catalytic asymmetric reactions, based on the idea of using chiral het-
erobimetallic lanthanoid complexes which function as both Brønsted base and
Lewis acid, just like an enzyme, makes possible a variety of efficient catalytic
asymmetric reactions.
2
Structural Requirements for an Efficient Bimetallic Catalyst:
What does the Catalysts Look Like?
For a better understanding of many successful applications of the rare earth-al-
kali metal containing LnM3tris(binaphthoxide) complexes (LnMB, Ln=rare
earth, M=alkali metal) to catalytic asymmetric synthesis, intense investigations
also focused on the determination of the structure. It has been shown that these
complexes, which can be readily prepared from the corresponding rare earth
trichlorides and/or rare earth isopropoxides [5], possess a structure as present-
ed in Fig. 2. This structure was supported by various NMR spectroscopic, MS
spectrometric, X-ray crystallographic and other analytic investigations of a va-
riety of LnMB complexes.
Beginning with LDI-TOF mass spectral analysis, these investigations revealed
that the structure was in fact a heterobimetallic complex consisting of one lan-
thanum, three lithium and three BINOL moieties [6]. Although LDI-TOF mass
Chiral
Heterobimetallic
Lanthanoid Catalysts
LnMB
M O O M
O Ln O
O O
M
(R)-LnMB
3
Asymmetric Catalytic C-C Bond Formation Using Heterobimetallic Lanthanoid
Complexes
3.1
Nitroaldol Reaction
3.1.1
Development of an Efficient Catalysis in Model Reactions
In 1992, Shibasaki et al. reported for the time an application of chiral heterobi-
metallic lanthanoid complexes (LnLB) as chiral catalysts in asymmetric cataly-
sis, namely the catalytic asymmetric nitroaldol reaction (Henry reaction),
which is one of the most classical C-C bond forming processes [11]. Additional-
ly, this work represents the first enantioselective synthesis of β-nitroalcohol
compounds by the way of enantioselective addition of nitroalkanes to aldehydes
in the presence of a chiral catalyst. The chiral BINOL based catalyst was prepared
starting from anhydrous LaCl3 and an equimolar amount of the dialkali metal
salt of BINOL in the presence of a small amount of water [9].
Starting from prochiral aldehydes 1–3, the desired products b were obtained
in good chemical yields and with enantioselectivities up to 90% ee (Scheme 1)
[11]. The amount of the catalyst is not shown in Scheme 1 due to the unknown
structure of the catalyst (at this time).
Investigations concerning the influence of the rare earth metal component
showed pronounced differences both in the reactivity and in the enantioselec-
204 Masakatsu Shibasaki, Harald Gröger
La-Li-(S)-BINOL complex OH
RCHO + CH3NO2 NO2
THF, -42 °C, 18 h R
(10 equiv)
Scheme 1. The first catalytic asymmetric nitroaldol reaction catalyzed by chiral lanthanoid
complexes.
ee (%) 100 OH
NO2
6
80 OH
NO2
Ph
60 4
OH
40 NO2
Ph
7
20
Eu
Gd Pr
Yb Y Tb Sm Nd La
0
0.8 0.9 1.0 1.1
Figure 3. Effects of the ionic radii of rare earth elements on the enantioselectivity.
tivity among the various rare earth metals used [12]. When benzaldehyde and
nitromethane were used as starting materials, the corresponding Eu complex
gave 7 in 72% ee (91%) in contrast to 37% ee (81%) in the case of the La complex
(–40 ˚C, 40 h). The unique relationship between the ionic radii of rare earth met-
als and the enantioselectivities of several nitroaldols 4, 6, 7 are depicted in Fig. 3.
Consequently, small changes in the structure of the catalyst (ca. 0.1 Å in ionic
radius of the rare earth cation) cause drastic changes in the optical purity of the
produced nitroaldols. Although in general nitroaldol reactions are regarded as
equilibrium processes, no detectable retro-nitroaldol reaction was observed in
the Ln-BINOL complex catalyzed asymmetric nitroaldol reactions.
The proposed mechanism for the asymmetric nitroaldol reaction catalyzed
by heterobimetallic lanthanoid complexes is shown in Scheme 2 [5]. In the ini-
tial step, the nitroalkane component is deprotonated and the resulting lithium
nitronate coordinates to the lanthanoid complex under formation of the inter-
Chiral Heterobimetallic Lanthanoid Complexes 205
Li *
O
O O OH
* La Li NO2
CH3NO2 O O R'
O
Li *
(R)-LLB
Li Li
* * * *
O O O O
O La O O La O H
Li O O H+ Li O O
R'
O N CH O O
Li 2 + H
* O— * Li N CH2
I O—
II
R'CHO
Scheme 2. A possible mechanism for catalytic asymmetric nitroaldol reactions with nitro-
methane.
Recently, Okamoto et al. showed that the reactivity and selectivity also de-
pends on the alkali metal component in the heterobimetallic catalysts [18]. Us-
ing the bulkier 2-nitropropane as starting material in a model reaction with ben-
zaldehyde, almost no reaction occurred at –30 ˚C in the presence of the lithium
containing catalyst LLB, whereas higher temperatures as well as the use of
HMPA as a co-solvent only led to racemic product. However, in the presence of
the corresponding potassium-containing catalyst LPB the desired reaction pro-
ceeded with 46% ee. In contrast, the use of LLB was connected with superior
enanioselectivity and chemical yield (compared to LPB) when replacing 2-nitro-
propane by the less bulkier nitromethane (LLB: 91% yield; 48% ee; LPB: 71%
yield; 6% ee) [18].
3.1.2
Enantioselective Catalytic Nitroaldol Reaction: Adducts with One Stereogenic Center
site sense to the enantiofacial selectivity which might have been expected on the
basis of the previous results (cf. Scheme 1). These results suggested that the pres-
ence of an oxygen atom at the β-position had a pronounced influence on the
enantiofacial selectivity.
The LLB type catalysts were also successfully applied in the asymmetric ni-
troaldol reaction of quite unreactive α,α-difluoro aldehydes. However, catalytic
asymmetric nitroaldol reaction of a broad variety of α,α-difluoro aldehydes
proceeded satisfactorily when using the heterobimetallic asymmetric catalysts
with modified, 6,6'-disubstituted BINOL ligands [22]. The best results were ob-
tained with the samarium (III) complex (5 mol%) generated from 6,6'-bis{(tri-
ethylsilyl)ethynyl}BINOL with enantioselectivities up to 95% ee. The (S)-config-
uration of one representative nitroaldol adduct showed that the nitronate react-
ed preferentially on the Si face of aldehyde in the presence of (R)-LLB (20 mol%;
74% yield; 55% ee). It is noteworthy that the enantiotopic face selection for α,α-
difluoro aldehydes is reverse to that for nonfluorinated aldehydes. The stereose-
lectivity for α,α -difluoro aldehydes is identical with that of β-oxa-aldehydes,
suggesting that the fluorine atoms at the α-position have a great influence on
enantioface selection.
3.1.3
Diastereoselective Catalytic Nitroaldol Reaction Starting from Chiral Aldehydes
Scheme 4. Diastereoselective nitroaldol reaction as key step in the synthesis of erythro-AHPA 19.
Scheme 5. Catalytic diastereoselective nitroaldol reaction promoted by the LPB type catalyst.
Chiral Heterobimetallic Lanthanoid Complexes 209
3.1.4
Diastereoselective and Enantioselective Nitroaldol Reaction
LnLB type catalysts are also able to promote diastereoselective and enantioselec-
tive nitroaldol reactions starting from prochiral materials. In preliminary work,
LLB gave unsatisfactory results in terms of both diastereoselectivity (syn/anti
ratio 63 : 37 to 77 : 23) and enantioselectivity (<78% ee) [19]. However, an effec-
tive asymmetric induction was obtained in the presence of LL(B-a) type cata-
lysts containing 6,6'-substituted BINOL (Scheme 6; for the structure of B-a, see
Scheme 7).
The application of the catalysts of type LL(B-a) (3.3 mol%) to diastereoselec-
tive nitroaldol reactions led to high syn-selectivity and enantioselectivity [29].
In all cases, much higher syn-selectivity (syn/anti ratio up to 94:6) and enanti-
catalyst OH OH
(3.3 mol %) R' R'
RCHO + R'CH2NO2 R + R
THF, -40 °C syn NO2 anti NO2
1: R = PhCH2CH2 23: R' = Et 25(syn), 26 (anti): R = PhCH2CH2, R' = Et
22: R = CH3(CH2)4 24: R' = CH2OH 27(syn), 28 (anti): R = PhCH2CH2, R' = CH2OH
29(syn), 30 (anti): R = CH3(CH2)4, R' = CH2OH
OH
catalyst (10 mol %)
CH3(CH2)14CHO OH CH3(CH2)14
+ O2N OH
-40 °C, 163 h
32 NO2
33 + anti-adduct
OH
Et3SiC≡C LL(B-a) catalyst:
H2, Pd-C 6
CH3(CH2)14 OH 78% (syn/anti = 91:9),
EtOH OH syn: 97% ee
NH2
OH
syn-dihydrosphingosine 31 LLB catalyst:
6'
Et3SiC≡C 31% (syn/anti =
86:14),
B-a syn: 83% ee
oselectivity (up to 97% ee) was obtained using the catalysts with 6,6'-substituted
BINOL instead of LLB (for comparison representative results are given in
Scheme 6). The optical purities of the minor anti-adducts 26, 28, and 30 were
lower than those of the syn-adducts 25, 27, and 29, indicating that the former
were not generated by epimerization of the nitro group. In fact, treatment of the
syn-adducts with catalysts such as LLB and its derivatives resulted in near-quan-
titative recovery of the starting materials with unchanged optical purities.
The syn-selective asymmetric nitroaldol reaction was successfully applied to
the catalytic asymmetric synthesis of threo-dihydrosphingosine 31, which elicits
a variety of cellular responses by inhibiting protein kinase C (Scheme 7) [30]. Ni-
troaldol reaction of hexadecanal 32 with 3 equiv of nitroethanol catalyzed by
LL(B-a) gave the corresponding nitroaldol adduct 33 in high syn-selectivity
(91:9) and 78% yield, with the syn-adduct 33 being obtained with up to 97% ee
[29]. In this case, under similar conditions the LLB-catalyzed reaction proceed-
ed only slowly to give an 86:14 ratio of the syn and anti-adducts in 31% yield
(with lower optical purity: 83% ee). The hydrogenation of 33 in the presence of
10% Pd on charcoal afforded threo-dihydrosphingosine 31 in 71% yield.
3.1.5
Tandem Inter-Intramolecular Catalytic Asymmetric Nitroaldol Reaction
O
OH O
O CH3NO2 (10 equiv)
(R)-PrLB (5 or 10 mol %) O2N *
OHC
* O2N * O*
THF, -40 °C OH
O
O
34 (S)-35a 35b
O O
room temperature
HO HO
OH OH
O2N O2N
36a 36b
up to 45% yield (cryst. product)
up to 65% ee (crude product)
up to 79% ee (cryst. product)
3.1.6
Recent Improvements (Second Generation of LnLB Catalyst) and Summary
OH
catalyst (1 mol %) R'
RCHO + R'CH2NO2 R
NO2
3: R = C6H11 37: R' = H 6: R = C6H11, R' = H
1: R = PhCH2CH2 38: R' = CH3 39: R = PhCH2CH2, R' = CH3
23: R' = Et 25: R = PhCH2CH2, R' = Et
Scheme 9. Comparison of the catalytic activity of LLB and second-generation LLB (LLB-II)
or LL(B-a) and LL (B-a)-II. [LLB-II: LLB + H2O (1 mol equiv) + BuLi (0.9 mol equiv.);
LL(B-a)-II: LL(B-a) + H2O (1 mol equiv) + BuLi (0.9 mol equiv)]; For the structure of the
BINOL derivative ligand B-a, see Scheme 7
OH
H
HO N
HO OH
40: arbutamine
64% overall yield
utamine 40, which is a useful β-agonist [33, 34], was achieved (Scheme 10) [8]. In
the key step, the nitroaldol adduct 42 was formed in 93% yield and with 92% ee [8].
In conclusion, since the first example of a catalytic asymmetric nitroaldol re-
action (Henry reaction) was reported in 1992 by Shibasaki et al., this reaction
has been developed into a highly efficient synthetic method for the stereoselec-
tive synthesis of nitroalkanols. Using alkali metal-containing heterobimetallic
Chiral Heterobimetallic Lanthanoid Complexes 213
3.2
Direct Asymmetric Aldol Reaction
3.3
Michael Addition Reaction
Catalytic asymmetric Michael reactions are one of the most important synthetic
methods for obtaining asymmetric centers by enantioselective construction of
carbon-carbon bonds [39, 40]. The first lanthanoid complex catalyzed Michael
(R)-LLB OH O
O (20 mol %)
R1CHO + R1 R2
H 3C R2 -20 °C, THF
yields up to 90%
ee up to 94%
R2
O M O O
O H
*
1
2 O LA R H
R II
R2
O M O M O
H
H
*
*
O LA I O LA O III
R1
R2
OH O
O LA : Lewis acid
1 O M
R R2 M : Metal of Brønsted
H base
O
*
O LA R1 O
IV : Chiral ligand
*
O
O
O LSB
1
COOR (10 mol %)
+ R2 COOR1
( )n
( )n COOR1 THF
R2 COOR1
n = 1,2 R1 = Et, Me, Bn 43
R2 = H, CH3 up to 98% yield
up to 92% ee
O LnSB O Ph
COOCH3 (10 mol %) H
+ COOCH3
Ph Ph COOCH3 toluene, Ph
-50°C COOCH3
44 45
Ln=La: 93% yield; 77% ee
Ln=Pr: 96% yield; 56% ee
Ln=Gd: 54% yield; 6% ee
addition reaction was realized using malonates as Michael donor in the catalytic
asymmetric Michael reaction of various enones [41, 42]. Although ineffective as
an asymmetric catalyst for nitroaldol reactions, LSB (L=lanthanum, S=sodium)
was found to be effective in this case, giving Michael adducts 43 in up to 92% ee
and almost quantitative yield (Scheme 13) [41]. In general, the use of THF as a
solvent led to the best results, whereas in the case of the corresponding LSB-cat-
alyzed reaction of trans-chalcone 44 the use of toluene was essential to give the
product in good enantiomeric excess. Center metal effects were also investigated
Chiral Heterobimetallic Lanthanoid Complexes 215
for this Michael reaction, indicating that LSB was the best catalyst for catalytic
asymmetric Michael reactions.
In order to clarify the nature of the interaction between the enone and the
asymmetric catalyst, the complexation was studied by 1H-NMR spectroscopy af-
ter mixing cyclohexenone and the asymmetric bimetallic complexes and ob-
serving the chemical shift of the α-proton of cyclohexenone [41]. Interestingly,
it was found that complexation with LSB induced a small downfield shift on the
α-proton of cyclohexenone whereas PrSB, a moderately effective asymmetric
catalyst for Michael reactions, induced a large upfield shift. In strong contrast,
in the case of either EuSB or LLB, which gave only near-racemic Michael ad-
ducts, the 1H-NMR spectra showed no changes in chemical shift of the α-proton
of cyclohexenone. These NMR studies indicated that the carbonyl group of the
enone coordinated to lanthanum and/or praseodymium metal in the LnSB mol-
ecule, while the enone did not coordinate to either LLB and/or EuSB. These
changes of chemical shift were observed even in the presence of dimethyl
malonate. The chemical phenomena described above might be understood by
considering the differing dihedral angles of binding of the BINOL moiety to the
center metal in each case.
Additionally, computational simulations of the enantioselection process us-
ing Rappé's universal force field (UFF) [43–45] were carried out which clearly in-
dicated that the (R)-LSB complex complexes better as a pro-(R) adduct than as
pro-(S) adduct (∆E=4.9 kcal/mol) [41]. The proposed catalytic cycle is shown in
Scheme 14. Thus, the basic LSB complex also acts as a Lewis acid, controlling the
orientation of the carbonyl function and so activating the enone to attack. It ap-
O
O
COOCH3 Na *
+ O
O O COOCH3
COOCH3 * La Na
O O COOCH3
Na O *
LSB
Na *
OCH3 O O
* O O Na
Na * La O
O Na O
O O O OCH3
* La O Na O OH
O OH * OH O
Na O
* OCHOCH
3 3
I II
Scheme 14. Proposed catalytic cycle for the asymmetric Michael reaction promoted by LSB.
216 Masakatsu Shibasaki, Harald Gröger
pears that the multifunctional nature of the LSB catalyst makes possible the for-
mation of Michael adducts with high ees even at room temperature. As final step
the resultant sodium enolates II of the optically active Michael adducts appear
to abstract a proton from an acidic OH so as to regenerate the LSB catalyst. In
both catalytic asymmetric Michael reactions and nitroaldol reactions, enones
and/or aldehydes appear to coordinate to the rare earth metal. The reason why
LSB is more effective for catalytic asymmetric Michael reactions whereas LLB is
more effective for catalytic asymmetric nitroaldol reactions is still unclear at
present. However, it seems that slight differences in bond lengths in chelate
structures such as I and II, as well as slight differences in “bite” angle for the BI-
NOL moiety caused by varying the alkali metal component may be responsible
for this effect.
Another type of an LSB catalyzed asymmetric Michael reaction, in which the
asymmetric center is induced on the side of the adduct originating from the
Michael donor, was also reported by Shibasaki et al. (Scheme 15) [46]. In a pre-
liminary study, it was found that the reaction of 46 with 3-buten-2-one in THF
using 10 mol% of LSB gave 47 with 23% ee, while carrying out the reaction in tol-
uene afforded 47 with 75% ee. However, when the amount of LSB was reduced to
5 mol%, the enantiomeric excess of 47 declined to a more modest 25% ee. To off-
set this decline while still maintaining the lower level of catalyst slow addition of
46 was essential. Accordingly, the use of syringe pump methods gave 47 with
high enantiomeric excess (89% ee). The reaction was further improved when us-
ing CH2Cl2 as a solvent. Thus, the asymmetric Michael reaction catalyzed by
5 mol% of LSB in CH2Cl2 gave 47 in 89% yield and with 91% ee, without the need
for slow addition previously encountered (Scheme 15). In addition, in this case
the catalytic asymmetric Michael reaction for 47 was not so affected by the
choice of rare earth metal.
In conclusion, as shown in Scheme 15, slow addition of β-keto ester and the
use of CH2Cl2 as solvent are generally quite effective methods for preventing re-
duction of enantiomeric excess for the various Michael adducts. On the other
solvent yield ee
toluene 83% 25%
toluenea) 76% 89%
CH2Cl2 89% 91%
a) Slow addition of nucleophile with
syringe-pump methods over 8h
hand, malonates give the adducts with high ees regardless of the solvent used
[41]. These results can be rationalized by comparison of the pKa of a β-keto ester
with that of a malonate; the former is significantly more acidic than the latter.
Thus, the concentration of the resulting Na-enolate can be expected to be greater
in the case of the β-keto ester, and moreover this Na-enolate will react with an
enone more slowly than the Na-enolate derived from a malonate. We suggest
that this combination of more rapid formation and longer lifetime increases the
likelihood of dissociation of the Na-enolate from the chelated ensemble, thus
giving a product of lower ee. On the other hand, in less polar CH2Cl2 the Na-eno-
late would, even in this case, remain part of the ensemble, thereby affording the
product with high ee (Fig. 4). Furthermore, it appears that slow addition of the
β-keto ester also acts to limit undesired ligand exchange between BINOL moie-
ties and the Michael donor.
In both types of catalytic asymmetric Michael reactions, the use of either sec-
ond-generation LSB or 6,6'-substituted BINOL derived LSB type catalysts did
not result in significantly improved results.
3.4
Diels-Alder Reaction
in CH2Cl2 in THF
Na * O O
Na * O O
O Na O O Na
O O O OEt * La O
La O O O
* O OH
O OH O
O OEt
Na *
Na *
high ee low ee
Figure 4. Proposed mechanism for the catalytic asymmetric Michael reaction promoted by
LSB.
218 Masakatsu Shibasaki, Harald Gröger
O O catalyst
(10 mol %) H
N O + N
—20 °C, 20 h O
O
48 O
49
(R)-LLB (R)-LLB*a)
yield: 82% 100%
endo/exo: 15:1 36:1
ee: 63% 86%
a) LLB*=LaL3tris(6,6’-dibromobinaphth-
oxide)
4
Catalytic Asymmetric C-O Bond Formation Using Heterobimetallic
Lanthanoid Complexes: Epoxidation
Catalytic asymmetric epoxidations are one of the most important asymmetric
processes [1]. In addition to previous successful achievements of other groups
with allylic alcohols [50, 51] and unfunctionalized olefins [52–55], the first effi-
cient catalytic asymmetric epoxidation of a variety of enones with broad gener-
ality has been developed when using chiral lanthanoid complexes [56].
In the presence of the sodium-containing heterobimetallic catalyst (R)-LSB
(10 mol%), the reaction of enone 52 with TBHP (2 equiv) was found to give the
desired epoxide with 83% ee and in 92% yield [56]. Unfortunately LSB as well as
other bimetallic catalysts were not useful for many other enones. Interestingly,
in marked contrast to LSB an alkali metal free lanthanoid BINOL complex,
which was prepared from Ln(O-i-Pr)3 and (R)-BINOL or a derivative thereof (1
or 1.25 molar equiv) in the presence of MS 4A (Scheme 17), was found to be ap-
plicable to a range of enone substrates. Regarding enones with an aryl-substitu-
ent in the α-keto position, the most effective catalytic system was revealed when
using a lanthanum-(R)-3-hydroxymethyl-BINOL complex La-51 (1–5 mol%)
and cumene hydroperoxide (CMHP) as oxidant. The asymmetric epoxidation
proceeded with excellent enantioselectivities (ees between 85 and 94%) and
yields up to 95%.
Chiral Heterobimetallic Lanthanoid Complexes 219
R
50: R = H
OH (1 equiv)
Ln(O-i-Pr)3 + OH 51: R = CH2OH
Ln = La or Yb (1.25 equiv)
MS 4A
THF, rt
In contrast to this result, the enones with an alkyl moiety in the α-keto posi-
tion were best converted to the corresponding epoxides when replacing lantha-
num by ytterbium in the corresponding catalytically active complex of the above
mentioned type. Using TBHP (1.5 equiv) as oxidant in the presence of 5 mol%
of Yb-51 gave enantioselectivities up to 94%. However, the use of either Yb-50
catalyst or a La-CMHP system afforded the product of type 53 with less satisfac-
tory results. It seems likely that the difference in ionic radius between lanthanum
and ytterbium, as well as the difference in Lewis acidities, accounts for the ob-
served center metal effects.
Although the structure of the catalytically active species could not have been
unequivocally determined, it was found that an almost 1:1 ratio of Ln(O-i-Pr)3
(Ln=La or Yb) and BINOL gave the maximum enantiomeric excesses. In addi-
tion, an asymmetric amplification of the catalytic asymmetric epoxidation has
been obtained (Fig. 5) [56], which led to the conclusion that an oligomeric struc-
ture of these lanthanoid-BINOL catalysts may play a key role in these catalytic
asymmetric epoxidations of enones. Regarding the mechanism, a Ln-alkoxide
moiety in the catalysts appears to act as a Brønsted base, activating a hydroper-
oxide moiety so as to make possible a Michael reaction. At the same time anoth-
er Ln metal ion seems to act as a Lewis acid, both activating and controlling the
orientation of the enone. Furthermore, it is noteworthy that this catalytic asym-
metric epoxidation can be carried out at room temperature using 1–8 mol% of
an asymmetric catalyst to give epoxides with good enantiomeric excesses.
220 Masakatsu Shibasaki, Harald Gröger
100
ee of 53a (%)
80
O
60 O
Ph CH3
40
53a
20
0
0 20 40 60 80 100
ee of 50 (%)
5
Catalytic Asymmetric C-P Bond Formation Using Heterobimetallic Lanthanoid
Complexes
The application of the heterobimetallic lanthanoid complexes of the LnMB type
led to a breakthrough in establishing a highly efficient asymmetric catalytic
route to α-hydroxy as well as α-amino phosphonic acid esters, which have at-
tracted much attention due to their wide ranging biological activity [57–64]. The
heterobimetallic catalysis described below represents the first and until now the
only highly efficient asymmetric catalytic approach to both α-hydroxy and α-
amino phosphonates by the attractive way of asymmetric catalytic hydrophos-
phonylation.
5.1
Hydrophosphonylation of Aldehydes
O OH
O (R)-LLB (10 mol %)
+ H P(OCH3)2 R P(OCH3)2
R H THF, -78 °C
O
R = Ph, p-R’C6H4,
(E)-R’CH=CR’’, 54
up to 94% yield
alkyl up to 95% ee
Li *
O
O O
O * La Li
O O OH
HP(OCH3)2 O
Li * R P(OCH3)2
LLB
O
Li Li
* * * *
O O O O
O La O O La O H R
Li H Li O O
O OO
OO H
P(OCH3)2
* Li * Li P(OCH3)2
RCHO
5.2
Hydrophosphonylation of Imines
5.2.1
Hydrophosphonylation of Acyclic Imines
α-Amino phosphonic acids are interesting compounds for the use in the design
of enzyme inhibitors [59, 60]. As in the case of α-hydroxy phosphonic acids, the
absolute configuration of the α-carbon strongly influences the biological prop-
erties. Although several (especially diastereoselective) methods for the synthesis
of optically active α-amino phosphonic acids have been known for a long time
[71, 72], the first catalytic asymmetric hydrophosphonylation of imines has
been reported recently by Shibasaki et al. [73] using potassium-containing
LnK3tris(binaphthoxide) complexes (LnPB) as most efficient catalysts.
Interestingly, almost the first results in asymmetric hydrophosphonylation
with acyclic imines revealed that the lanthanum-potassium-BINOL catalyst LPB
was more effective than the analog sodium and lithium complexes LSB and LLB,
both of which have been shown to be highly efficient in asymmetric C-C bond
formations (see Sect. 3). As a representative example, in the presence of LPB
(20 mol%) and 5 equiv of dimethyl phosphite the hydrophosphonylation of im-
224 Masakatsu Shibasaki, Harald Gröger
ine 57a proceeds at room temperature under formation of the desired product
in 91% ee, whereas the use of LLB and LSB gave decreased enantioselectivities
(LSB: 49% ee; LLB: 38% ee) and yields. Moreover, attempts to improve the effi-
ciency of this asymmetric catalysis were successful. In the presence of only 5 or
10 mol% catalytic amount of LPB and 1.5 equiv of phosphite, another α-amino
phosphonate (bearing a benzhydryl group as N-substituent) was obtained in ex-
cellent 92 or 96% ee and good yields up to 82% when using 57b as imine compo-
nent (Scheme 21).
The broad generality of this asymmetric hydrophosphonylation method us-
ing catalytic amounts of LPB was shown by the effective conversion of several
types of imines to the corresponding optically active α-amino phosphonates in
satisfactory to excellent enantioselectivities up to 96% ee and yields up to 87%.
Concerning the substitution pattern of the imine component, the authors fo-
cused on the use of alkyl and alkenyl groups as C-substituents, whereas the ben-
zhydryl group was used as N-substituent in most of the cases due to the possibil-
ity to cleave this group from the products in order to obtain the pure α-amino
phosphonic acids with a primary amino function. In case of two imines, the use
of GdPB or PrPB gave the best results. Once again, these results confirm impres-
sively the ability of the LnMB type catalysts as a flexible asymmetric catalytic
system which can be easily modified and “designed” according to the needs for
the desired reaction by changing lanthanoid center ion, alkali metal component
and/or BINOL derivative. However, in this reaction neither the use of second-
generation LPB catalyst nor of LPB derivatives derived from modified BINOLs
gave improved results.
The proposed mechanism of this catalytic asymmetric hydrophosphonyla-
tion is shown in Scheme 22. The first step of this reaction is the deprotonation of
dimethyl phosphite by LPB to generate the potassium dimethyl phosphite. This
potassium phosphite immediately coordinates to a rare earth to give I due to the
strong oxophilicity of rare earth metals [74]. I then reacts with an imine to give
an optically active potassium salt of the α-amino phosphonate, which leads via
proton-exchange reaction to an α-amino phosphonate and LPB, thereby com-
pleting the catalytic cycle and giving the desired asymmetric hydrophospho-
nylation.
O R
R HN
N H P(OCH3)2
(R)-LPB P(OCH3)2
H (5 - 20 mol %)
O
57 58
a: R = CH(p-CH3OC6H4)2 up to 82% yield
b: R = CHPh2 up to 96% ee
* O
O K O O K R1 * P(OCH3)2
HP(OCH3)2 O La O
HNR2
* O O *
K
LPB complex
* *
K
K O O K O O K
O O
O La O * La *
O O
* O O * O K
O H HNR2
KP(OCH3)2 P(OCH3)2
R1*
I III
*
2
N R K
O O K
O O
R1 H * La
O O *
OH
KNR2
P(OCH3)2
R1 *
II
5.2.2
Hydrophosphonylation of Cyclic Imines
O
(H3CO)2P
N (R)-LnPB (5 - 20 mol %) 4 NH
H3C 4 CH3 H 3C CH3
H3C S CH3 (H3CO)2PHO, H3C S CH3
THF/toluene (1:7)
59 (S)-60
up to 97% yield
up to 98% ee
ee 100
(%)
80
60
40
20
Gd
Yb Dy Sm Pr La
0
0.8 0.9 1.0 1.1
100
90
80
Yield (%); ee (%)
70
60
50
40
30
(R)-SmPB (% Yield)
20
(R)-SmPB (% ee)
(R)-YbPB (% Yield)
10 (R)-YbPB (% ee)
0
0 5 10 15 20
Catalytic Amount
(mol %)
Scheme 24. Proposed reaction mechanism (for graphical reasons, the substituents at the
thiazoline 59 are not shown in this scheme).
6
Summary
In conclusion, chiral heterobimetallic lanthanoid compexes LnMB, which were
recently developed by Shibasaki et al., are highly efficient catalysts in stereose-
lective synthesis. This new and innovative type of chiral catalyst contains a Lewis
acid as well as a Brønsted base moiety and shows a similar mechanistic effect as
observed in enzyme chemistry. A broad variety of asymmetric transformations
were carried out using this catalysts, including asymmetric C-C bond forma-
tions like the nitroaldol reaction, direct aldol reaction, Michael addition and
Diels-Alder reaction, as well as C-O bond formations (epoxidation of enones).
Thereupon, asymmetric C-P bond formation can also be realized as has been
successfully shown in case of the asymmetric hydrophosphonylation of alde-
hydes and imines. It is noteworthy that all above-mentioned reactions proceed
with high stereoselectivity, resulting in the formation of the desired optically ac-
tive products in high to excellent optical purity.
7
References and Notes
1. Noyori R (1994) In: Asymmetric catalysis in organic synthesis. Wiley, New York
2. Ojima I (1993) In: Catalytic asymmetric synthesis. VCH, New York
3. Steinhagen H, Helmchen G (1996) Angew Chem 208:2489; Angew Chem Int Ed Engl
35:2337
4. Shibasaki M, Sasai H (1996) Pure & Appl Chem 68:523
5. Shibasaki M, Sasai H, Arai T (1997) Angew Chem 109:1290; Angew Chem Int Ed Engl
36:1236
6. Sasai H, Suzuki T, Itoh N, Tanaka K, Date T, Okamura K, Shibasaki M (1993) J Am
Chem Soc 115:10,372
7. Sasai H, Arai T, Satow Y, Houk KN, Shibasaki M (1995) J Am Chem Soc 117:6194
8. Takaoka E, Yoshikawa N, Yamada YMA, Sasai H, Shibasaki M (1997) Heterocycles
46:157
9. Sasai H, Suzuki T, Itoh N, Shibasaki M (1993) Tetrahedron Lett 34:851
10. Gröger H, Saida Y, Sasai H, Yamaguchi K, Martens J, Shibasaki M (1998) J Am Chem
Soc (in press)
11. Sasai H, Suzuki T, Arai S, Arai T, Shibasaki M (1992) J Am Chem Soc 114:4418
12. Sasai H, Suzuki T, Itoh N, Arai S, Shibasaki M (1993) Tetrahedron Lett 34:2657
13. Although the lithium nitronate is first generated, there also appears to be a significant
possibility that the aldehyde coordinates to La first
14. A 6,6'-disubstituted BINOL derivative was also used in the asymmetric Diels-Alder re-
action by:Terada M, Motoyama Y, Mikami K (1994) Tetrahedron Lett 35:6693
15. Lingfelter DS, Hegelson RC, Cram DJ (1981) J Org Chem 46:393
16. Maruoka K, Itoh T, Araki Y, Shirasaka T, Yamamoto H (1988) Bull Chem Soc Jpn
61:2975
17. Yamamoto K, Noda K, Okamoto Y (1985) J Chem Soc Chem Commun 1065
18. Oshida JI, Okamoto M, Azuma S, Tanaka T (1997) Tetrahedron: Asymmetry 8:2579
19. Sasai H, Itoh N, Suzuki T, Shibasaki M (1993) Tetrahedron Lett 34:855
20. Sasai H, Yamada YMA, Suzuki T, Shibasaki M (1994) Tetrahedron 50:12,313
21. Sasai H, Suzuki T, Itoh N, Shibasaki M (1995) Appl Organomet Chem 9:421
Chiral Heterobimetallic Lanthanoid Complexes 231
60. Kafarski P, Lejczak B (1991) Phosphorus Sulfur Silicon Relat Elem 63:193
61. Powers JC, Boduszek B, Oleksyszyn J (1995) Georgia Tech Research Corp PCT Int Appl
WO 95 29691; Chem Abstr (1996) 124:203,102m
62. Boduszek B, Oleksyszyn J, Kam CM, Selzer J, Smith RE, Powers JC (1994) J Med Chem
37:3969
63. Andrews KJM (1981) Hoffmann-LaRoche. Eur Pat 33,919; Chem Abstr (1982)
96:52,498r
64. Drauz K, Koban HG, Martens J, Schwarze W (1985) Degussa AG. US Pat 4,524,211;
Chem Abstr (1984) 101:211,458x
65. Yokomatsu T, Yamagishi T, Shibuya S (1993) Tetrahedron:Asymmetry 4:1783
66. Rath NP, Spilling CD (1994) Tetrahadron Lett 35:227
67. Sasai H, Bougauchi M, Arai T, Shibasaki M (1997) Tetrahedron Lett 38:2717
68. Yokomatsu T, Yamagishi T, Shibuya S (1997) J Chem Soc, Perkin Trans I 1783
69. The Sr(E) values would be a measure of the susceptibility of the substrate to attack by
the catalysts based on the distribution of electrons in the frontier orbital
70. Fukui K, Yonezawa T, Nagata C (1954) Bull Chem Soc Jpn 27:423
71. Kukhar VP, Soloshonok VA, Solodenko VA (1994) Phosphorus Sulfur Silicon Relat.
Elem 92:239
72. Yager KM, Taylor CM, Smith III AB (1994) J Am Chem Soc 116:9377 and references cit-
ed therein
73. Sasai H, Arai S, Tahara Y, Shibasaki M (1995) J Org Chem 60:6656
74. Molander GA (1992) Chem Rev 92:292
75. Hoppe I, Schöllkopf K, Nieger M, Egert K (1985) Angew Chem 97:1066; Angew Chem
Int Ed Engl 24:1036
76. Gröger H, Martens J (1996) Synth Commun 26:1903
77. Gröger H, Saida Y, Arai S, Martens J, Sasai H, Shibasaki M (1996) Tetrahedron Lett.
37:9291
Reactions of Ketones with Low-Valent Lanthanides:
Isolation and Reactivity of Lanthanide Ketyl and Ketone
Dianion Complexes
Zhaomin Hou* and Yasuo Wakatsuki
The Institute of Physical and Chemical Research (RIKEN), Hirosawa 2–1, Wako,
Saitama 351–0198, JapanYasuo Wakatsuki
*E-mail: [email protected]
Recent progress in the chemistry of structurally well-defined lanthanide ketyl and ketone
dianion complexes is reviewed, with particular emphasis on the ligand effects on the reac-
tivity of these complexes. It has been demonstrated that the stability and reactivity of the
ketyl radical and ketone dianion species strongly depend on the steric and electronic prop-
erties of the ancillary ligands, the structure of their parent ketones, as well as the nature of
the metals to which they are bound. Fine-tuning these factors can control the reactivity of
these species. Generation and reactions of dianionic thioketone and imine species are also
briefly described.
Keywords: Ketyl radicals, Ketone dianions, Lanthanides, pinacol coupling, Ligand effects
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
7 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
1
Introduction
Reactions of ketones with low-valent lanthanides such as SmI2 and the Ln (Ln=
Sm, Yb) metals constitute one of the most useful reactions in lanthanide-medi-
ated organic synthesis, and occupy an important position in modern organic
chemistry [1, 2]. It is well known that one-electron reduction of ketones by low-
valent lanthanides easily produces the corresponding radical anion species, or
ketyls. In the case of diaryl ketones, further one-electron transfer is also achiev-
able to afford the corresponding ketone dianions. These highly reactive species
play a very important role, as key intermediates, in a variety of useful reactions
involving ketones, such as reductions, ketyl-olefin couplings, pinacol couplings,
intermolecular cross couplings, and the Barbier-type reactions. Although other
reducing metals such as alkali, alkaline earth, and low-valent titanium metals
can also produce similar ketyls or ketone dianions upon reaction with ketones
[3], the lanthanide species usually behave differently from those metal analogs
and in many cases show higher reactivity and chemo-, regio- and stereo-selec-
tivity.
Despite the extensive applications and importance of the lanthanide ketyl and
ketone dianion species in organic synthesis, very little was known about their
structures. Our understanding about the reactivity of this important class of
species had relied solely on the analysis of the organic products obtained by hy-
drolysis work-up of the final reaction mixtures. Attempts to isolate these species
were greatly hampered by their extremely high reactivity and air and moisture
sensitivity [4].
It was not until very recently that significant progress has been made in this
field. By using sterically demanding ancillary ligands under an extremely dry
and oxygen-free inert atmosphere, several different types of lanthanide ketyl
and ketone dianion complexes have been successfully isolated and many of them
structurally characterized. Detailed studies on the reactivity of these well-de-
fined complexes have offered unprecedented insights into the mechanistic as-
pects of the reactions of ketones with low-valent lanthanides, and also shed new
lights on similar reactions with other reducing agents. This article is intended to
highlight the recent progress in this area, with particular emphasis on the struc-
ture-reactivity relation of lanthanide ketyl and ketone dianion complexes.
2
One-Electron Reduction of Ketones
The most difficult problem in isolating a ketyl species generated in one-electron
reduction of ketones is due to its rapid hydrogen abstraction and/or coupling re-
actions. To suppress these decomposition reactions, it is essential to use sterical-
ly demanding reducing agents and highly conjugated ketones [5]. It has been
found that several types of lanthanide reducing agents are able to afford isolable
and structurally characterizable ketyl complexes in the reactions with benzo-
Reactions of Ketones with Low-Valent Lanthanides 235
phenone and fluorenone. The structures and reactivity (stability) of the lantha-
nide ketyl complexes are strongly dependent on the ancillary ligands, solvents,
and the structure of their parent ketones.
X-ray analyses have shown that the C-O bond distance of a ketyl unit is gen-
erally around 1.30 Å, which is longer than that of a free ketone (ca. 1.20 Å) and
shorter than that of an alkoxide (ca. 1.40 Å). The radical carbon atom of a ketyl
is still in a sp2-hybrid state.
2.1
Fluorenone Ketyl Complexes with an Aryloxide (ArO) Ligand
Scheme 1
236 Zhaomin Hou, Yasuo Wakatsuki
Scheme 2
OH OH
l
HC
2N nt.
ArO OAr qua 4
Sm THF O2
THF
quant.
O
1a O
2N
H
qu Cl
2a an
t.
H OH
Scheme 3
ly influenced by the ancillary ligands bound to the central metal, and HMPA is a
better ligand than THF for the stabilization of a ketyl species [5, 7, 9].
Reflecting the typical reactivity of a ketyl species, hydrolysis of 2a gives the
corresponding pinacol-coupling product 4, while air oxidation of 2a yields flu-
orenone almost quantitatively (Scheme 3). Reaction of 2a with one equivalent of
1a produced a THF-insoluble purple precipitate which upon hydrolysis afforded
fluorenol quantitatively, suggesting that a fluorenone dianion intermediate was
formed [7, 9].
Reactions of Ketones with Low-Valent Lanthanides 237
2.2
Fluorenone Ketyl Complexes with a C5Me5 Ligand
C5Me5 O
THF hexane
(C5Me5)2Ln(THF)2 + Ln no reaction
Ln = Sm, Yb C5Me5 THF
O
5a: Ln = Sm, 87%
5b: Ln = Yb, 85%
HMPA
C5Me5 O
Ln
C5Me5 HMPA
5c: Ln = Sm, 85%
5d: Ln = Yb, 86%
Scheme 4
Scheme 5
238 Zhaomin Hou, Yasuo Wakatsuki
2.3
Fluorenone Ketyl Complexes with an N(SiMe3)2 Ligand
The ketyl species stabilized by N(SiMe3)2 behaves a little differently from those
by C5Me5 and the ArO ligand. Reaction of Sm(N(SiMe3)2)2(THF)2 with
1 equivalent of fluorenone in THF gave a brown solution whose UV-Vis spec-
trum was almost identical to that of the ArO-ligated ketyl complex 2a, suggest-
ing that the corresponding ketyl species 7a was formed [7]. However, attempts
to isolate the ketyl species 7a from THF were unsuccessful and always resulted
in the formation of the light yellow pinacolate 8a (Scheme 6).
The central C-C bond in 8a could also be cleaved by addition of a strongly co-
ordinative solvent. Dissolving the light yellow 8a in THF gave immediately a
brown solution whose UV-Vis spectrum was the same as that of the solution
originally obtained by the reaction of Sm(N(SiMe3)2)2(THF)2 with fluorenone in
THF (Scheme 6). Addition of 4 equivalents of HMPA to this THF solution afford-
ed the bis(HMPA)-coordinated ketyl complex 7b together with the mono(HM-
PA)-coordinated pinacolate complex 8b (Scheme 7) [7].
The isolation of the pinacolate 8a rather than the ketyl 7a from THF
(Scheme 6) and the isolation of the mono(HMPA)-coordinated pinacolate 8b to-
gether with the bis(HMPA)-coordinated ketyl 7b in the reaction of 8a or 7a with
HMPA (Scheme 7) contrast sharply with the similar reactions of the ArO-ligated
complex 2a or 3, in which only the bis(THF or HMPA)-coordinated ketyl com-
plexes were isolated (see also Sect. 2.1). These differences probably result from
the difference in electron-donating ability between N(SiMe3)2 and the ArO lig-
and. Since N(SiMe3)2 is more electron-donating than the ArO ligand, dissocia-
tion of a neutral L (L=THF or HMPA) ligand from the central Sm atom in the
(Me3Si)2N
THF
(Me3Si)2N N(SiMe3)2 (Me3Si)2N
Sm
THF Sm THF O
Sm(N(SiMe3)2)2(THF)2 + THF – THF
O +THF
O O
Sm
N(SiMe3)2
THF
7a N(SiMe3)2
brown in THF
8a, 79%
light-yellow
Scheme 6
Reactions of Ketones with Low-Valent Lanthanides 239
(Me3Si)2N
HMPA
(Me3Si)2N Sm
(Me3Si)2N N(SiMe3)2
O
HMPA Sm HMPA
HMPA +
O O
Sm
N(SiMe3)2
HMPA
N(SiMe3)2
(Me3Si)2N 7b
8b
THF greenish-brown
(Me3Si)2N light-yellow
Sm 50% 20%
O
ArO OAr
O ArOH Sm
THF THF
Sm – HN(SiMe3)2
N(SiMe3)2 90% O
THF
N(SiMe3)2
8a
light-yellow 2a
purple-brown
ArO OAr
ArOH/HMPA
HMPA Sm HMPA
– HN(SiMe3)2
88% O
2c
dark-green
Scheme 7
(Me3Si)2N
THF
(Me3Si)2N
Sm (Me3Si)2N N(SiMe3)2 (Me3Si)2N N(SiMe3)2
O ∆H =11 kcal/mol Sm + THF Sm
THF THF
2 THF O 2
in toluene – THF O
O
Sm
N(SiMe3)2
THF 7c
N(SiMe3)2 7a
8a
Scheme 8
2.4
Benzophenone Ketyl Complexes
Scheme 9
O TpMe2 Ph
toluene Sm O C
Sm(TpMe2)2 + C
Ph 80% Ph
Ph TpMe2
TpMe2 = BH(3,5-dimethylpyrazolyl)3 10, purple
Scheme 10
Reactions of Ketones with Low-Valent Lanthanides 241
2.5
Complexes Bearing Three Independent Ketyl Ligands
OH
H+
O 4, 98%
HMPA HMPA
3 equiv HMPA
Ln + 3 Ln
THF O
HMPA
O O
+ 0.5 equiv 4
12, 60%
light-yellow
– HMPA
O O
1 HMPA
HMPA
2 HMPA
O Sm
O HMPA HMPA O
Sm O
O H O
13 HMPA
O
H
Scheme 11
Zhaomin Hou, Yasuo Wakatsuki
Reactions of Ketones with Low-Valent Lanthanides 243
trast with those of Na/HMPA [5] or Ca/HMPA [5, 15] with benzophenone, in
which the corresponding benzophenone ketyl species have been successfully
isolated and structurally characterized. These results show that the nature of the
reducing metals also strongly influences on the reactivity and stability of a ketyl
species and the lanthanide ketyl species are more reactive than those of alkali
and alkaline earth metals.
2.6
Structure-Reactivity Relation of Lanthanide Ketyl Complexes
The results described above have demonstrated that the stability and reactivity
of a ketyl species are extremely susceptible to the environment around the cen-
tral metal ion. Both anionic and neutral (solvent) ligands play an important role
in determining the behavior of a ketyl species. The sterically demanding
bis(pentamethylcyclopentadienyl) ligand set (C5Me5)2 together with a neutral L
(L=THF or HMPA) ligand offers a good stabilizing environment for a lanthanide
fluorenone ketyl species, while the less bulky (ArO)2 or ((Me3Si)2N)2 ligand set
requires two L (L=THF or HMPA) ligands for the stabilization of the same ketyl
species. In the latter case, dissociation of one of the two L ligands from the cen-
tral metal ion easily occurs to cause pinacol-coupling of the ketyl, and this takes
place more easily for the (Me3Si)2N-ligated complexes than for the ArO-ligated
ones, due to the stronger electron-donating ability of N(SiMe3)2. Re-coordina-
tion of an L ligand to the metal atom of the resulting pinacolate easily cleaves the
central C-C bond and regenerates the ketyl, which thus makes the pinacol-cou-
pling process reversible. The relatively small dissociation enthalpy for the pina-
colate complexes (e.g., DHdiss=11 kcal/mol for 8a) could well account for this re-
versibility. HMPA, as a neutral and strongly coordinating ligand, offers an excel-
lent stabilizing moiety for both mono- and multi-ketyl complexes. The transfor-
mation of the tris(ketyl) complex 11a to the fluorenoxide/pinacolate complex 12
provides unprecedented insights into the elementary steps of the hydrogenation
and pinacol-coupling reactions of a ketyl species. Ketone-dependence of the sta-
bility of ketyls is also observed. In contrast to fluorenone ketyl complexes, the
isolation of a structurally characterizable benzophenone ketyl complex is more
difficult due to its less conjugated structure which makes it more reactive and
less stable. However, by choosing an appropriate ligand set such as bis(hydrot-
ris(3,5-dimethylpyrazolyl)borate), a structurally characterizable lanthanide
benzophenone ketyl species can indeed be isolated. It is obvious from these re-
sults that the stability and reactivity of a ketyl radical species can be controlled
through tuning the ancillary ligands bound to the central metal ion.
3
Two-Electron Reduction of Ketones
Among all low-valent lanthanides, the zero valent Sm and Yb metals are the most
often used reducing agents for the study of two-electron reduction of ketones.
244 Zhaomin Hou, Yasuo Wakatsuki
Although earlier studies showed that the formation of a samarium ketone dian-
ion species was not as efficient as that of ytterbium [18–20], it was later found
that samarium ketone dianion species could be generated similarly if the metal
surface was sufficiently activated, e.g., by ICH2CH2I [21, 22]. Since the negatively
charged carbon and oxygen atoms in a ketone dianion species are adjacent, the
use of a sterically demanding lanthanide(II) reducing agent does not necessarily
give a good result for the isolation of such species because of the steric repulsion
between the resultant two closely located bulky metal moieties. Therefore, the
control of the reactivity or stability of a ketone dianion species through tuning
the ancillary ligands may not be as effective as in the case of ketyls. Moreover,
compared to the mono-anionic ketyls, ketone dianions are more difficult to gen-
erate. They are more sensitive to air and moisture, and thus more difficult to
handle. Probably due to these reasons, structurally characterized examples of
ketone dianion species remain very scarce [23–25], and only one lanthanide ke-
tone dianion complex, [Yb(m-h1,h2-OCPh2)(HMPA)2]2, has been isolated and
structurally characterized to date (see below) [24]. The known ketone dianions
are limited solely to those bearing two aromatic groups on the carbonyl carbon
atom.
Despite these limitations, it has been found that the lanthanide ketone dian-
ions are an interesting class of species which show unique reactivities. As de-
scribed below, in lanthanide ketone dianion species the carbonyl carbon atom is
completely changed from electrophilic to nucleophilic, and shows strong nucle-
ophilicity towards a variety of organic substrates. Due to the delocalization of
the negative charges in the aromatic groups, some reactions can also take place
at the aromatic ring.
3.1
Isolation of an Ytterbium(II) Benzophenone Dianion Complex
Scheme 12
Reactions of Ketones with Low-Valent Lanthanides 245
O O
C C
...
Scheme 13
Scheme 14
that a benzophenone dianion uses its O atom to make a Yb-O bond with one
Yb(II) ion and its C atom to make a Yb-C bond with the other Yb(II) ion to which
the lone electron pair of the oxygen atom is also donated (Scheme 12). The C-O
bond of the benzophenone dianion (1.39(6) Å) is longer than that of benzophe-
none ketyl (1.31(2) Å) [15, 16] and that of free benzophenone (1.23(1) Å) [26].
Similar to benzophenone ketyl, the carbonyl carbon atom in benzophenone di-
anion is still in an sp2-hybrid state, which can thus allow good conjugation of the
negative charges with the phenyl rings.
Consistent with the above structural observations, the 1H NMR signals for the
phenyl groups in 14 in THF-d8 are greatly upfield shifted to as high as d 5.63–
7.04, demonstrating that the negative charges of the ketone dianion are indeed
highly delocalized into the phenyl rings, especially to the para-positions (d 5.63)
(Scheme 13) [21,22]. As described in Sect. 3.3, this delocalization of electrons
can have great influence on the reactivity of benzophenone dianion species.
3.2
Reactions of Ketone Dianions with Organic Electrophiles
It has been found that lanthanide ketone dianion species are excellent nucle-
ophiles, which react smoothly with a variety of organic substrates such as ke-
246 Zhaomin Hou, Yasuo Wakatsuki
tones, nitriles, epoxides, CO2, etc. to give after hydrolysis the corresponding un-
symmetrical pinacols, a-hydroxy ketones, 1,3-diols, a-hydroxy carboxylic acids,
etc., in good yields, respectively (Scheme 14) [18–20]. Compared to alkali metal
ketone dianions, they are less basic and more nucleophilic, and selectively un-
dergo addition reactions even with substrates bearing active a-protons, such as
aliphatic ketones and nitriles.
3.3
Reactions of Ketone Dianions with 2,6-Di-tert-Butylphenols
Most of the reactions of ketone dianion species, including those of main group
and titanium metals, are known to occur at the carbonyl unit, as demonstrated
in Sect. 3.2. However, it was found that the protonation reactions of lanthanide
benzophenone dianion species with 2,6-di-tert-butylphenols can take place not
only at the carbonyl group but also at the phenyl ring, due to the delocalization
of the negative charges onto the aromatic groups. The selectivity between reac-
tions at the carbonyl group and those at the aromatic ring in lanthanide diaryl
ketone dianion species depends greatly on the nature of both metals and ke-
tones.
3.3.1
Benzophenone Dianion Species
C HMPA OAr
HMPA O ArO HMPA
ArOH
Yb Yb + HMPA Yb HMPA
HMPA Yb
O – Ph2CHOH
HMPA ArO O
C HMPA O
C C
1d, 80% H Ph Ph H
H
14 Ar = C6H2-tBu2-2,6 -Me-4 H
15, 5%
Scheme 15
Reactions of Ketones with Low-Valent Lanthanides 247
OAr
O C HMPA
THF/HMPA HMPA O
Sm + Ph C Ph ArOH HMPA Sm HMPA
Sm Sm
O HMPA 60% O O
HMPA
C C C
H Ph Ph H
H H
16
Ar = C6H2-tBu2-2,6 -Me-4
Scheme 16
OAr
O
THF/HMPA ArOD HMPA Sm HMPA
Sm + Ph C Ph
60% O O
C C
H Ph Ph H
D D
17
Scheme 17
Scheme 18
OAr OAr
toluene
HMPA Sm HMPA HMPA Sm HMPA
180 °C
O O O O H(D)
C C C
C H(D)
H Ph Ph H (D)H
D D
17 H(D)
19
Scheme 19
3.3.2
Fluorenone Dianion Species
The protonation reaction of the Yb(II) fluorenone dianion species with ArOH is
similar to the major reaction of the Yb(II) benzophenone dianion 14 with ArOH,
occurring at both the C and O atoms of the carbonyl unit to afford 1d and fluore-
nol (Scheme 20). However, the protonation reaction of the Sm(II) fluorenone di-
anion species with ArOH takes place only at the C atom and is accompanied by
the oxidation of the Sm(II) ion to give the corresponding samarium(III) ary-
loxide/fluorenoxide complex 20 (Scheme 20), which again demonstrates the
metal dependence of the reactivity of lanthanide ketone dianion species [22].
4
Ketone-Reduction Paths Based on Isolated Intermediates
As described above, with the combination of appropriate lanthanide reducing
agents, solvents, and substrates, almost all types of the key intermediates in-
volved in the reduction of ketones have been isolated and structurally character-
ized (Scheme 21), such as the one-electron reduction product (ketyl radical 21),
its reversible coupling product (pinacolate 24), hydrogen abstraction product
(alkoxide 23) and further one-electron reduction product (ketone dianion 22),
and the selective protonation products of the ketone anion (alkoxide 23 and eno-
(Ln = Yb)
ArO HMPA
ArOH
Yb
– fluorenol
ArO HMPA
O ArO 1d, 85%
Ln(HMPA)x Ln(HMPA)x
THF/HMPA ArOH
Ln + O H
O OAr
(Ln = Sm)
ArOH Sm HMPA
HMPA
– H2
ArO O H
Reactions of Ketones with Low-Valent Lanthanides
20, 64%
Scheme 20
249
250 Zhaomin Hou, Yasuo Wakatsuki
O
Ar C H
Ar
23 H+
solvent-H
O O O OH
+e +e
C C E H+
–e –e C Ar C E
Ar Ar Ar Ar Ar Ar Ar
21 22 E = electrophiles or H+
H+
Ar = Ph H+
OH OH O O O
Ar H+ O
C C Ar Ar C C Ar C
Ph Ph C H
Ar Ar Ar Ar
H Ph
24 H
25
late 25). Structural and reactivity studies of these isolated intermediates have
given rise to an ever clearer picture of the reaction paths in the reaction of ke-
tones with low-valent lanthanides (Scheme 21), and thus offered unprecedented
insights into the mechanistic aspects of these and related reactions.
5
Dianionic Thioketone and Imine Species
Similar to diaryl ketones, aromatic thioketones and imines can also be reduced
to the corresponding dianionic species. Fujiwara and coworkers have reported
that the ytterbium diaryl thioketone dianion species, generated by reaction of Yb
metal with the thioketones in THF/HMPA at low temperature, show good nucle-
ophilicity towards organic substrates such as acetone and alkyl halides
(Scheme 22) [31]. The thioketone dianion species seemed to be less stable than
those of ketones. At room or higher temperatures, C-S bond cleavage reaction
took place.
Reaction of N-(diphenylmethylidene)aniline with 1 equivalent of Yb metal in
THF/HMPA gives the corresponding dianionic complex Yb(h2-Ph2CNPh)(HM-
PA)3 (26) whose structure has been crystallographically determined (Scheme 23)
[32]. The imine dianion complex 26 is more basic and less nucleophilic than ke-
tone- and thioketone-dianion species. It reacts with CO2 to give after hydrolysis
the corresponding Yb(III) amino acid derivative [33], but with acetone to release
Reactions of Ketones with Low-Valent Lanthanides 251
SH OH
CH3COCH3
Ar C C CH3
72%
Ar CH3
S S
THF/HMPA Yb
C + Yb Ar C
Ar Ar 0 °C
Ar
50-66% Ar ( )n
S
n = 1, 2
Scheme 22
CH3COCH3
Ph2CHNHPh
> 99%
Ph Ph
N
THF/HMPA N
+ Yb Yb(HMPA)3
C room temp. C
Ph Ph Ph
Ph
CO2 H 2O
26, red-black Ph2C(NHPh)COO)3Yb • 2HMPA •
45% 2H2O
92%
Scheme 23
6
Conclusions and Perspectives
The results described in this article have clearly demonstrated that the highly re-
active lanthanide ketyl and ketone dianion species can not only be isolated but
also their reactivity be controlled through adjusting the environment around the
central metals. For example, a ketyl species can be “deactivated” (stabilized) by
binding to a bulky metal moiety and “reactivated” by reducing the bulkiness
around the metal. Although the examples presented here are limited to those of
aromatic ketones, these data should be conceptually useful for the understand-
ing and design of reactions involving other organic carbonyl compounds. In
view of the versatile uses of ketyls in organic synthesis and the high nucle-
ophilicity of lanthanide ketone dianion species, if the “deactivation” and “reacti-
vation” of alkyl ketone and/or aldehyde ketyl species can be achieved in a con-
trollable way, and if alkyl ketone and/or aldehyde dianion species can also be
generated, numerous classical C-C bond formation processes will be substan-
252 Zhaomin Hou, Yasuo Wakatsuki
tially improved and our capability to access a variety of new target compounds
will be dramatically enhanced. These topics remain to be challenged.
7
References
1. For some reviews on lanthanide-mediated organic synthesis, see (a) Molander GA,
Harris CR (1996) Chem Rev 96:307; (b) Molander GA (1992) Chem Rev 92:26; (c)
Molander GA (1991) Samarium and ytterbium reagents. In: Trost BM, Fleming I, Sch-
reiber S (eds) Comprehensive organic synthesis, vol 1. Pergamon, Oxford, p251; (d)
Imamoto T (1991) Organocerium reagents. In: Trost BM, Fleming I, Schreiber S (eds)
Comprehensive organic synthesis, vol 1. Pergamon, Oxford, p 231; (e) Kagan HB,
Namy JL (1986) Tetrahedron 42:6573
2. Imamoto T (1994) Lanthanides in organic synthesis. Academic Press, London
3. For examples of the formation and reactions of other metal ketyl and ketone dianion
speices, see: (a) Wirth T (1996) Angew Chem Int Ed Engl 35:61; (b) Fürstner A,
Bogdanovic B (1996) Angew Chem Int Ed Engl 35:2442; (c) Huffman JW (1991) Reduc-
tion of C=X to CHXH by dissolving metals and related methods. In: Trost BM, Fleming
I (eds) Comprehensive organic synthesis, vol 8. Pergamon, Oxford, p107; (d) Robert-
son GM (1991) Pinacol coupling reactions. In: Trost BM, Fleming I, Pattenden G (eds)
Comprehensive organic synthesis, vol 3. Pergamon, Oxford, p563; (e) McMurry JE
(1989) Chem Rev 89:1513; (f) Kahn BE, Riecke RT (1988) Chem Rev 88:733; (g) Prad-
han SK (1986) Tetrahedron 42:6351
4. For a spectroscopic study of in situ generated lanthanide ketyl species, see Hirota N,
Weissman SI (1964) J Am Chem Soc 86:2538
5. For a recent overview on metal ketyl complexes, see Hou Z, Wakatsuki Y (1997) Chem
Eur J 3:105
6. Hou Z, Miyano T, Yamazaki H, Wakatsuki Y (1995) J Am Chem Soc 117:4421
7. Hou Z, Fujita A, Zhang Y, Miyano T, Yamazaki H, Wakatsuki Y (1998) J Am Chem Soc
120:754
8. (a) Hou Z, Zhang Y, Wakatsuki Y (1997) Bull Chem Soc Jpn 70:149; (b) Hou Z, Wakat-
suki Y (1994) J Chem Soc Chem Commun 1205
9. Hou Z, Wakatsuki Y (1995) J Synth Org Chem Jpn 53:906
10. Hou Z, Zhang Y, Yoshimura T, Wakatsuki Y (1997) Organometallics 16:2963
11. (a) Neumann WP, Schroeder B, Ziebarth M (1975) Liebigs Ann Chem 2279; (b) Ziebar-
th M, Newmann WP (1978) Liebigs Ann Chem 1765
12. Neumann WP, Uzick W, Zarkadis AK (1986) J Am Chem Soc 108:3762
13. Covert KJ, Wolczanski PT, Hill SA, Krusic PJ (1992) Inorg Chem 31:66
14. Nolan SP, Stern D, Marks TJ (1989) J Am Chem Soc 111:7844
15. Hou Z, Jia X, Wakatsuki Y (1997) Angew Chem Int Ed Engl 36:1292
16. (a) Takats J (private communication). See also Takats J (1997) J Alloys Compd 249:51;
(b) Clegg W, Eaborn C, Izod K, O’Shaughnessy P, Smith JD (1997) Angew Chem Int Ed
Engl 36:2815.
17. Hou Z, Fujita A, Yamazaki H, Wakatsuki Y (1996) J Am Chem Soc 118:7843
18. Hou Z, Takamine K, Fujiwara Y, Taniguchi H (1987) Chem Lett 2601
Reactions of Ketones with Low-Valent Lanthanides 253
19. Hou Z, Takamine K, Aoki O, Shiraishi H, Fujiwara Y, Taniguchi H (1988) J Chem Soc
Chem Commun 668
20. Hou Z, Takamine K, Aoki O, Shiraishi H, Fujiwara Y, Taniguchi H (1988) J Org Chem
53:6077
21. Hou Z, Yamazaki H, Fujiwara Y, Taniguchi H (1992) Organometallics 11:2711
22. Yoshimura T, Hou Z, Wakatsuki Y (1995) Organometallics 14:5382
23. Bogdanovic B, Kruger C, Wermeckes B (1980) Angew Chem Int Ed Engl 19:817
24. Hou Z, Yamazaki H, Kobayashi K, Fujiwara Y, Taniguchi H (1992) J Chem Soc Chem
Commun 722
25. Hou Z, Fujita A, Yamazaki H, Wakatsuki Y (1998) Chem Commun 669
26. Fleischer EB, Sung N, Hawkinson S (1968) J Phy Chem 72:4311
27. Hou Z, Yoshimura T, Wakatsuki Y (1994) J Am Chem Soc 116:11,169
28. cf. Eo(Sm3+/Sm2+)=–1.55 V, Eo(Yb3+/Yb2+)=–1.15 V in aqueous medium
29. Sm(II) is ca. 0.14 Å bigger than Yb(II) in radius when both have the same coordination
number. See Shannon RD (1976) Acta Crystallogr Sect A 32:751
30. For examples of Birch reductions, see: Mander LN (1991) Partial reduction of aromatic
rings by dissolving metals and by other methods. In: Trost BM, Fleming I (eds) Com-
prehensive organic synthesis, vol 8. Pergamon, Oxford, p489 and references cited
therein
31. (a) Makioka Y, Uebori S. Tsuno M, Taniguchi Y, Takaki K, Fujiwara Y (1996) J Org
Chem 61:372; (b) Makioka Y, Uebori S. Tsuno M, Taniguchi Y, Takaki K, Fujiwara Y
(1994) Chem Lett 611
32. Makioka Y, Taniguchi Y, Fujiwara Y, Takaki K, Hou Z, Wakatsuki Y (1997) Organome-
tallics 15:5476
33. Takaki K, Tanaka S, Fujiwara Y (1991) Chem Lett 493
34. Makioka Y, Saiki A, Takaki K, Taniguchi Y, Kitamura T, Fujiwara Y (1997) Chem Lett 27
Organo Rare Earth Metal Catalysis for the Living
Polymerizations of Polar and Nonpolar Monomers
Hajime Yasuda
This article deals with the rare earth metal initiated polymerization of polar and nonpolar
monomers in a living fashion. For example, [SmH(C5Me5)2]2 or LnMe(C5Me5)2(THF) (Ln=
Sm, Y and Lu) conducted the polymerization of methyl methacrylate(MMA) to give high
molecular weight syndiotactic polymers (Mn>500,000, syndiotacticity>95%) quantitative-
ly at low temperature (–95 °C). The initiation mechanism was discussed on the basis of X-
ray analysis of the 1:2 adduct of [SmH(C5Me5)2]2 with MMA. Synthesis of high molecular
weight isotactic poly(MMA) with very narrow molecular weight distribution was for the
first time realized by the efficient catalytic function of Yb[C(SiMe3)3]2. Living polymeriza-
tions of alkyl acrylates (methyl acrylate, ethyl acrylate, and butyl acrylate) were also possi-
ble by the excellent catalysis of LnMe(C5Me5)2(THF) (Ln=Sm, Y). By taking advantages of
the living polymerization ability, we attempted ABA triblock copolymerization of
MMA/butyl acrylate/MMA to obtain rubber-like elastic polymers. Organo rare earth metal
complexes such as LnOR(C5R5)2 or LnR(C5R5)2 conducted the living polymerizations of
various lactones such as β-propiolactone, δ-valerolactone and ε-caprolactone, and also
conducted the block copolymerizations of MMA with various lactones. Lanthanum alkox-
ide(III) has good catalytic activity for the polymerization of alkyl isocyanates. Monodis-
perse polymerizations of lactide and various oxiranes were also achieved by the use of rare
earth metal complexes. C1 symmetric bulky organolanthanide(III) complexes such as
SiMe2[2(3),4-(SiMe)2C5H2]2LnCH(SiMe3)2 (Ln=La, Sm, and Y) show high catalytic activity
towards linear polymerization of ethylene. Organolanthanide(II) complexes such as ra-
cemic SiMe2(2-SiMe3-4-tBu-C5H2)2Sm(THF)2 as well as C1 symmetric SiMe2[2(3),4-
(SiMe3)2C5H2]2Sm(THF)2 were found to have high activity for the polymerization of ethyl-
ene to give Mn>106 with Mw/Mn =1.6. Utilizing the high polymerization activity of rare
earth metal complexes towards both polar and nonpolar monomers, block copolymeriza-
tions of ethylene with polar monomers such as methyl methacrylate and lactones were for
the first time realized. 1,4-cis-Conjugated diene polymerization of 1,3-butadiene and iso-
prene became available by the efficient catalytic activity of NdCl(C5H5)2/AlR3 or Nd(oc-
tanoate)3/AlR3. The Ln(naphthenate)3/AliBu3 system allows selective polymerization of
acetylene in cis-fashion.
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
2 Highly Stereospecific Living Polymerization of
Alkyl Methacrylates . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
3 Living Polymerization of Alkyl Acrylates . . . . . . . . . . . . . . . 262
256 Hajime Yasuda
1
Introduction
Recent development of various living polymerizations of polar and nonpolar
monomers allows the synthesis of high molecular weight polymers with very
narrow molecular weight distribution. The indispensable conditions for the ide-
al living polymers are (1) number-average molecular weight Mn>100,000, (2)
polydispersity index Mw/Mn<1.05, (3)stereoregularity >95%, and (4) high con-
version in a short period of time >95%. Even the famous living polystyrene sys-
tem [1] and the Group Transfer system [2] do not simultaneously satisfy require-
ments (1) and (2). Recently, to satisfy a long desire of many polymer chemists,
all these four requirements were found to be met in the rare earth metal-initiated
polymerization of methyl methacrylate (MMA), which gave Mn>500,000 with
Mw/Mn=1.02–1.03 (syndiotacticity >95%) [3]. More recently, we have found iso-
tactic polymerization of MMA by the efficient catalysis of non-metallocene type
single component complex, Yb[C(SiMe3)3]2, to give the poly(MMA) of
Mn>200,000 with Mw/Mn=1.1 in high yield (isotacticity >97%).
Alkyl acrylates were for the first time polymerized in a living fashion with the
aid of the unique catalytic action of rare earth metal complexes [4]. Since these
monomers have an acidic α-H, termination and chain transfer reactions occur
so frequently that their polymerizations generally do not proceed in a living
manner. By taking advantages of the living polymerization ability of both MMA
and alkyl acrylate, ABA or ABC type tri-block copolymerization was performed
to obtain thermoplastic elastomers.
Living polymerization of lactones has been successfully performed by the ca-
talysis of rare earth metal complexes producing Mw/Mn values of 1.07–1.08 [5].
Polymerizations of acrylonitrile and alkyl isocyanates have been successfully re-
alized using La[CH(SiMe3)2]2(C5Me5) as initiator, and those of various oxiranes
have been made using Ln(acac)3/AlR3/H2O system [6].
Organo rare earth metal initiators also show good activity towards non-polar
monomers such as ethylene, 1-olefins, styrene, conjugated dienes, and acetylene
Organo Rare Earth Metal Catalysis for the Living Polymerizations of Polar and Nonpolar Monomers 257
2
Highly Stereospecific Living Polymerization of Alkyl Methacrylates
Although various living polymerization systems have been proposed, any of an-
ionic [7], cationic [8], Group Transfer [2], and metal carbene initiated polymer-
izations [9] achieved no success in synthesizing living polymers of Mn>500,000
with Mw/Mn<1.05. On the other hand, high molecular weight poly(MMA) hav-
ing an unusually low polydispersity has been synthesized by utilizing the unique
initiating function of organolanthanide(III) complexes (Fig. 1) [3]. The relevant
complexes include lanthanide hydrides, bulky alkyl lanthanide, bis-Me-bridged
trimethylaluminum complexes of alkyllanthanides, and simple alkyl complexes
(Fig. 2). The results of the polymerization of MMA with [SmH(C5Me5)2]2 ini-
tiator at different temperatures are summarized in Table 1. The most striking is
very narrow molecular weight distributions, i.e. Mw/Mn=1.02–1.04 for
Mn>60×103. Remarkably, [SmH(C5Me5)2]2 complexes give high conversion (pol-
ymer yield) in a relatively short period, and allow the polymerization to proceed
over a wide range of reaction temperatures from –78 to 40 °C. Furthermore, syn-
diotacticity exceeding 95% is obtained when the polymerization temperature is
lowered to –95 °C.
The typical initiator systems reported so far for the synthesis of highly
syndiotactic poly(MMA) are bulky alkyllithium CH3(CH2)4CPh2Li [10], Grig-
nard reagent in THF [11], and some AlR3 complexes [12]. Although the
On the other hand, monodisperse isotactic poly(MMA) has been obtained with
tBuMgBr in toluene at –40 °C (Mn=10.1×103, Mw/Mn=1.08, isotacticity =96.7%)
[16] or secBuMgBr in toluene at –78 °C (Mn=4.9×103, Mw/Mn=1.29, isotacticity
=95.5%) [16]. Synthesis of isotactic poly(MMA) with much higher molecular
weight is required. Recently, high molecular weight isotactic poly(MMA) (mm=
97%, Mn=500,000, Mw/Mn=1.12) was first obtained quantitatively by the use of a
non-metallocene type rare earth complex [(Me3Si)3C]2Yb (Fig. 4) [17]. We pro-
pose here the initiation mechanism as noted below through noncyclic interme-
diate for this isotactic polymerization of MMA (Fig. 5). The polymerization
should proceed by the enantiomorphic site control in toluene. In contrast, syn-
diotactic polymer was obtained when THF was used as solvent (rr=87%, Mn=
3.2×104, Mw/Mn=1.76) presumably due to chain end control through cyclic in-
termediate. By utilizing the resulting isotactic polymer, we have prepared a ster-
eo-complex between isotactic and syndiotactic polymers by mixing these two
polymers in acetone in a 1:2 ratio (Fig. 6). The resulting stereo-complex shows
the intermediate physical property between the homo-isotactic and homo-syn-
dotactic polymers (Table 2). Isotactic polymerization of MMA has also been
achieved (mm=94%, Mn=134×103, Mw/Mn=6.7) by using Me2Si(C5Me4)[C5H3-
(1S),(2S),(5R)-neomenthyl]LaR (R=CH(SiMe3)2 or N(SiMe3)2) [18], while
Me2Si(C5Me4)[C5H3-(1S),(2S),(5R)-menthyl]LnR (Ln=Lu, Sm; R=CH(SiMe3)2
or N(SiMe3)2) complex produces syndiotactic poly(MMA) (rr=69%, Mn=177×
260103, Mw/Mn=15.7). Polydispersity is rather wide in these cases.
Boffa and Novak found a divalent rare earth metal complex, Sm(C5Me5)2, to
be a good catalyst for polymerization of MMA [19]. The initiation started with
the coupling of two coordinated MMA molecules to form Sm(III) species. The
bis-allyl initiator, Sm(η3-CH2-CH-CH)2(C5Me5)2, was also effective for living
polymerization of MMA. In this case, MMA must add to both ends of the hexa-
diene group. In the polymerization of MMA initiated with methylaluminum tet-
raphenylporphyrine, crowded Lewis acid such as MeAl(ortho-substituted phe-
nolate)2 serves as a very effective accelerator without damaging the living char-
acter of polymerization [20]. Thus, the polymer produced has narrow polydis-
persity (Mw/Mn=1.09) and sufficiently high molecular weight, Mn=25,500, but
the stereoregularity is very poor. The use of a simple organoaluminum such as
trimethylaluminum causes the occurrence of preferred termination.
The organolanthanide initiators also allowed stereospecific polymerization
of ethyl, isopropyl, and tert-butyl methacrylates. The rate of polymerization and
syndiotacticity decreased with increasing bulkiness of the alkyl group in the or-
der Me>Et>iPr>>tBu. Butyl methacrylate was also polymerized using Nd(oc-
tanoate)3/AliBu3 (Al/Nd=7–10), but the molecular weight distribution and ster-
eoregularity were not reported [21].
In general, Ziegler-Natta catalysts such as TiCl4/MgCl2/AlR3 and Kamin-
sky catalysts such as Cp2ZrCl2/(AlMe2-O-)n do not catalyze the poly-
merization of polar monomers. However, a mixture of cationic species
Cp2ZrMe(THF)+ and Cp2ZrMe2 has been found to do so for MMA [17], al-
lowing syndiotactic poly(MMA) (rr=80%, Mn=120,000, Mw/Mn=1.2–1.3)
[22]. Recently, Soga et al. [23] reported the syndio rich polymerization of
MMA catalyzed by Cp2ZrMe2/Ph3CB(C6F5)4/ZnEt2 and also the isotactic po-
262 Hajime Yasuda
3
Living Polymerization of Alkyl Acrylates
Living polymerization of alkyl acrylates is usually very difficult because the
chain transfer or temination occurs preferentially, owing to a high sensitivity of
the acidic α-proton to the nucleophilic attack. Exceptions are the living polym-
erization of a bulky acrylic ester catalyzed by alkyllithium/inorganic salt (LiCl)
systems as well as the Group Transfer polymerizations of ethyl acrylate using
ZnI2 as the catalyst. Porphyrinatoaluminum initiator systems also induced the
living polymerization of tert-butyl acrylate [24], but the upper limit of molecu-
lar weight attained was ca. 20, 000. We have found the efficient initiating proper-
ties of SmMe(C5Me5)2(THF) and YMe(C5Me5)2(THF) for living polymerization
of acrylic esters (Table 3), i.e. methyl acrylate (MeA), ethyl acrylate (EtA), butyl
acrylate (Bu), and tert-butyl acrylate (tBuA), although the reactions were non-
stereospecific (Fig. 7) [25]. The initiator efficiency exceeded 90% except for
tBuA. We therefore concluded that the reactions occur in living fashion. In fact,
the Mn of poly(BuA) initiated by SmMe(C5Me5)2(THF) increased linearly in pro-
4
Block Copolymerization of Hydrophobic and Hydrophilic Acrylates
Trimethylsilyl metacrylate (TMSMA) was found to proceed the living polymer-
ization using SmMe(C5Me5)2(THF) as initiator to give high molecular weight
syndiotactic polymer (Mn=56,000, rr=92%) with very low polydispersity
(Mw/Mn=1.09). By utilizing this nature, we have performed the block copolym-
erization of TMSMA with MMA or butyl acrylate to obtain the adhesive materi-
als upon hydrolysis of the resulting polymer [26]. The result is shown in Table 5.
Thus, block copolymerization of TMSMA with MMA (or BuA) gave thermally
stable adhesive materials for the first time (high thermal stability originates
from high syndiotacticity) (Fig. 9).
5
Polymerization of Alkyl Isocyanates
Polyisocyanates have attracted much attention owing to their liquid crystalline
properties, stiff-chain solution characteristics, and induced optical activities as-
sociated with the helical chain conformation. Pattern and Novak discovered that
such titanium complexes as TiCl3(OCH2CF3) and TiCl2(C5H5)(OCH2CF3) initi-
ate the living polymerization of isocyanates at ambient temperature, giving pol-
ymers with narrow molecular weight distribution [27]. When hexyl isocyanate
was added to TiCl3(OCH2CF3) the polymerization took place, to give Mn (Mw/Mn=
1.1–1.3) increasing linearly with the initial monomer-to-initiator mole ratio or
the monomer conversion over a wide range. Recently, Fukuwatari et al. found
lanthanum isopropoxide to serve as a novel anionic initiator for the polymeriza-
tion of hexyl isocyanate at low temperature (–78 °C), which led to very high mo-
lecular weight (Mn>106) and rather narrow molecular weight distribution
(Mw/Mn=2.08–3.16) [28]. Other lanthanide alkoxides such as Sm(OiPr)3,
Yb(OiPr)3, and Y(OiPr)3 also induced the polymerization of hexyl isocyanate.
Furthermore, it was shown that butyl, isobutyl, octyl and m-tolyl isocyanates
were polymerized using lanthanum isopropoxide as initiator. However, tert-
butyl and cyclohexyl isocyanates failed to polymerize with these initiators under
the same reaction conditions. When the reaction temperature was raised to am-
bient temperature, only cyclic trimers were produced at high yields. More re-
cently we have found that La(C5Me5)2[CH(SiMe3)]2 also initiates the polymeri-
zation of hexyl isocyanate in 90–96% yields in THF at –20 to –40 °C (Mn=59×104,
Mw/Mn=1.57–1.90) [29] (Fig. 10).
6
Living Polymerization of Lactones
AlEt3-H2O or AlEt3-catalyzed polymerizations of 3-methyl-β-propiolactone and
ε-caprolactone have been reported [30, 31] but this type of polymerization gen-
erally gives a broad molecular weight distribution. We have explored the polym-
erization of various lactones including β−propiolactone (PL), 3-methyl-β-propi-
olactone (MePL), δ-valerolactone (VL) and ε-caprolactone initiated by single
organolanthanides, and found that VL and CL led to the living polymerization
(Fig. 11) , yielding polymers with Mw/Mn=1.05–1.10 at quantitative yields
(Table 6). For ε-caprolactone, Mn obtained with the SmMe(C5Me5)2(THF) or
7
Stereospecific Polymerization of Oxiranes
Organolanthanide(III) complexes such as LnMe(C5Me5)2 or LnH(C5Me5)2 do
not initiate the polymerization of oxirans, but more complex systems like
Ln(acac)3/AlR3/H2O and Ln(2-EP)3/AlR3/H2O (2-EP; di-2-ethylhexylphos-
phate) polymerize with good initiation activity [33]. High molecular weight
poly(ethylene oxide) is one of the common water-soluble polymers useful as
adhesives, surfactants, plasticizers, and dispersants as well as for sizing materi-
al. Poly(ethylene oxide) of Mn=2.85×106 was obtained with the Y(2-
EP)3/AliBu3/H2O system at the ratio of Y/Al/H2O=1/6/3. The initiator activity
varies with the molar ratio of the components. Polymerization of propylene ox-
ide was reported to proceed with the Ln(acac)3/AlEt3/H2O system, and light
rare earth elements (Y, La, Pr, Nd, Sm) produced very high molecular weight
poly(propylene oxide) at Al/Ln=6 in a short period of time (2 h) in toluene [34].
The Nd(acac-F3)3/AlR3/H2O (Ln=Y, Nd) systems [35] gave isotactic poly(pro-
pylene oxide), while the Cp2LnCl/AlR3/H2O, Sm(OiPr)3/AlR3/H2O or Y(2-ethyl-
hexanoate)3/AlR3/H2O system produced relatively low molecular weight isotac-
tic species of this polymer (Fig. 14).
268 Hajime Yasuda
and the ratio of these ring openings changes little with the initiator system and
the polymerization temperature. The polymers obtained were amorphous ac-
cording to DSC and XRD analyses.
8
Polymerization of Lactide
Shen et al. [44] succeeded ring opening polymerization of D,L-lactide (racemic
species) using Nd(naphthenate)3/AliBu3/H2O (1:5:2.5), Nd(P204)3/AliBu3/H2O
(P204=[CH3(CH2)3CH(CH2CH3)CH2O]2P(O)OH], and Nd(P507)3/AliBu3 /H2O
(P507=(iC8H17O)2P(O)OH) systems, obtaining the polymers whose molecular
weights were Mn=3.1–3.6×104 and the conversions larger than 94%. When the
Ln(naphthenate)3/AliBu3/H2O system was used, nearly the same results were
obtained irrespective of the metals used (La, Pr, Sm, Gd, Ho, Tm). Divalent (2,6-
tBu2-4-Me-phenyl)2Sm(THF)4 was also found to be active for the polymeriza-
tion of D,L-lactide at 80 °C in toluene, giving Mn of 1.5–3.5×104 [45]. A more re-
cent finding is that the Ln(O-2,6-tBu2-C6H3)3/iPrOH (1:1–1:3) system initiates a
smooth homo-polymerization of L-lactide, CL (caprolactone) and VL (valerol-
actone) to give relatively high molecular weights (Mn>24×103) with low polydis-
persity indices (Mw/Mn=1.2–1.3) (Fig. 16) [46]. Ring-opening polymerization of
D,L-lactide was also carried out by using Ln(OiPr3)3 as the catalyst at 90 °C in tol-
uene. The catalytic activity increased in the order La>Nd>Dy>Y and the molec-
ular weight reaches 4.27×104(conversion 80%) [47]. However, the molecular
weight distribution is not clear at present. The block copolymerization of CL
with L-lactide proceeded effectively and gave a polymer of very narrow molecu-
lar weight distribution (Mw/Mn=1.16). On the other hand the addition of CL to
the living poly(L-lactide) end led to no success.
9
Stereospecific Polymerization of Olefins
Bulky organolanthanide(III) complexes such as LnH(C5Me5)2 (Ln=La, Nd) were
found to catalyze with high efficiency the polymerization of ethylene. These hy-
drides are, however, thermally unstable and cannot be isolated as crystals.
Therefore, thermally more stable bulky organolanthanides were synthesized by
Fig. 19. Initiation mechanism for polymerization of ethylene catalyzed by rare earth metal
(III) complexes
Fig. 20. X-ray and NMR analysis of racemic, meso and C1 symmetric complexes(II)
Fig. 21. Initiation mechanism for polymerization of ethylene catalyzed by rare earth meta
(II) complexes
highest activity for the polymerization of ethylene, but the molecular weights of
the resulting polymers are the lowest. On the other hand, the racemic and C1
symmetric complexes produce much higher molecular weight polyethylene but
the activity is rather low.
Particularly, the very high molecular weight polyethylene (Mn>100×104) ob-
tained with C1 complex deserves attention (Fig. 21). For the polymerization of
α-olefins, only the racemic divalent complex showed good activity at 0 °C in tol-
uene: poly(1-hexene) Mn=24,600, Mw/Mn=1.85; poly(1-pentene) Mn=8,700,
Mw/Mn=1.58. Thus, we see that the reactivity of divalent organolanthanide com-
plexes varies depending on their structure. The poly(1-alkene) obtained re-
vealed highly isotactic structure (>95%) when examined by 13C NMR (Fig. 22).
The dihedral angles of Cp'-Ln-Cp' of racemic and meso type divalent complexes
were 117 and 116.7°, respectively. These values are much smaller than those of
Organo Rare Earth Metal Catalysis for the Living Polymerizations of Polar and Nonpolar Monomers 273
10
Block Copolymerization of Ethylene with Polar Monomers
Block copolymerization of ethylene or propylene with polar monomers is yet to
be attained in polyolefin engineering. The success of this type of block copoly-
merization should give hydrophobic polymeric materials having remarkably
high adhesive, dyeing, and moisture adsorbing properties. The following is the
first example of a well-controlled block copolymerization using the unique dual
catalytic function of LnR(C5Me5)2 (Ln=Sm, Yb, and Lu; R=H, Me) complexes
toward polar and nonpolar olefins [55]. Ethylene was copolymerized with
MMA first by the homopolymerization of ethylene (17–20 mmol) with
SmMe(C5Me5)2(THF) or [SmH(C5Me5)2]2(0.05 mmol) at 20 °C in toluene un-
der atmospheric pressure, and then sequential addition of MMA (10 mmol)
(Table 9). The initial step proceeded very rapidly, completed in 2 min, and gave
a polymer of Mn=ca. 10,100 and Mw/Mn=1.42–1.44. However, the second step
was rather slow, with the reaction taking 2 h at 20 °C (Fig. 24). The polymer ob-
tained was soluble in 1,2-dichlorobenzene and 1,2,4-trichlorobenzene at 100 °C
but insoluble in THF and CHCl3. This fact indicates quantitative conversion to
the desired linear block copolymer. Repeated fractionation in hot THF did not
change the molar ratio of the polyethylene and poly(MMA) blocks, though po-
ly(MMA) blended with polyethylene can easily be extracted with THF. With the
copolymerization, the elution maximum in GPC shifted to a higher molecular
Organo Rare Earth Metal Catalysis for the Living Polymerizations of Polar and Nonpolar Monomers 275
weight region, with its initial unimodal pattern unchanged. The relative molar
ratio of the polyethylene and poly(MMA) blocks was controllable at will in the
range of 100:1 to 100:103 if the Mn of the initial polyethylene was fixed to ca.
10,300. 1H and 13C NMR spectra for the copolymers as well as their IR absorption
spectra were superimposable onto those of the physical mixtures of the respective
homopolymers. The molar ratio of the poly(MMA) and polyethylene blocks,
however, decreased as the Mn of the prepolymer increased, especially when Mn
exceeds over ca. 12,000 at which value polyethylene began precipitating as fine
colorless particles. It is noteworthy that smooth block copolymerization of ethyl
acrylate or methyl acrylate to the polyethylene growing chain (Mn=6600–24,800)
can be realized by the sequential addition of the two monomers.
We have extended the above work to the block copolymerization of ethylene
with lactones. δ-Valerolactone and ε-caprolactone were incorporated to the
growing polyethylene end at ambient temperature and the expected AB type co-
polymers (100:1 to 100:89) were obtained in high yield.
The treatment of the resulting block copoly(ethylene/MMA) (100:3, Mn=
35,000) and block copoly(ethylene/ε−caprolactone) (100:11, Mn=12,000) with
dispersed dyes (Dianix AC-E) made them deeply dye with three primary colors,
though polyethylene itself was inert to these dyes. Reversed addition of the mon-
omers (first MMA or lactones and then ethylene) induced no block copolymer-
ization at all, even in the presence of excess ethylene, and only homo-po-
ly(MMA) and homo-poly(lactone) were produced. Hence, these copolymers can
be said to have a very desirable chemical reactivity.
More recently, Yang et al. [56] have examined a new approach in which a reac-
tive functional group was introduced into polyolefins using methylenecyclopro-
pane. Thus, ethylene (1.0 atm) was copolymerized with methylenecyclopropane
(0.25–2.5 ml) using [LnH(C5Me5)2]2 (Ln=Sm, Lu) in toluene at 25 °C, and it was
shown that 10–65 of exo-methylenes were incorporated per 1000 -CH2- units.
The resulting polymer had a Mw of 66–92×103. Yet its Mw/Mn was larger than 4.
276 Hajime Yasuda
11
Polymerization of Styrene
Styrene polymerization was performed by using binary initiator systems such as
Nd(acac)3/AlR3 or Nd(P507)3/AlR3. Syndio-rich polystyrene was obtained at a ra-
tio of Al/Nd=10–12 [57]. More recently, it was shown that the
Gd(OCOR)3/iBu3Al/Et2AlCl catalytic system initiates the copolymerization of
styrene with butadiene, but gives only atactic polystyrene [58]. The
Sm(OiPr)3/AlR3 or Sm(OiPr)3/AlR2Cl (Sm/Al=1–15) catalytic system also initi-
ates the polymerization of styrene to give a high molecular weight polymer (Mn=
300,000), low in polydispersity but atactic in stereoregularity [59]. The cationic
polymerization of styrene using Ln(CH3CN)9(AlCl4)3(CH3CN) was also exam-
ined [60], with the finding that the activity increased in the order La (conversion
73%)>Tb=Ho>Pr=Gd>Nd>Sm=Yb>Eu (conversion 54%), while Mn decreased
with increasing the polymerization temperature from 0 (20×103) to 60 °C
(13×103). A more recent study showed that the single component initiator
[(tBuCp)2LnCH3]2 (Ln=Pr, Nd, Gd) initiated the polymerization of styrene at rel-
atively high temperature, 70 °C, with a conversion of 96% and the Mn of 3.3×104
for [(tBuCp)2NdCH3]2 [61], though stereoregularity was very poor. The activity
varied greatly with the lanthanide element; and catalytic activity increased in the
order Nd>Pr>Gd>>Sm=Y (the Sm and Y complexes showed practically no activ-
ity). Therefore, the reaction is supposed to follow the radical initiation me-
chanism. Styrene polymerization was also performed successfully using the
single component initiators, [(Me3Si)2N]2Sm(THF)2, [(Me3Si)2CH]3Sm, and
La(C5Me5)[CH(SiMe3)2]2(THF) at 50 °C in toluene without addition of any cocat-
alyst (Fig. 25). The resulting polymers had Mn=1.5–1.8×104 and Mw/Mn=1.5–1.8,
and show atactic properties [62].
Thus no success has yet been achieved in synthesizing syndiotactic polysty-
rene with rare earth metal complexes, in contrast to the synthesis of highly
syndiotactic polystyrene with TiCl3(C5Me5)/(AlMe-O-)n (syndiotacticity
>95%) [63,64].
12
Stereospecific Polymerization of Conjugated Dienes
Organolanthanide(III) based binary initiator systems were used by Yu et al. [65]
for stereospecific polymerization of butadiene and isoprene. Typically, the po-
Organo Rare Earth Metal Catalysis for the Living Polymerizations of Polar and Nonpolar Monomers 277
system, i.e. 97.5% cis-selectivity for butadiene and 96.7% for isoprene. Al-
though bimetallic species like (CF3COO)EtNd(µ-Cl)(µ-H)AlEt2 are proposed
as active one, their exact structures are still unknown [69]. The molecular
structure of dimer complexes, [(CF3C)(µ2-O)2(µ3-O)4YAlEt2(THF2)]2 and
[(CF3)( µ2-O)3(µ3-O)3NdAlEt2(THF2)]2, generated during the reaction be-
tween (CF3COO)2NdC1 and AlEt3, has recently been elucidated by X-ray anal-
ysis. However, these complexes are considered to be byproducts because they
give polymers of low stereoregularity only at low yield [70].
A remarkable solvent effect was observed on the activity of the Nd
(OCOC7H15)3/Et2AlCl/iBu3Al system, which initiates the cis-1,4-polymerization
of butadiene (>98%) and isoprene (>95%). Aliphatic compounds such as pen-
tane, 1-pentene, and 2-pentene behave as good solvents, while aromatic com-
pounds such as toluene and mesitylene act as inhibitors [71]. The coordination
of aromatic compounds to the metal center may be responsible for the remarka-
ble suppression of the polymerization, as was the case in the polymerization us-
ing cobalt catalysts [72]. A binary initiator system, LnCl3[(BuO)3PO]3/AliBu3,
and a ternary system, Ln(naphthenate)3/AliBu3/Al2Et3Cl3 (naphthenate;
C10H7COO), also behave as good initiators for the cis-1,4-polymerization of iso-
prene [73]. In both cases, the polymerization activity varies with the nature of the
metal in the order Nd>Pr>Ce>Gd>La>>Sm>Eu. Thus, the activity increases
with increasing electronegativity and does so with decreasing M3+ ionic radius,
except for Sm3+ and Eu3+ which can be easily converted to the M2+ species in the
presence of a reducing agent. No relationship was observed between the initiating
activity and the Ln-O or Ln-Cl bond energy determined by IR and laser Raman
spectroscopy.
Block copolymerization of butadiene with isoprene (32:68–67:33) giving high
cis-1,4-polymers have also been successfully made with the Ln(naphthen-
ate)3/AliBu3/Al2Et3Cl3 system at temperature –78 to 33 °C. Noteworthy is the rel-
atively long life time of this initiator. Thus, it was possible to copolymerize iso-
prene 1752 h after the polymerization of butadiene. The (iPrO)2HLn2Cl3HAlEt2
species (Ln=Gd, Dy, Er, Tm) [74] prepared from either Ln(iPrO)3/Et2AlCl/Et3Al
or (iPrO)2LnCl/Et3Al also can initiate the cis-1,4-polymerization of butadiene
and isoprene. The most probable structure of this complex as evidenced by X-
ray analysisis is (iPrO)HLnEt(Cl)AlHEt(Cl)LnCl(OiPr). The polymerization ac-
tivity decreased in the order Gd>Dy>Er>Tm and the cis-content ranged from 92
to 95% in the case of Gd derivatives. Random copolymerization of butadiene
with isoprene was also performed using Nd(C6H6)(AlCl4)3/AlR3 (Al/Nd=30) in
benzene. Both monomers were incorporated in the copolymer selectively with
cis–1,4-butadiene 96.1–96.4% and cis–1,4-isoprene 97.5–98.3%. The conversion
increased with increase in the polymerization temperature from 0 (10%) to
80 °C (80–100%) [75].
Some rare earth metal based initiators induce the trans-1,4-polymerization
of conjugated dienes at high yield. The Ce(acac)3/AlEt2Cl system as well as the
CeCl3/AlEt2Cl and GdCl3/AlEt3 systems were most effective catalyst system to
carry out this type of polymerization for isoprene with high selectivity (91–
Organo Rare Earth Metal Catalysis for the Living Polymerizations of Polar and Nonpolar Monomers 279
97%) [76]. The marked difference in the selectivity between the Ce(acac)3/AlEt3
and the C5H5LnCl2/AlR3 or (iPrO)2LnCl/AlR3 initiator systems may be due to a
specific action of small amounts of water present in the system. Actually, metal
compound hydrates like NdCl3•6H2O, PrCl3•6H2O and UO2(C2H3O2)2•6H2O
can initiate the trans-polymerization in the presence of AlR3 [77]. Prolonged
aging of the initiator system decreased the activity significantly, presumably
owing to an irreversible self-reaction of the intermediate generated from the
organolanthanide and water. A precise recent study on the effect of water has
revealed that a maximum conversion was attainable at an H2O/AlEt2Cl ratio of
1.1–1.2, which produces (AlEt-O-)n(AlCl-O)m species. However, the molecular
weight was independent of the amount of water added to AlEt2Cl.
Butadiene-styrene copolymerization was attempted using the L3Ln-RX-AlR3
system [78]. Especially, (CF3COO)3Nd/C5H11Br/AliBu3 (1:3:15) was found to be
active for this type of copolymerization, with the cis-content of butadiene unit
amounting 97.8% and the styrene content to ca. 32%. However, for the system of
isoprene/styrene, the trans-1,4-polyisoprene copolymer was produced exclu-
sively.
13
Stereospecific Polymerization of Acetylene Derivatives
Polyacetylene (PA) is one of the simplest conjugated polymers, useful for man-
ufacturing lightweight high energy density plastics for storage batteries, solar
energy cells, etc. Acetylene can be polymerized to give high cis PA film with
Ti(OiBu)4/AlEt3 [79] or Co(NO3)2/NaBH4 [80] at temperatures lower than –
78 °C. Recently, it has been reported that Ln(naphthenate)3/AlR3/Donor (1:10:2–
3) (Donor=acetone, ether, ethyl acetate) systems also can initiate stereoregular
cis polymerization of acetylene at 30 °C, which leads to silvery metallic film [81,
82]. The polymer yield increased with increasing polymerization temperature
from –15 to 45 °C. A cis-polyacetylene with 95% selectivity was obtained when
the Al/Ln ratio was adjusted to ca. 5. The polymerization activity decreased in
the order Y=Ce>Nd=Tb>Pr>La>Lu>Gd>Tm=Er>Ho=Yb=Eu>Sm>Dy. The
trans content of the film amounted to 100% when the temperature was raised to
180 °C. The elements leading to PA film with a cis content exceeding 95% are La,
Pr, Nd, Sm, Gd, Tb, Dy, Ho, Er, Tm, and Y. The electrical conductivity of the films
was 294×10–8 Scm–1 for La, 181×10–8 Scm–1 for Nd, 194×10–8 Scm–1 for Gd,
490×10–8 Scm–1 for Tb, and 184×10–8 Scm–1 for Tm. Differential scanning calor-
imetry revealed two exothermic peaks at 200 and 380 °C and an endothermic
peak at 460 °C. These peaks were attributed to cis-trans isomerization, hydrogen
migration, and chain decomposition, respectively [82].
Sc(naphthenate)3/ROH/AlR3 (1/2/7) has been found to exhibit an activity
similar to the lanthanide series catalyst [83] (Fig. 27). The cis PA film obtained
with it showed an electrical conductivity of 14.4 Scm–1 when the polymer was
doped with I2 at a ratio of (CHI0.04)n, and the TEM measurement suggested the
formation of fibrils of about 20–30 nm in size.
280 Hajime Yasuda
The (P204)3Ln/AlR3 system also exhibited good activity for the polymeriza-
tion of acetylene when Al/Ln ratio was 5 [83]. The polymerization was conduct-
ed by the conventional method, and polymers with silver metallic appearance
were obtained. The addition of an oxygen-containing donor was effective for en-
hancing the polymerization rate and the cis-content. The effects were especially
marked for P204 (PO/Nd=1.1). The activity decreased in the order Nd=
Tb>Ce>Pr=Y>La>Er>Ho>Sm=Eu>Yb=Lu>Gd>Tm>Dy and the cis-content
decreased in the order Pr (95%)>Lu=Tb=Dy (92%)>Er=Y=Sm=Gd (87–89%).
The polymerization activities of Nd(P507)3, Nd(P204)3, and Nd(P215) were com-
pared and found to increase in this order, the result is consistent with the basic-
ities of the ligands (P507H=pKa 4.10, P204H=pKa 3.32, P215H=pKa 3.22). The M-
C bond is supposed to weaken as the electron donating ability of the ligand in-
creases. The Nd(iPrO)3/AlEt3 (Al/Nd=10) system [84] was also shown to be a
good initiator for the polymerization of acetylene. The soluble fraction obtained
was considered to be trans-polyacetylene, which was shown to have a molecular
weight of 277–540. Its 1H NMR spectrum revealed methyl groups at δ=
0.826 ppm and terminal vinyl groups at 4.95 ppm.
Phenylacetylene was polymerized to give a polymer of high cis configuration
by the use of the Ln(naphthenate)3/AlEt3 system [85, 86], with the activity de-
creasing in the order Gd>Lu>Nd=Ce>Ho>Sm>Dy=Eu>Er>Pr>La>Y=Tm>Yb,
and the cis-content exceeding 90%. It had Mn and Mw of 2×105 and 4×105, respec-
tively, and was crystalline according to XRD and SEM measurement. Its soften-
ing point was in the range 215–230 °C. Other terminal alkynes such as 1-hexyne,
1-pentyne, 3-methyl-1-pentyne, 4-methyl-1-pentyne, 3-methyl-1-butyne, and
phenylacetylene were found to polymerize quantitatively in the cis-fashion with
the Ln(naphthenate)3/AlR3/C2H5OH (Ln=Sc, Nd) or Ln(P204)3/ AlR3/C2H5OH
(1:7:3) system. The highest molecular Mn obtained was 16.8×104 for poly(1-pen-
tyne). Trimethylsilylacetylene was oligomerized to H(Me3SiC=CH)nCH2CHMe2
(n=2–3) by the use of LnX3(Donor)/AliBu3 (Ln=Gd, Pr, Nd, Tb, Dy, Lu; X=Cl, Br)
[87, 88]. The catalytic dimerization of terminal alkynes using
(C5Me5)2LnCH(SiMe3)2 (Ln=Y, La, Ce) has been reported recently. Here the
dimer was a mixture of 2,4-disubstituted 1-buten-3-yne and 1,4-disubstituted 1-
14
References
1. Klein JW, Lamps JP, Gnaonou Y, Remp P (1991) Polymer 32:2278
2. (a) Webster OW, Hertler WR, Sogah DY, Farnham WB, RajanBabu TV (1983) J Am
Chem Soc 105:5706; (b) Sogah DY, Hertler WR, Webster OW, GM Cohen (1987) Mac-
romolecules 20:1473
3. (a)Yasuda H, Yamamoto H, Yokota K, Miyake S, Nakamura A (1992) J Am Chem Soc
114:4908; (b)Yasuda H, Yamamoto H, TakemotoY, Yamashita M, Yokota K, Miyake S,
Nakamura A (1993)Makromol Chem Macromol Symp 67:187; (c) Yasuda H, Yamamoto
H, Yamashita M, Yokota K, Nakamura A, Miyake S, Kai Y, Kanehisa N (1993) Macro-
molecules 22:7134; (d) Yasuda H, Tamai H (1993) Prog Polym Sci 18:1097; (e) Yasuda
H, Ihara E (1997) Adv Polym Sci 133:53
4. Ihara E, Morimoto M, Yasuda H (1995) Macromolecules 28:7886
5. Yamashita M, Takemoto Y, Ihara E, Yasuda H (1996) Macromolecules 29:1798
6. Tanaka K, Ihara E, Yasuda H (1996) (unpublished results)
7. (a)Szwarc M (1983) Adv Polym Sci 49:1; (b) Nakahama S, Hirao A (1990) Prog Polym
Sci (1990) 13:299; (c)Inoue S (1988) Macromolecules 21:1195
8. (a) Sawamoto M, Okamoto O, Higashimura T (1987) Macromolecules 20:2693; (b) Ko-
jima K, Sawamoto M, Higashimura T (1988) Macromolecules 21:1552
9. (a) Gillon LR, Grubbs RH (1986) J Am Chem Soc (1986) 108:733; (b)Grubbs RH, Tumas
W (1989) Science 243:907; (c) Schrock RR, Feldman J, Canizzo LF, Grubbs RH (1989)
Macromolecules 20:1169
10. Cao ZK, Okamoto Y, Hatada K (1986)Kobunshi Ronbunshu 43:857
11. (a) Joh Y, Kotake Y (1976) Macromolecules 3:337; (b) Hatada K, Nakanishi H, Ute K,
Kitayama T (1986) Polym J 18:581
12. Abe H, Imai K, Matsumoto M (1968) J Polym Sci C23:469
13. Nakano N, Ute K, Okamoto Y, Matsuura Y, Hatada K (1989) Polym J 21:935
14. Bawn CEH, Ledwith A (1962) Quart Rev Chem Soc 16:361
15. Cram DJ, Kopecky KR (1959) J Am Chem Soc 81:2748
16. Hatada K, Ute K, Tanaka Y, Kitayama T, Okamoto (1986) Polym J 18:1037
17. Nitto Y, Hayakawa T, Ihara E, Yasuda H (1998) (unpublished result)
18. Yamamoto Y, Giardello MA, Brard L, Marks TJ (1995) J Am Chem Soc 117:3726
19. Boffa LS, Novak BM (1994) Macromolecules 27:6993
20. (a)Kuroki M, Watanabe T, Aida T, Inoue S (1991) J Am Chem Soc 113: 5903; (b) Aida T,
Inoue S (1996) Acc Chem Res 29:39
21. Sun J, Wang G, Shen Z (1993)Yingyong Huaxue 10:1
22. Collins S, Ward DG, Suddaby KH (1994) Macromolecules 27:7222
23. Soga K, Deng H, Yano T, Shiono T (1994) Macromolecules 27:7938
24. Hosokawa Y, Kuroki M, Aida T, Inoue S (1991) Macromolecules 24:8243
25. (a) Ihara E, Morimoto M, Yasuda H (1996) Proc Japan Acad 71:126; (b) Ihara E, Mori-
moto M, Yasuda H (1995) Macromolecules 28:7886; (c)Suchoparek M, Spvacek J (1993)
Macromolecules 26:102; (d)Madruga EL, Roman JS, Rodriguez MJ (1983) J Polym Sci
Polym Chem Ed 21: 2739
26. Kakehi T, Yasuda H (1997) (unpublished result)
282 Hajime Yasuda
27. (a)Patten TE, Novak BM (1991) J Am Chem Soc 113:5065; (b)Patten TE, Novak BM
(1993) Makromol Chem Macromol Symp 67:203
28. Fukuwatari N, Sugimoto H, Inoue S (1996) Macromol Rapid Commun 17:1
29. Tanaka K, Ihara E, Yasuda H (1998) (unpublished result)
30. (a) Hofman A, Szymanski R, Skomkowski S, Penczek S (1980) Makromol Chem
185:655; (b) Agostini DE, Lado JB, Sheeton JR (1971) J Polym Sci A1:2775
31. Cherdran H, Ohse H, Korte F (1962) Makromol Chem 56:187
32. Evans JE, Shreeve JL, Doedens RJ (1993) Inorg Chem 32:245
33. Zhang Y, Chen X, Shen Z (1989) Inorg Chim Acta 155:263
34. Wu J, Shen Z (1990) J Polym Sci Polym Chem Ed 28:1995
35. Itoh H, Shirahama H, Yasuda H (1998) (unpublished results)
36. Shen Z, Wu J, Wang G (1990) J Polym Sci Polym Chem Ed 28:1965
37. Wu J, Shen Z (1990) Polymer J 22:326
38. (a) Inoue S, Tsuruta T (1969) J Polym Sci Polym Lett Ed 7:287; (b) Kobayashi M., Inoue
S, Tsurtuta T (1973) J Polym Sci Polym Chem Ed 11:2383
39. (a) Shen X, Zhang Y, Shen Z (1994) Chinese J Polym Sci 12:28; (b)Shen Z, Chen X,
Zhang Y (1994) Macromol Chem Phys 195:2003
40. (a) Machoon JP, Sigwalt P (1965) Compt Rend 260:549; (b)Belonovskaja GP (1979) Eu-
rop Polym J 15:185
41. Dumas P, Spassky N, Sigwalt P (1972) Makromol Chem 156:65
42. Guerin PB, Sigwalt P (1974) Eur Polym J 10:13
43. Shen Z, Zhang Y, Pebg J, Ling L (1990) Sci China 33:553
44. Shen Z, Sun J, Wu Li, Wu L (1990) Acta Chim Sinica 48:686
45. Yao Y, Shen Q (private communication)
46. Stevels WM, Ankone JK, Dijkstra PJ, Feijen J (1996) Macromolecules 29:3332
47. Shen Y, Zhang F, Zhang Y, Shen Z (1995) Acta Polym Sinica 5:222
48. (a) Yasuda H, Ihara E (1993) J Synth Org Chem Jpn 51:931; (b)Yasuda H, Ihara E, Yosh-
ioka S, Nodono M, Morimoto M, Yamashita M (1994) Catalyst design for tailor-made
polyolefins. Kodansha-Elsevierp, p237
49. Nodono M, Ihara E, Yasuda H (unpublished result)
50. Fu P, Marks TJ (1995) J Am Chem Soc 117:10,747
51. Ihara E, Nodono M, Yasuda H, Kanehisa N, Kai Y (1996) Macromol Chem Phys
197:1909
52. Evans WJ, Ulibbari TA, Ziller JW (1988) J Am Chem Soc 110:6877
53. Evans WJ, Ulibbari TA, Ziller JW (1990) J Am Chem Soc 112:219
54. Lang DP, Bianconi PA (1996) J Am Chem Soc 118:52,453
55. Yasuda H, Furo M, Yamamoto H, Nakamra A, Miyake S, Kibino N (1992) Macromole-
cules 25:5115
56. Yang Y, Seyam AM, Fuad PF Marks TJ (1994) Macromolecules 27:4625
57. Yang M, Chan C, Shen Z (1990) Polym J 22:919
58. Kobayashi E, Aida S, Aoshima S, Furukawa J (1994) J Polym Sci Polym Chem Ed
32:1195
59. Hayakawa T, Ihara E, Yasuda H (1995) 69th National Meeting of Chem Soc Jpn 2B530
60. Cheng XC, Shen Q (1993) Chinese Chem Lett 4:743
61. Hu J, Shen Q (1993) Cuihua Xuebao 11:16
62. Tanaka K, Yasuda H (1997) (unpublished result)
63. Ishihara N, Seimiya T, Kuramoto M, Uoi M (1986) Macromolecules 19:2464
64. Ishihara N, Kuramoto M, Uoi M (1988) Macromolecules 21: 3356
65. Yu G, Chen W, Wang Y (1981) Kexue Tongabo 29:412
66. Ji X, Png S, Li Y, Ouyang J (1986) Scientia Sinica 29: 8
67. Jin S, Guan J, Liang H, Shen Q (1993) J Catalysis (Cuihua Xuebao) 159:14
68. Hu J, Liang H, Shen Q (1993) J Rare Earths 11:304
Organo Rare Earth Metal Catalysis for the Living Polymerizations of Polar and Nonpolar Monomers 283
69. (a) Jin Y, Li F, Pei F, Wang F, Sun Y (1994) Macromolecules 27:4397; (b) Li F, Jin Y, Pei K,
Wang F (1994) J Macromol Sci Pure Appl Chem A31:273
70. (a)Jin Y, Li X, Sun Y, Ouyang J (1982) Kexue Tongbau 27:1189; (b) Jin Y, Li X, Lin Y, Jin
S, Shi E, Wang M (1989) Chinese Sci Bull 34:390
71. Ricci G, Boffa G, Porri L (1986) Makromol Chem Rapid Commun 7:355
72. (a) Natta G, Porri L (1964) Polym Prepr 5:11,630; (b)Natta G, Polli L (1966) Adv Chem
Ser 52:24
73. Wong F (1994) J Macromol Sci Pure Appl Chem A31:273
74. Li X, Sun Y, Jin Y (1986) Acta Chim Sinica 44:1163
75. Hu Y, Ze L, Shen Q (1993) J Rare Earths 11:304
76. Lee DH, Wang JK, Ahn TO (1987) J Polym Sci Polym Chem Ed 25:1407
77. Lee DH, Ahn TO (1988) Polymer 29:71
78. Wang P, Jin Y, Pei F, Jing F, Sun Y (1994) Acta Polym Sinica 4:392
79. Ito T, Shirakawa H, Ikeda S (1974) J Polym Sci Polym Chem Ed 12:11
80. Luttinger LB (1962) J Org Chem 27:159
81. Shen Z, Wang Z, Can Y (1985) Inorg Chim Acta 110:55
82. Shen Z, Yang M, Shi M, Cai Y (1982) J Polym Sci Polym Chem Ed 20:411
83. Shen Z, Yu L, Yang M (1984) Inorg Chim Acta 109: 55
84. Hu X, Wang F, Zhao X, Yan D (1987) Chin J Polym Sci 5: 221
85. ZhaoJ, Yang M, Yuan Y, Shen Z (1988) Zhonggono Xitu Xuebao 6:17
86. Shen Z, Farona MF (1984) J Polym Sci Polym Chem Ed 22:1009
87. Shen Z, Farona MF (1983) Polym Bull 10:298
88. MullagalievI R, Mudarisova RK (1988) Izv Akad Nauk SSSR Ser Khim 7:1687
89. HeersH J, Teuben JH (1991) Organometallics 10:1980
90. Akita M, Yasuda H, Nakamura A (1984) Bull Chem Soc Jpn 57: 480
Polymer-Supported Rare Earth Catalysts Used in Organic
Synthesis
Shu Kobayashi
Abstract. Three types of polymer-supported rare earth catalysts, Nafion-based rare earth
catalysts, polyacrylonitrile-based rare earth catalysts, and microencapsulated Lewis acids,
are discussed. Use of polymer-supported catalysts offers several advantages in preparative
procedures such as simplification of product work-up, separation, and isolation, as well as
the reuse of the catalyst including flow reaction systems leading to economical automation
processes. Although the use of immobilized homogeneous catalysts is of continuing inter-
est, few successful examples are known for polymer-supported Lewis acids. The unique
characteristics of rare earth Lewis acids have been utilized, and efficient polymer-support-
ed Lewis acids, which combine the advantages of immobilized catalysis and Lewis acid-me-
diated reactions, have been developed.
Keywords: Lewis acids, Polymer-supported catalysts, Rare earth triflate, Combinatorial syn-
thesis, Carbon–carbon bond-forming reactions
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
2 Nafion-Based Rare Earth Catalysts . . . . . . . . . . . . . . . . . . 286
3 Polyacrylonitrile-Based Rare Earth Catalysts . . . . . . . . . . . . 289
4 A Microencapsulated Rare Earth Lewis Acid.
(A New Type of Polymer-Supported Catalyst) . . . . . . . . . . . . 295
5 References and Notes . . . . . . . . . . . . . . . . . . . . . . . . . . 303
1
Introduction
Use of polymer-supported catalysts offers several advantages in preparative pro-
cedures. Simplification of product work-up, separation, and isolation as well as
reuse of the catalyst including use of flow reaction systems could lead to eco-
nomical automation processes. Although the use of immobilized homogeneous
catalysts is of continuing interest [1], few successful examples are known for pol-
ymer-supported Lewis acids [2,3]. This is probably due to the instability of most
Lewis acids to air (moisture) and water. During preparation of polymer-sup-
ported catalysts, many manipulations have to be carried out in air or in the pres-
ence of water. On the other hand, it was recently found that some rare earth
Lewis acids are stable in water [4]. Utilizing this very unique characteristic, sev-
eral efforts to develop efficient polymer-supported Lewis acids combining the
advantages of immobilized catalysis and Lewis acid-mediated reactions, have
286 Shu Kobayashi
been made. In this chapter, three types of polymer-supported rare earth cata-
lysts, Nafion-based rare earth catalysts, polyacrylonitrile-based rare earth cata-
lysts, and microencapsulated Lewis acids are discussed [5].
2
Nafion-Based Rare Earth Catalysts
Recently, scandium triflate [Sc(OTf)3] was found to be stable in water and suc-
cessful Lewis acid catalysis was carried out in both water and organic solvents
[6–8]. Sc(OTf)3 coordinates to Lewis bases under equilibrium conditions, and
thus activation of carbonyl compounds using a catalytic amount of the acid has
been achieved [6,7]. In addition, effective activation of nitrogen-containing
compounds such as imines, amino aldehydes, etc. has been performed success-
fully [8]. Encouraged by the characteristics and the usefulness of Sc(OTf)3 as a
Lewis acid catalyst, a polymer-supported scandium catalyst was prepared.
Nafion (NR-50, Du Pont) was chosen as the supporting framework [9]. Three
equivalents of Nafion were treated with ScCl3◊6H2O in acetonitrile under reflux
conditions [10]. After 40 h, 96% of the ScCl3◊6H2O was consumed and the poly-
mer thus prepared (Nafion-Sc) contained 1.3% Sc, according to ICP analysis.
Choice of solvent is important at this stage; only 27% of the ScCl3◊6H2O was con-
sumed when 1,2-dichloroethane was used as the solvent. This Nafion-Sc catalyst
was then tested in several synthetic reactions [11]. First, allylation reactions of
carbonyl compounds were investigated. Allylation reactions of carbonyl com-
pounds are among the most important carbon–carbon bond-forming reactions,
and the products, homoallylic alcohols having hydroxyl and double bond
groups, are synthetically useful intermediates [12]. Nafion-Sc was also found to
be effective in the allylation reactions of carbonyl compounds with tetraallyltin,
and selected examples are shown in Table 1 [13]. In all cases, the reactions pro-
ceeded smoothly in both organic and aqueous solvents to afford the desired ho-
moallylic alcohols in high yields. Not only aldehydes, but also ketones worked
well. Moreover, use of aqueous solvents enabled the reactions of non-protected
carbohydrates [14]. Non-protected sugars reacted directly with tetraallyltin to
give the adducts, which are useful intermediates for the synthesis of higher sug-
ars. Salicylaldehyde and 2-pyridinecarboxaldehyde reacted with tetraallyltin to
afford the corresponding homoallylic alcohols in good yields. These compounds
are known to react with the Lewis acids rather than the nucleophile under gen-
eral Lewis acidic conditions.
Nafion-Sc was also found to be effective in some other reactions (Schemes 1–
3). In typical Lewis acid-mediated reactions, such as Diels–Alder, Friedel–Crafts
acylation, and imino Diels–Alder reactions, Nafion-Sc worked efficiently to af-
ford the corresponding adducts in high yields.
It was also found that Nafion-Sc could be easily recovered and reused. The
catalyst was recovered simply by filtration and washing with a suitable solvent,
and the activity of the recovered Nafion-Sc was comparable to the fresh catalyst;
in the reaction of benzaldehyde with tetraallyltin in H2O/MeOH/toluene(1:7:4)
Polymer-Supported Rare Earth Catalysts Used in Organic Synthesis 287
OH A 91
PhCHO B 90
Ph C 91
OH
CHO A 83
Ph C 83
Ph
OH
CHO A 93
Ph C 98
Ph
N CHO OH
N A 78
O
OH
Ph A 57 c
Ph
O
OH A 64 c
Ph Ph
C 87 c
O OH A 95 c
Ph Ph
C 85 c
O MeO2C OH A 81 c
Ph CO2Me Ph
C 84 c
OAc OAc
D-arabinose AcO A 76 c,d,e
OAc OAc
OAc OAc
D-ribose AcO A 91 c,d,f
OAc OAc
OAc OAc OAc
D-glucose
AcO A 64 c,d,g
OAc OAc
a The reaction was carried out at rt unless otherwise noted.
b A: H O/THF (1:9); B: H O/MeOH/toluene (1:7:4); C: CH CN.
2 2 3
c The reaction was carried out at 60 °C.
d The products were isolated after acetylation.
e syn/anti=59:41.
f syn/anti=65:35.
g Diastereomeric ratio=68:32. Relative configuration assignment was not made.
288 Shu Kobayashi
O O
Nafion-Sc O
+
N O CH2Cl2, rt, 2 h
92% CO N O
H2O/MeOH/Toluene
(1:7:4), rt, 2 h
97%
Scheme 1
Nafion-Sc
MeO + Ac2O MeO COCH3
LiClO4-CH3NO2
50 °C, 6 h
68%
Scheme 2
NH2
PhCOCHO•H2O + +
MeO
H
Nafion-Sc MeO
H
H2O/MeOH/Toluene
(1:7:4), rt, 2 h N COPh
92% H
Scheme 3
[15] at rt; lst, 91% (15 h); 2nd, 81% (48 h); 3rd, 93% (40 h); in the reaction of 3-
acryloyl-1,3-oxazolidin-2-one with cyclopentadiene at rt; 1st, 92% (2 h); 2nd,
87% (2 h); 3rd, 91% (2 h).
Nafion-Sc was also used in a flow system. Nafion-Sc was packed in a glass tube
and substrates were passed through the column using a motor pump (Fig. 1)
[16]. Results of the reaction of benzaldehyde with tetraallyltin in aqueous solu-
tion are shown in Scheme 4. It is noted that even higher yields were obtained in
the 2nd, 3rd, and 4th runs than in the 1st run.
Polymer-Supported Rare Earth Catalysts Used in Organic Synthesis 289
Nafion-Sc
pump
3
Polyacrylonitrile-Based Rare Earth Catalysts
One of the drawbacks of polymer-supported catalysts is their low reactivity.
Bearing in mind that the low reactivity might be ascribed to insolubility of the
catalysts, a new polymer-supported catalyst, which was partially soluble in an
appropriate solvent and is precipitated after completion of the reaction and re-
covered quantitatively by filtration, was searched. After several trials, a new
scandium catalyst, polyallylscandium trifylamide ditriflate (PA-Sc-TAD), was fi-
nally developed. PA-Sc-TAD was readily prepared according to Scheme 5. Poly-
acrylonitrile was treated with BH3◊SMe in diglyme for 36 h at 150 °C. The result-
ing amine (1) was reacted with Tf2O in the presence of Et3N in 1,2-dichlo-
290 Shu Kobayashi
CN n CH2NH2 n
1
CH2 CH 1. KH CH2 CH
CH2NH n 2. Sc(OTf)3 CH2NTf n
Tf
2 Sc(OTf)2
PA-Sc-TAD (3)
Scheme 5. Preparation of PA-Sc-TAD (3)
NH2
R3 R5
R1CHO + +
R4 R6
R3 R4
R2
PA-Sc-TAD (3) R5
R6
N R1
R2 H
Ph O
Cl Cl
N Ph N Ph N Ph
H H H
75%, 90/10 80%, 86/14 83%, 62/38
SPh SBu
Cl Cl
H,H
N Ph N Ph HN
H H
Ph
65%, nd 84%a 92%, 95/5
Cl
HN OMe
MeO HN HN
Ph Ph
Cl
Cl Cl
HN
HN HN
S
Cl
70%, 100/0 96%, 100/0 91%, 100/0
Cl Cl
Cl
HN HN
HN
COPh
O
78%, 100/0 80%, 100/0 99%, 100/0
NH2
PhCOCHO•H2O + +
PA-Sc-TAD (3)
CH2Cl2-CH3CN
(2:1) N COPh
H
1st use, 90%; 2nd use, 91%; 3rd use, 93% yield
Scheme 7
that more than 100-mg scale syntheses with a large array of diverse molecular
entries have been achieved with high purities (high yields and high selectivi-
ties). The number of commercially available aromatic aldehydes, aliphatic alde-
hydes, heterocyclic aldehydes, and glyoxals and glyoxylates that could be appli-
cable to this synthesis is more than 200, and more than 200 aromatic amines and
50 alkenes (and alkynes) [21] are commercially available. Therefore, a quinoline
library of more than a million compounds with high quantity and quality could
be prepared by using an automation system based on this method. Moreover, the
tetrahydroquinoline derivatives thus obtained are easily oxidized to dihydroqui-
noline or quinoline derivatives, which could double or triple the size of the li-
brary. Similarly, diverse amino ketone, amino ester, and amino nitrile deriva-
tives were synthesized using PA-Sc-TAD as a catalyst (Schemes 8 and 9) [22].
In addition to several advantages of polymer-supported catalysts in synthesis,
such as simplification of reaction procedures, easy separation of catalysts and
products, application to automation systems, etc., it is desirable to achieve high
selectivities by utilizing the bulkiness of polymers, the stability of polymer-co-
ordinated complexes, etc. [23]. Such examples have recently been reported using
a polymer-supported scandium catalyst (PA-Sc-TAD) [24]. It is well recognized
that aldimines are less reactive than aldehydes towards nucleophilic addition be-
cause of the difference in electronegativity between oxygen and nitrogen, the
steric hindrance of aldimines, etc. [25]. In fact, in the model competition reac-
tion of benzaldehyde and N-benzylideneaniline with propiophenone lithium
enolate, the enolate attacks the aldehyde exclusively in the coexistence of the ald-
imine. On the other hand, it was recently reported that in the presence of a cat-
alytic amount of Yb(OTf)3, aldimines react with silyl enolates selectively to af-
ford the corresponding adducts in high yields, even in the coexistence of alde-
hydes [26]. Scandium salts were then used in the same competition reactions.
Although both scandium and ytterbium elements belong to group 3, the compe-
tition reaction of benzaldehyde and N-benzylideneaniline with the silyl enol
ether of propiophenone in the presence of 20 mol% of Sc(OTf)3 gave a mixture
of aldehyde-adduct 4a (13%) and aldimine-adduct 5a (58%). Interestingly, when
Polymer-Supported Rare Earth Catalysts Used in Organic Synthesis 293
OSiMe3
2
R1CHO + R NH2 + R3
R4
R2HN O
PA-Sc-TAD (3)
R1 R4
3
R
Scheme 8
4a 5a
4 5
Yield (%)
R1 R2 R3 R4 4/5
4 5
Ph Ph Ph Me 1 97 1/99
c-C 6 H 11 Ph Ph Me trace 91 <1/>99
2-furyl Ph Ph Me trace >99 <1/>99
c-C 6 H 11 (p-Cl)Ph Ph Me trace 97 <1/>99
Ph Ph SEt H trace 58 <1/>99
R2
ScX3 O N ScX3 R2
k1 + k2
O N
R1 H R1 H
R1 H R1 H
+ ScX3
6
Scheme 12
OSiMe3
R4 O OH
3
R
6 R3 R1
4
R
Scheme 13
OSiMe3
R2
R4 O HN
R3
6
R3 R1
R4
Scheme 14
Polymer-Supported Rare Earth Catalysts Used in Organic Synthesis 295
Ph Ph
O OH O HN
O N OSiMe3 Catalyst
+
+ +
CH2Cl2, -23 °C Ph Ph Ph Ph
Ph H Ph H Ph
4a 5a
Sc(OTf)3 13 58 18/82
Sc(OTf) 2 (NTf 2 ) 16 77 17/83
Sc(NTf2 )3 3 77 4/96
ScCl 3 trace 43 <1/>99
Sc(PF 6 )3 2 76 3/97
Sc(SbF 6 )3 3 64 5/95
Sc(AsF 6 )3 26 49 35/65
Sc(ClO 4 )3 trace 87 <1/>99
Sc(BPh 4 )3 18 51 16/84
{[(3,5-CF3 )2 Ph] 4 B}3 Sc 10 66 13/87
Scheme 15
acid first coordinates an aldimine, but the coordination occurs under equilibri-
um conditions (Schemes 12–14). Aldehyde/Lewis acid or aldimine/Lewis acid
complexes should be stabilized by counter anions of scandium compounds
(Scheme 15), and the ratios of 4/5 shown in Scheme 15 could depend on the sta-
bility of the complexes influenced by the counter anions. On the other hand,
when the polymer-supported scandium catalyst (PA-Sc-TAD) was used, the cat-
alyst also coordinated aldimines first. In this case, however, the aldimine/poly-
mer-supported catalyst complexes are more stable than aldimine/nonpolymer
Lewis acid complexes due to the polymer effect. Hence, the aldimines activated
by the polymer-supported catalyst reacted with silyl enolates to afford aldimine
adducts exclusively.
These aldimine-selective reactions using the polymer-supported catalyst will
be applied to other nucleophilic addition reactions in organic chemistry.
4
A Microencapsulated Rare Earth Lewis Acid (A New Type of Polymer-
Supported Catalyst)
In Sects. 1 and 2, polymer-supported scandium Lewis acids based on Nafion and
a polyacrylonitrile derivative were discussed. While several useful reactions
296 Shu Kobayashi
Fig. 2. Scanning electron microscopy (SEM) micrograph (left) and scandium energy dis-
persive X-ray (EDX) map of MC Sc(Otf)3
have been developed, it was found that their reactivity, especially towards alde-
hydes and ketones, was lower than that of the nonpolymer Lewis acids. In gen-
eral, catalysts are immobilized on polymers via coordinate bonds or covalent
bonds. While the stability of polymer catalysts having coordinate bonds is often
low, preparation of polymer catalysts having covalent bonds can be troublesome
and their activities are generally lower than those of the nonpolymer catalysts
[27]. As an unprecedented polymer-supported Lewis acid, attention is focused
on a microencapsulated Lewis acid.
Microcapsules have been used for coating and isolating substances until such
time as their activity is needed, and their application to medicine and pharmacy
has been extensively studied [28]. Recently, much progress has been made in this
field, and, for example, the size of the microcapsules that can be achieved has
been reduced from a few µm to nanometers [28,29]. The idea is to apply this mi-
croencapsulation technique to the immobilization of Lewis acid catalysts onto
polymers. That is, Lewis acids would be physically enveloped by polymer thin
films and, at the same time, immobilized by the interaction between p electrons
of benzene rings of the polystyrene used as a polymer backbone and vacant or-
bitals of the Lewis acids [30]. Sc(OTf)3 was first chosen as a Lewis acid to be im-
mobilized [31]. Preparation of the microencapsulated Sc(OTf)3 [MC Sc(OTf)3]
was performed as follows: Polystyrene (1.000 g) was dissolved in cyclohexane
(20 ml) at 40 °C, and to this solution was added Sc(OTf)3 (0.200 g) as a core. The
mixture was stirred for 1 h at this temperature and then slowly cooled to 0 °C.
Coacervates were found to envelop the core dispersed in the medium, and hex-
ane (30 ml) was added to harden the capsule walls. The mixture was stirred at rt
for 1 h, and the capsules were washed with acetonitrile several times and dried
at 50 °C [32,33]. A scanning electron microscopy (SEM) micrograph and a scan-
dium energy dispersive X-ray (EDX) map of MC Sc(OTf)3 prepared are shown
in Fig. 2. It was found that small capsules of MC Sc(OTf)3 adhered to each other,
probably due to the small size of the core. Judging from the scandium mapping,
Polymer-Supported Rare Earth Catalysts Used in Organic Synthesis 297
Ph
Ph OSiMe3 MC Sc(OTf)3 NH O
N
+
Ph CH3CN, rt, 3 h Ph Ph
Ph H
Use a Yield/%
1 90
2 90
3 88
4 89
5 89
6 88
7 90
a
Recovered catalyst was used
successfully (Runs 2,3,4...)
Scheme 16. Imino aldol reaction (flow system)
Sc(OTf)3 was located all over the polymer surface. The importance of the ben-
zene rings of the polystyrene in immobilizing Sc(OTf)3 was demonstrated by
control experiments using polybutadiene or polyethylene instead of polysty-
rene. While 43% of Sc(OTf)3 [100%=the amount of Sc(OTf)3 immobilized by
polystyrene] was bound using polybutadiene, no Sc(OTf)3 was observed in the
microcapsules prepared using polyethylene. These results demonstrate that the
interaction between Sc(OTf)3 and the benzene rings of polystyrene is a key to
immobilizing Sc(OTf)3. Steric factors (physical envelopment) would also con-
tribute to the remarkable immobilizing ability of polystyrene compared to poor
immobilization of polybutadiene and polyethylene.
MC Sc(OTf)3 was used in several fundamental and important Lewis acid cat-
alyzed carbon–carbon bond-forming reactions. All the reactions were carried
out on a 0.5 mmol scale in acetonitrile [34] using MC Sc(OTf)3 containing ca.
0.120 g Sc(OTf)3. The reactions were carried out in both batch (using normal
vessels) and flow systems (using circulating columns). It was found that
MC Sc(OTf)3 effectively activated aldimines. Imino aldol (Sheme 16), aza Diels–
Alder (Scheme 17), cyanation (Scheme 18), and allylation (Scheme 19) reactions
of aldimines proceeded smoothly using MC Sc(OTf)3 to afford, respectively, the
synthetically useful b-amino ester, tetrahydroquinoline, a-aminonitrile, and
homoallylic amine derivatives in high yields.
Although it is well known that most Lewis acids are trapped and sometimes
decomposed by basic aldimines, MC Sc(OTf)3 worked well in all cases. One of
the most remarkable and exciting points is that the activating ability of
298 Shu Kobayashi
Use Yield/%
1 80
2 78
3 78
Use Yield/%
1 77
2 77
3 76
Use Yield/%
1 85
2 87
3 83
Fig. 3. Comparison of MC Sc(OTf)3 (O) and Sc(OTf)3 (") in the reaction of N-benzylide-
neaniline with (Z)-1-phenyl-1-trimethylsiloxypropene. The reaction was carried out using
N-benzylideneaniline (0.5 mmol), (Z)-1-phenyl-1-trimethylsiloxypropene (0.60 mmol),
and MC Sc(OTf)3 [1.167 g, containing Sc(OTf)3 0.120 g] or Sc(OTf)3 (0.120 g) in CH3CN
(5 ml) at rt
Use Yield/%
1 90
2 96
3 93
Use Yield/%
1 70
2 71
3 75
Use Yield/%
1 68
2 69
3 69
Use Yield/%
1 92
2 97
3 95
dium trifylamide ditriflate (PA-Sc-TAD) were found to have only a low ability to
activate aldehydes, as mentioned previously. On the other hand, MC Sc(OTf)3
efficiently activated aldehydes, and aldol (Scheme 23), cyanation (Scheme 24)
and allylation (Scheme 25) reactions of aldehydes proceeded smoothly to give
the corresponding aldol, cyanohydrin, and homoallylic alcohol derivatives in
high yields.
Polymer-Supported Rare Earth Catalysts Used in Organic Synthesis 301
Use Yield/%
1 79
2 78
3 74
Use a Yield/%
1 92
2 91
3 90
Use Yield/%
1 92
2 97
3 95
Use Yield/%
1 77
2 79
3 80
Use Yield/%
1 76
2 76
3 81
trial chemistry [37]. While the alkylation reaction proceeds in the presence of a
catalytic amount of a Lewis acid, the acylation reaction generally requires more
than a stoichiometric amount of a Lewis acid, usually aluminum chloride
(AlCl3), due to consumption of the Lewis acid by coordination to the aromatic
ketone products. It should be noted that MC Sc(OTf)3 has high activity in the
Friedel–Crafts acylation and can be recovered easily by simple filtration and re-
used without loss of activity.
Similarly, microencapsulated ytterbium triflate (MC Yb(OTf)3) has been pre-
pared [38]. MC Yb(OTf)3 was found to be effective in the aza-Diels–Alder reac-
tion using ethyl vinyl ether as a dienophile (Scheme 29). In the presence of
MC Yb(OTf)3, N-benzylideneaniline reacted with ethyl vinyl ether to afford the
corresponding tetrahydroquinoline derivative in a good yield. MC Yb(OTf)3
could also be recovered quantitatively and could be reused. The yield of the 2nd
and 3rd runs were comparable to that of the 1st run. In a similar reaction using
Sc(OTf)3 or MC Sc(OTf)3 as a catalyst, ethyl vinyl ether polymerized rapidly and
no desired adduct was obtained.
Polymer-Supported Rare Earth Catalysts Used in Organic Synthesis 303
Use Yield/%
1 78
2 70
3 77
5
References and Notes
1. (a) Bailey DC, Langer SH (1981) Chem Rev 81:109. (b) Akelah A, Sherrington DC (1981)
Chem Rev 81:557. (c) Frechet JM (1981) J Tetrahedron 37:663. (d) Smith K (ed) (1992)
Solid supports and catalysts in organic synthesis. Ellis Horwood, Chichester. (e) Sher-
rington DC, Hodge P (eds) (1988) Synthesis and separations using functional polymers.
John Wiley, New York. (f) Blazer HU, Pugin B (1995) In: Jannes G, Dubois J (eds) Chiral
reactions in heterogeneous catalysis. Plenum Press, New York
2. Recently, Lewis acid mediated reactions have been especially focused upon because of
their unique reactivities, selectivities, and mild conditions. Schinzer D (ed) (1989) Se-
lectivities in Lewis acid promoted reactions. Kluwer, Dordrecht; Santelli M, Pons J-M
(1995) Lewis acids and selectivity in organic synthesis. CRC Press, Boca Raton
3. Immobilization of AlCl3 onto polymers or inorganic support materials has been report-
ed. (a) Neckers DC, Kooistra DA, Green GW (1972) J Am Chem Soc 94:9284. (b) Krzy-
wicki A, Marczewski M (1980) J Chem Soc Faraday Trans 1:1311. (c) Drago RS, Getty EE
(1988) J Am Chem Soc 110:3311. (d) Clark JH, Martin K, Teasdale AJ, Barlow SJ (1995)
J Chem Soc Chem Commun 2037. Montmorillonite-supported zinc chloride. (e) Clark
JH, Cullen SR, Barlow SJ, Bastock TW (1994) J Chem Soc Perkin Trans 2 1117. Immobi-
304 Shu Kobayashi
lized trityl salts. (f) Mukaiyama T, Iwakiri H (1985) Chem Lett 1363. Polymer- or den-
drimer-bound Ti-TADDOLates. (g) Seebach D, Marti RE, Hintermann T (1996) Helv
Chim Acta 79:1710. (h) Rheiner PB, Sellner H, Seebach D (1997) Helv Chim Acta
80:2027. (i) Comina PJ, Beck AK, Seebach D (1998) Org Pro. Res Dev 2:18
4. (a) Kobayashi S (1991) Chem Lett 2187. (b) Kobayashi S, Hachiya I (1994) J Org Chem
59:3590. (c) Kobayashi S (1994) Synlett 689
5. Besides these examples, rare earth catalysts supported on ion-exchange resins have re-
cently been reported. Yu L, Chen D, Li J, Wang PG (1997) J Org Chem 62:3575
6. Kobayashi S, Hachiya I, Araki M, Ishitani H (1993) Tetrahedron Lett 34:3755
7. (a) Kobayashi S, Hachiya I, Ishitani H, Araki M (1993) Synlett 472. (b) Hachiya I, Koba-
yashi S (1993) J Org Chem 58:6958. (c) Hachiya I, Kobayashi S (1994) Tetrahedron Lett
35:3319. (d) Kawada A, Mitamura S, Kobayashi S (1994) Synlett 545. (e) Kobayashi S, Ar-
aki M, Hachiya I (1994) J Org Chem 59:3758. (f) Kobayashi S, Moriwaki M, Hachiya I
(1995) J Chem Soc Chem Commun 1527. (g) Kobayashi S, Moriwaki M, Hachiya I (1995)
Synlett 1153. (h) Kobayashi M, Hachiya I, Yasuda M (1996) Tetrahedron Lett 37:5569. (i)
Kobayashi S, Moriwaki M, Hachiya I (1997) Bull Chem Soc Jpn 70:267. (j) Kobayashi S,
Wakabayashi T, Nagayama S, Oyamada H (1997) Tetrahedron Lett 38:4559. (k) Koba-
yashi S, Wakabayashi T, Oyamada H (1997) Chem Lett 831
8. (a) Kobayashi S, Araki M, Ishitani H, Nagayama S, Hachiya I (1995) Synlett 233. (b)
Kobayashi S, Ishitani H, Nagayama S (1995) Chem Lett 423. (c) Kobayashi S, Ishitani H,
Nagayama S (1995) Synthesis 1195. (d) Kobayashi S, Hachiya I, Suzuki S, Moriwaki M
(1996) Tetrahedron Lett 37:2809. (e) Kobayashi S, Ishitani H, Komiyama S, Oniciu DC,
Katritzky AR (1996) Tetrahedron Lett 37:3731. (f) Kobayashi S, Moriwaki M, Akiyama
R, Suzuki S, Hachiya I (1996) Tetrahedron Lett 37:7783. (g) Kobayashi S, Ishitani H,
Ueno M (1997) Synlett 115. (h) Kobayashi S, Moriwaki M (1997) Tetrahedron Lett
38:4251. (i) Kobayashi S, Akiyama R, Moriwaki M (1997) Tetrahedron Lett 38:4819. (j)
Kobayashi S, Akiyama R, Kawamura M, Ishitani H (1997) Chem Lett 1039. (k) Koba-
yashi S, Busujima T, Nagayama S (1998) J Chem Soc Chem Commun 19. (l) Oyamada H,
Kobayashi S (1998) Synlett 249
9. As for metal Nafions, Hg: (a) Olah GA, Meidar D (1978) Synthesis 671. (b) Saimoto H, Hi-
yama T, Nozaki H (1985) Bull Chem Soc Jpn 56:3078. Si: (c) Murata S, Noyori R (1980)
Tetrahedron Lett 21:767. Cr, Ce: (d) Kanemoto S, Saimoto H, Oshima K, Nozaki H (1984)
ibid. 25:3317. Al,: (e) Waller FJ (1984) Br Polym J 16:239. Ta: (f) Olah GA, Gupta B, Farina
M, Felberg, JD, Ip WM, Husain A, Karpeles R, Lammertsma K, Melhotra AK, Trivedi NJ
(1985) J Am Chem Soc 107:7097. See also: (g) Waller F (1986) J Catal Rev Sci Eng 28:1
10. ScCl3◊6H2O (519 mg, 2.0 mmol) and Nafion (5 g, 1.2 meq/g) were combined in ace-
tonitrile (10 ml) under reflux for 40 h. After cooling to rt, the polymer was filtered,
washed with acetonitrile (20 ml¥3), and then dried under reduced pressure for 24 h. A
trial to prepare Nafion-Sc from Sc2O3 and Nafion failed (only 0.1% Sc was included in
the polymer (ICP analysis))
11. Kobayashi S, Nagayama S (1996) J Org Chem 61:2256
12. Yamamoto Y, Asao N (1993) Chem Rev 93:2207
13. Nafion 50 itself has much lower activity in this reaction. For example, only 19% yield of
the adduct was obtained in the reaction of acetophenone with tetraallyltin
14. (a) Schmid W, Whitesides GM (1991) J Am Chem Soc 113:6674. (b) Kim, E, Gordon DM,
Schmid W, Whitesides GM (1993) J Org Chem 58:5500. (c) Li C-J, Chan T-H (1997) Or-
ganic reactions in aqueous media. Wiley, New York
15. Kobayashi S, Hachiya I, Yamanoi Y (1994) Bull Chem Soc Jpn 67:2342
16. Nafion-Sc (2 g) and sea sand (2 g, 15–20 mesh) were combined and the mixture was
packed in a 6-mm glass tube. A carbonyl compound (0.5 mmol) and tetreallyltin
(0.5 mmol) were dissolved in acetonitrile (4 ml) and passed through the column using
a peristaltic pump (EYELA, MP-3)
Polymer-Supported Rare Earth Catalysts Used in Organic Synthesis 305
17. (a) Takahashi H, Kawakita T, Ohno M, Yoshioka M, Kobayashi S (1992) Tetrahedron Lett
48:5691. (b) Corey EJ, Imwinkelried R, Pikul S, Xiang YB (1991) J Am Chem Soc 111:5493
18. PA-Sc-TAD (3) is fully characterized by elementary analysis. 3: C, 23.51; H, 3.62; N, 3.33;
F, 12.5; S, 6.74; Sc, 2.0%
19. Kobayashi S, Nagayama S (1996) J Am Chem Soc 118:8977
20. Reviews: (a) Thompson LA, Ellman JA (1996) Chem Rev 96:555. (b) Früchtel JS, Jung G
(1996) Angew Chem Int Ed Engl 35:17. (c) Terrett NK, Gardner M, Gordon DW, Kobylecki
RJ, Steele J (1995) Tetrahedron 51:8135. (d) Lowe G (1995) Chem Soc Rev 37:309. (e) Gal-
lop MA, Barrett RW, Dower WJ, Fodor SPA, Gordon EM (1994) J Med Chem 37:1233. (f)
Gordon EM, Barrett RW, Dower WJ, Fodor SPA, Gallop MA (1994) J Med Chem 37:1385
21. Electron-deficient dienophiles will not work under the conditions, because the present
reactions are based on inverse electron-demand aza-Diels–Alder reactions
22. Kobayashi S, Nagayama S, Busujima T (1996) Tetrahedron Lett 37:9221
23. For example, (a) Bailar JC Jr (1974) Catal Rev Sci Eng 10:17. (b) Pittman CU, Jacobson
SE (1978) J Mol. Catal 3:293. (c) Dawans F, Morel D (1978) ibid. 3:403. (d) Carlini C,
Braca G, Ciardelli F, Sbrana G (1977) ibid. 2:379. (e) Capka M, Svoboda P, Cerny M, Het-
flejs J (1971) Tetrahedron Lett 3787. (f) Brown JM, Morinari H (1979) Tetrahedron Lett
2933. (g) Pinna F, Bonivento M, Strukul G, Graziani M, Cernia E, Palladino N (1976) J
Mol Catal 1:309
24. Kobayashi S, Nagayama S (1997) Synlett 653
25. For example, Yamaguchi M (1991) In: Trost BM (ed) Comprehensive organic synthesis,
vol 1. Pergamon, New York, chap 1.11
26. (a) Kobayashi S, Nagayama S (1997) J Org Chem 62:232. (b) Kobayashi S, Nagayama S
(1997) J Am Chem Soc 119:10049
27. Ciardelli F, Tsuchida E, Wöhrle D (eds) (1996) Macromolecule-metal complexes.
Springer, Berlin Heidelberg New York
28. Donbrow M (1992) Microcapsules and nanoparticles in medicine and pharmacy. CRC
Press, Boca Raton
29. Marty JJ, Oppenheim RC, Speiser P (1978) Pharm Acta Helv 53:17; Kreuter J (1978)
Pharm Acta Helv 53:33
30. (a) Greoffrey F, Cloke N (1995) In: Lappert MF (ed) Comprehensive organometallic
chemistry II. Pergamon, Oxford, vol 4, p 1. (b) Edelmann FT (1995) In: Lappert MF (ed)
Comprehensive organometallic chemistry II. Pergamon, Oxford, vol 4, p 11
31. Kobayashi S, Nagayama S (1998) J Am Chem Soc 120:2985
32. Judging from the recovered Sc(OTf)3 (0.080 g), 0.120 g of Sc(OTf)3 was microencapsulat-
ed according to this procedure. The weight of the capsules was 1.167 g which contained
acetonitrile. MC Sc(OTf)3 thus prepared can be stored at rt for more than a few months
33. IR (KBr) 3062, 3030 (vCH), 1946, 1873, 1805 (dCH), 1601, 1493 (benzene rings). 1255
(vasSO2), 1029 (vsSO2), 756 (vC-S), 696 (vS-O) cm–l. Cf. Sc(OTf)3: 1259 (vasSO2), 1032
(vsSO2), 769 (vC-S), 647 (vS-O) cm–l; polystyrene: 3062, 3026 (vCH), 1944, 1873, 1803
(dCH), 1600, 1491 (benzene rings) cm–l
34. Nitromethane was used in Friedel–Crafts acylation (Scheme 28)
35. Although the origin of the high activity of MC Sc(OTf)3 is not clear at this stage, it was
reported that aldimine–Lewis acid complexes were stabilized by using a polymer-sup-
ported Lewis acid. Cf. [24]
36. Cf. Ugi I, Domling A, Hörl W (1994) Endeavour 18:115; Armstrong RW, Combs AP, Tem-
pest PA, Brown SD, Keating TA (1996) Acc Chem Res 29:123
37. For leading references on Friedel–Crafts acylation, Olah GA (1973) Friedel–Crafts
chemistry. Wiley-Interscience, New York; Heaney H (1991) In: Trost BM (ed) Compre-
hensive organic synthesis. Pergamon, Oxford, vol 2, p 733; Olah GA, Krishnamurti R,
Prakash GKS (1991) In: Trost BM (ed) Comprehensive organic synthesis. Pergamon,
Oxford, vol 3, p 293
38. Kobayashi S, Nagayama S, Endo M, unpublished results
Lanthanide Triflate-Catalyzed Carbon-Carbon
Bond-Forming Reactions in Organic Synthesis
Shu Kobayashi
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
2 Aldol Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3 Diels-Alder Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4 Allylation Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5 Mannich-Type Reactions . . . . . . . . . . . . . . . . . . . . . . . . 81
10 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
1
Introduction
Lewis acid-catalyzed carbon-carbon bond forming reactions have been of great
interest in organic synthesis because of their unique reactivities, selectivities,
and for the mild conditions used [1]. While various kinds of Lewis acid-promot-
ed reactions have been developed and many have been applied in industry, these
reactions must be carried out under strict anhydrous conditions. The presence
of even a small amount of water stops the reaction because most Lewis acids im-
mediately react with water rather than the substrates and decompose or deacti-
vate, and this has restricted the use of Lewis acids in organic synthesis.
On the other hand, lanthanide compounds were expected to act as strong
Lewis acids because of their hard character and to have strong affinity toward
carbonyl oxygens [2]. Among these compounds, lanthanide triflates (Ln(OTf)3)
were expected to be one of the strongest Lewis acids because of the electron-
withdrawing trifluoromethanesulfonyl group. Their hydrolysis was postulated
to be slow, based on their hydration energies and hydrolysis constants [3]. In
fact, while most metal triflates are prepared under strict anhydrous conditions,
Lanthanide triflates are reported to be prepared in aqueous solution [4, 5]. The
large radius and the specific coordination number of lanthanide(III) are also
unique, and many investigations using Ln(OTf)3 as Lewis acids have been per-
formed.
In this paper, use of Ln(OTf)3 as Lewis acid catalysts in carbon-carbon bond-
forming reactions in both aqueous and organic solvents is overviewed.
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 65
2
Aldol Reactions
2.1
Aldol Reactions in Aqueous Media
The titanium tetrachloride-mediated aldol reaction of silyl enol ethers with al-
dehydes was first reported in 1973 [6]. The reaction is notably distinguished
from the conventional aldol reactions carried out under basic conditions; it pro-
ceeds in a highly regioselective manner to afford cross aldols in high yields [7].
Since this pioneering effort, several efficient activators such as trityl salts [8],
Cray montmorillonite [9], fluoride anions [10], etc. [11] have been developed to
realize high yields and selectivities, and now the reaction is considered to be one
of the most important carbon-carbon bond forming reactions in organic syn-
thesis. These reactions are usually carried out under strictly anhydrous condi-
tions. The presence of even a small amount of water causes lower yields, proba-
bly due to the rapid decomposition or deactivation of the promoters and the hy-
drolysis of the silyl enol ethers. Furthermore, the promoters cannot be recovered
and reused because they decompose under usual quenching conditions.
On the other hand, the water-promoted aldol reaction of silyl enol ethers with
aldehydes was reported in 1986 [12]. While the report that the aldol reactions
proceeded without catalyst in water was unique, the yields and the substrate
scope were not satisfactory. In 1991 the first example of Lewis acid catalysis in
aqueous media was reported; that is the hydroxymethylation reaction of silyl
enol ethers with commercial formaldehyde solution using Ln(OTf)3. Formalde-
hyde is a versatile reagent as one of the most highly reactive C1 electrophiles in
organic synthesis [13]. Dry gaseous formaldehyde required for many reactions
has some disadvantages because it must be generated before use from solid pol-
ymer paraformaldehyde by way of thermal depolymerization and it itself-po-
lymerizes easily [14]. On the other hand, commercial formaldehyde solution,
which is an aqueous solution containing 37% formaldehyde and 8–10% metha-
nol, is cheap, easy to handle, and stable even at room temperature. However, the
use of this reagent is strongly restricted due to the existence of a large amount of
water. For example, the titanium tetrachloride (TiCl4)-promoted hydroxymeth-
ylation reaction of silyl enol ethers was carried out by using trioxane as a HCHO
source under strict anhydrous conditions [6b, 15]. Formaldehyde water solution
could not be used because TiCl4 and the silyl enol ether reacted with water rather
than HCHO in that water solution.
The effects of Ln(OTf)3 in the reaction of 1-phenyl-1-trimethylsiloxypropene
(1) with commercial formaldehyde solution were examined [16]. In most cases
the reactions proceeded smoothly to give the corresponding adducts in high
yields. The reactions were most effectively carried out in commercial formalde-
hyde solution-THF media under the influence of a catalytic amount of Yb(OTf)3.
Several examples of the hydroxymethylation reaction of silyl enol ethers with
commercial formaldehyde solution are shown in Table 1, and the following char-
66 Shu Kobayashi
Table 1. Reaction of silyl enol ethers with commercial formaldehyde solution catalyzed by
Yb(OTf)3
Entry Silyl enol ether Product Yield/%
1 94
2 85
3 77
4 82
5 86
6 92
7 92
8 88
9 83
10 90
a Z/E=>98/2.
b Z/E=1/4.
c The mixture of the hydroxy thioester and the lactone (2:1) was obtained. The other dias-
tereomers were not observed.
d The mixture of the hydroxy thioester and lactone (3:1) was obtained. Less than 3% yield of
the other diastereomers was observed.
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 67
acteristic features of this reaction are noted. (1) In every case, the reactions pro-
ceeded smoothly under extremely mild conditions (almost neutral) to give the
corresponding hydroxymethylated adducts in high yields. Sterically hindered
silyl enol ethers also worked well and the diastereoselectivities were high. (2) Di-
and poly-hydroxymethylated products were not observed [17]. (3) The absence
of equilibrium (double bond migration) in silyl enol ethers allowed for a re-
giospecific hydroxymethylation reaction. (4) Only a catalytic amount of
Yb(OTf)3 was required to complete the reaction. The amount of the catalyst was
examined by taking the reaction of 1 with commercial formaldehyde solution as
a model, and the reaction was found to be catalyzed by even 1 mol% of
Yb(OTf)3: 1 mol% (90% yield); 5 mol% (90% yield); 10 mol% (94% yield);
20 mol% (94% yield); 100 mol% (94% yield).
The use of Ln(OTf)3 in the activation of aldehydes other than formaldehyde
was also investigated [18]. The model reaction of 1-trimethylsiloxycyclohexene
(2) with benzaldehyde under the influence of a catalytic amount of Yb(OTf)3
(10 mol%) was examined. The reaction proceeded smoothly in H2O-THF (1:4),
but the yields were low when water or THF was used alone. Among several
Ln(OTf)3 screened, neodymium triflate (Nd(OTf)3), gadolinium triflate
(Gd(OTf)3), Yb(OTf)3, and lutetium triflate (Lu(OTf)3) were quite effective,
while the yield of the desired aldol adduct was lower in the presence of lantha-
num triflate (La(OTf)3), praseodymium triflate (Pr(OTf)3) or thulium triflate
(Tm(OTf)3) (Table 2).
Lanthanide triflates were found to be effective for the activation of formalde-
hyde water solution. The effect of ytterbium salts was also investigated (Table 3)
[19]. While the Yb salts with less nucleophilic counter anions such as OTf– or
ClO4– effectively catalyzed the reaction, only low yields of the product were ob-
tained when the Yb salts with more nucleophilic counter anions such as Cl–,
OAc–, NO3–, and SO42– were employed. The Yb salts with less nucleophilic coun-
68 Shu Kobayashi
ter anions are more cationic and the high Lewis acidity promotes the desired re-
action.
Several examples of the present aldol reaction of silyl enol ethers with alde-
hydes are listed in Table 4. In every case, the aldol adducts were obtained in high
yields in the presence of a catalytic amount of Yb(OTf)3, Gd(OTf)3, or Lu(OTf)3
in aqueous media. Diastereoselectivities were generally good to moderate. One
feature in the present reaction is that water soluble aldehydes, for instance,
acetaldehyde, acrolein, and chloroacetaldehyde, can be reacted with silyl enol
ethers to afford the corresponding cross aldol adducts in high yields (entries 5–
7). Some of these aldehydes are commercially supplied as water solutions and
are appropriate for direct use. Phenylglyoxal monohydrate also worked well
(entry 8). It is known that water often interferes with the aldol reactions of metal
enolates with aldehydes and that in the cases where such water soluble aldehydes
are employed, some troublesome purifications including dehydration are neces-
sary. Moreover, salicylaldehyde (entry 9) and 2-pyridinecarboxaldehyde
(entry 10) could be successfully employed. The former has a free hydroxy group
which is incompatible with metal enolates or Lewis acids, and the latter is gen-
erally difficult to use under the influence of Lewis acids because of the coordina-
tion of the nitrogen atom to the Lewis acids resulting in the deactivation of the
acids.
The aldol reactions of silyl enol ethers with aldehydes were also found to pro-
ceed smoothly in water-ethanol-toluene [20]. Some reactions proceeded much
faster in this solvent system than in THF-water. Furthermore, the new solvent
system realized continuous use of the catalyst by a very simple procedure.
Several water-organic solvent systems were examined in the model reaction
of 1-phenyl-1-trimethylsiloxypropene (1) with 2-pyridinecarboxaldehyde un-
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 69
1 91 a)
2 89 b)
3 93 c)
4 81 d)
5 93 e, f)
6 82 e, g)
7 95 h)
8 67 i)
9 81 j, k)
10 87 j, l)
a syn/anti=73/27. g syn/anti=60/40.
b syn/anti=63/37. h syn/anti=45/55.
c syn/anti=71/29. i syn/anti=27/73.
d syn/anti=53/47. j Lu(OTf) was used instead of Yb(OTf)3.
3
e Gd(OTf) was used instead of Yb(OTf)3. k syn/anti=55/45.
3
f syn/anti=46/54. l syn/anti=42/58.
70 Shu Kobayashi
OH O
CHO OSiMe3 Yb(OTf) 3 (10 mol%)
+ Ph
N Ph solvent, rt, 4 h
N
der the influence of 10 mol% Yb(OTf)3 (Table 5). While the reaction proceeded
sluggishly in a water-toluene system, the adduct was obtained in a good yield
when ethanol was added to this system. The yield increased in accordance with
the amount of ethanol, and it was noted that the reaction proceeded much faster
in the water-ethanol-toluene system than in the original water-THF system.
Although the water-ethanol-toluene (1:7:4) system was one phase, it easily be-
came two phases by adding toluene after the reaction was completed. The prod-
uct was isolated from the organic layer by a usual work up. On the other hand,
the catalyst remained in the aqueous layer, which was used directly in the next
reaction without removing water. It is noteworthy that the yields of the second,
third, and fourth runs were comparable to that of the first run (Eq. 1).
(1)
1st run: 86% (syn /anti = 38/62)
2nd run: 82% (syn /anti = 38/62)
3rd run: 90% (syn /anti = 38/62)
4th run: 82% (syn /anti = 39/61)
Several examples of the present aldol reactions of silyl enol ethers with alde-
hydes in water-ethanol-toluene are listed in Table 6. 3-Pyridinecarboxaldehyde
as well as 2-pyridinecarboxaldehyde, salicylaldehyde, and formaldehyde water
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 71
Table 6. b(OTf)3-catalyzed aldol reactions of silyl enol ethers with aldehydes in water-etha-
nol-toluene
a)
89
95
87
82
96
90
a syn/anti=74/26.
solution worked well. As for silyl enol ethers, not only ketone enol ethers but also
silyl enolates derived from thioesters were used. In every case, the adducts were
obtained in high yields in the presence of 10 mol% Yb(OTf)3.
72 Shu Kobayashi
reaction mixture
water (quench)
(extraction)
product Ln(OTf)3
Scheme 1
2.2
Recovery and Reuse of the Catalyst
2.3
Aldol Reactions in Organic Solvents
Lanthanide triflates were found to be excellent Lewis acids not only in aqueous
media but also in organic solvents. The reaction of ketene silyl acetal 3 with ben-
zaldehyde proceeded smoothly in the presence of 10 mol% of Yb(OTf)3 in
dichloromethane at –78 °C, to afford the corresponding aldol-type adduct in a
94% yield. The same reaction at room temperature also went quite cleanly with-
out side reactions and the desired adduct was obtained in a 95% yield. No adduct
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 73
Scheme 2
1 Sc(OTf) 3 81
2 Y(OTf) 3 trace
3 Yb(OTf) 3 trace
Sc(OTf)3 (5 mol%) OH O
OSiMe3 CH2Cl2, -78 °C
CHO +
Ph Ph OMe
OMe 80%
3
Sc(OTf)3 (5 mol%)
OMe OSiMe3 CH2Cl2 OMe O
+
Ph OMe OMe -78 °C; 0% Ph OMe
3 0 °C to rt; 97%
Scheme 3
Sc(OTf)3 (5 mol%) OH O
OSiMe3 CH2Cl2, -78 °C, 1 h
PhCHO + (2)
88% (1st run) Ph OMe
OMe
3 89% (2nd run)
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 75
2.4
Aldol Reactions in Micellar Systems
Sc(OTf)3-catalyzed aldol reactions of silyl enol ethers with aldehydes were suc-
cessfully carried out in micellar systems. While the reactions proceeded sluggishly
in pure water, remarkable enhancement of the reactivity was observed in the pres-
ence of a small amount of a surfactant. In these systems, versatile carbon-carbon
bond-forming reactions proceeded in water without using any organic solvents.
Lewis acid catalysis in micellar systems was first found in the model reaction
of 1 with benzaldehyde (Table 8) [26]. While the reaction proceeded sluggishly
in the presence of 0.2 equiv. Yb(OTf)3 in water, remarkable enhancement of the
reactivity was observed when the reaction was carried out in the presence of
0.2 equiv. Yb(OTf)3 in an aqueous solution of sodium dodecylsulfate (SDS,
0.2 equiv., 35 mmol/l), and the corresponding aldol adduct was obtained in a
50% yield. In the absence of the Lewis acid and the surfactant (water-promoted
conditions), only 20% yield of the aldol adduct was isolated after 48 h, while a
33% yield of the aldol adduct was obtained after 48 h in the absence of the Lewis
acid in an aqueous solution of SDS. The amounts of the surfactant also influ-
Yb(OTf) 3 /0.2 48 17
Yb(OTf) 3 /0.2 SDS/0.04 48 12
Yb(OTf) 3 /0.2 SDS/0.1 48 19
Yb(OTf) 3 /0.2 SDS/0.2 48 50
Yb(OTf) 3 /0.2 SDS/1.0 48 22
Sc(OTf) 3 /0.2 SDS/0.2 17 73
Sc(OTf) 3 /0.1 SDS/0.2 4 88
Sc(OTf) 3 /0.1 TritonX-100 a)/0.2 60 89
Sc(OTf) 3 /0.1 CTAB/0.2 4 trace
76 Shu Kobayashi
88 a)
86 b)
88 c)
82 d)
88 e)
75 f,g)
94
84 g)
a syn/anti=50/50. e syn/anti=57/43.
b syn/anti=45/55. f Sc(OTf)3 (0.2 equiv.) was used.
c syn/anti=41/59. g Additional silyl enolate (1.5 equiv.) was
d Commercially available HCHO aq. (3 ml), charged after 6 h.
1 (0.5 mmol), Sc(OTf)3 (0.1 mmol), and
SDS (0.1 mmol) were combined.
enced the reactivity and the yield was improved when Sc(OTf)3 was used as a
Lewis acid catalyst. Judging from the critical micelle concentration, micelles
would be formed in these reactions, and it is noteworthy that the Lewis acid-cat-
alyzed reactions proceeded smoothly in micellar systems [27]. It was also found
that the surfactants influenced the yield, and that TritonX-100 was effective in
the aldol reaction (but required long reaction time), while only a trace amount
of the adduct was detected when using cetyltrimethylammonium bromide
(CTAB) as a surfactant [28].
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 77
3
Diels-Alder Reactions
Although many Diels-Alder reactions have been carried out at higher reaction
temperatures without catalysts, heat sensitive compounds in complex and mul-
tistep syntheses cannot be employed. While Lewis acid catalysts allow the reac-
tions to proceed at room temperature with satisfactory yields, they are often ac-
companied by diene polymerization and excess amounts of the catalyst are often
needed to catalyze carbonyl-containing dienophiles [29].
Lanthanide triflates were also found to be efficient catalysts in the Diels-Alder
reactions of carbonyl-containing dienophiles with cyclopentadiene [30]. A catalytic
amount of Yb(OTf)3 was enough to promote the reactions to give the correspond-
ing adducts in high yields, and the catalyst could be easily recovered and reused.
In the Diels-Alder reactions, Sc(OTf)3 was clearly the most effective among
Ln(OTf)3 as a catalyst [31]. While in the presence of 10 mol% of Y(OTf)3 or
Yb(OTf)3, only a trace amount of the adduct was obtained in the Diels-Alder re-
action of MVK with isoprene, and the reaction proceeded quite smoothly to give
the adduct in a 91% yield in the presence of 10 mol% of Sc(OTf)3.
Several examples of the Sc(OTf)3-catalyzed Diels-Alder reactions are shown
in Table 10. In every case, the Diels-Alder adducts were obtained in high yields
with endo selectivities.
The present Diels-Alder reactions proceeded even in aqueous media (Eq. 3)
[32]. Thus, naphthoquinone reacted with cyclopentadiene in THF-H2O (9:1) at
room temperature to give the corresponding adduct in a 93% yield (endo/exo=
100/0).
O
Sc(OTf)3 (10 mol%) O
+ (3)
THF : H2O ( 9 : 1 )
O O
93% yield, endo/exo = 100/ 0
Recovery and reuse of the catalyst were also possible in this reaction. After the
reaction was completed, the aqueous layer was concentrated to give the catalyst.
The recovered catalyst was effective in subsequent Diels-Alder reactions, and it
78 Shu Kobayashi
O O
O
95 87 / 13
N O O
CON
O
89 100 / 0
CON O
O
90
CON O
O
86
CON O
O O
O
97 84 / 16
N O O
CON
O
96 89 / 11
O
83 >95 / 5
O
O
91
O
73
O
O
83 100 / 0
O
O
O
89 94 / 3
O
O
92
O
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 79
should be noted that the yields of the second and even the third runs were com-
parable to that of the first run.
4
Allylation Reactions
OH
CHO Sn Yb(OTf)3 (10 mol%)
Ph +
4 Ph
H2O/EtOH/toluene
(1 : 7 : 4) (4)
rt, 8 h
1st run: 90%; 2nd run: 95%; 3rd run: 96%; 4th run: 89%
1 92
96
96
86
94
2 94
82
3 98
94
4 78
5 87
6 82
c)
7 81
d)
89
e)
93
f)
8 89
9 88f)
g)
10 78
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 81
11 quant.
90
quant.
12
84
13 87
14
90
ing homoallylic alcohols in high yields. The reactions were successfully carried
out in the presence of a small amount of a surfactant in water without using any
organic solvents.
5
Mannich-Type Reactions
5.1
Reactions of Imines with Silyl Enolates
The Mannich and related reactions provide one of the most fundamental and
useful methods for the synthesis of b-amino ketones or esters. Although the clas-
sical protocols include some severe side reactions, new modifications using pre-
formed iminium salts and imines have been developed [38]. Among them, reac-
tions of imines with enolate components, especially silyl enolates, provide useful
and promising methods leading to b-amino ketones or esters. The first report
using a stoichiometric amount of TiCl4 as a promoter appeared in 1977 [39], and
since then some efficient catalysts have been developed [40].
82 Shu Kobayashi
OSiMe3 Ln(OTf)3 R2
R2 (5 mol%) NH O
N R 3
+ R5
CH2Cl2, 0 °C R1 R5
R1 H R4 R3 R4
OSiMe3 Yb 65
SEt Sc 80
OSiMe3
Yb 80 b)
OBn
Ph
N
3 Yb 95
H
O OSiMe3
OBn
Yb 67 c)
N Ph
3 Yb 88
H
O
Ts
N
3 Yb 88
Ph H
Ph
N
Ph 3 Yb 60
H
O
N Ph
3 Yb 47
C4H9 H
a Second use=96% yield; third use=98% yield.
b syn/anti=18/82.
c syn/anti=21/79.
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 83
5.2
One-Pot Synthesis of β-Amino Esters from Aldehydes
While the Lewis acid-catalyzed reactions of imines with silyl enolates are one of
the most efficient methods for the preparation of b-amino esters, many imines
are hygroscopic, unstable at high temperatures, and difficult to purify by distil-
lation or column chromatography. It is desirable from a synthetic point of view
that imines, generated in situ from aldehydes and amines, immediately react
with silyl enolates and provide b-amino esters in a one-pot reaction. However,
most Lewis acids cannot be used in this reaction because they decompose or de-
activate in the presence of the amines and water that exist during imine forma-
tion. Due to the unique properties of Ln(OTf)3, their use as catalysts for the
above one-pot preparation of b-amino esters from aldehydes was planned.
A general scheme of the one-pot synthesis of b-amino esters from aldehydes
is shown in Scheme 4 [42]. In the presence of a catalytic amount of Yb(OTf)3 and
an additive (a dehydrating reagent such as MS 4 A or MgSO4), an aldehyde was
treated with an amine and then with a silyl enolate in the same vessel. Several
examples are shown in Tables 13 and 14, and the following characteristic fea-
tures of this reaction are noted.
1. In every case, b-amino esters were obtained in high yields. Silyl enolates de-
rived from esters as well as thioesters reacted smoothly to give the adducts.
R2
R3 OSiMe3 Yb(OTf)3 (5-10 mol%) NH O
R1CHO + R2NH 2 +
R4 R5 Additive, CH2Cl2, rt R1 R5
R3 R4
Scheme 4
84 Shu Kobayashi
MS4A 90
1 Ph Ph
MgSO4 89
2 Ph Bn MS4A 85
3 Ph p-MeOPh MgSO4 91 b)
4 Ph o-MeOPh MS4A 96
5 Ph Ph MS4A 90
6 Ph Bn MS4A 62 b)
p-MeOPh 79
7 Ph MS4A
84,b) 87 b,c)
8 Ph C4H9 MS4A 89
9 PhCO d) Ph MgSO4 82
10 PhCO d) Ph MgSO4 87
12 Ph(CH2)2 Bn MgSO4 83 f)
13 C4H9 Bn MgSO4 77 f)
14 C8H17 Bn MgSO4 81 f)
No adducts between aldehydes and the silyl enolates were observed in any re-
action.
2. A silyl enol ether derived from a ketone also worked well to afford the b-ami-
no ketone in a high yield (Table 13, entry 10).
3. Aliphatic aldehydes reacted with amines and silyl enolates to give the corre-
sponding b-amino esters in high yields. In some reactions of imines, it is
known that aliphatic enolizable imines prepared from aliphatic aldehydes
gave poor results.
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 85
R2
R3 OSiMe3 Yb(OTf)3 (5-10 mol%) NH O
1
R CHO + R 2NH
2
+
4 MS4A, C2H5CN, -78 °C
R R1 R4
R3
BnO OSiMe3
3 Ph(CH2) 2 Ph2CH 88 8.1 / 1
OMe
BnO OSiMe3
4 C4H9 Ph2CH 90 8.1 / 1
OMe
BnO OSiMe3
5 (CH3)2CHCH2 Ph2CH 86 7.3 / 1
OMe
a Determined by 1H NMR analysis.
4. Phenylglyoxal monohydrate also worked well in this reaction. The imine de-
rived from phenylglyoxal is unstable and a troublesome treatment is known
to be required for its use [43].
5. The catalyst could be recovered after the reaction was completed and could be
reused (first run, 91%; second run, 92%, in the reaction of benzaldehyde, p-
anisidine, and silyl enolate 3 (Table 13, entry 3)).
6. As for the diastereoselectivity of this reaction, good results were obtained af-
ter examination of the reaction conditions. While anti adducts were produced
preferentially in the reactions of benzaldehyde, syn adducts were obtained
with high selectivities in the reactions of aliphatic aldehydes (Table 14).
7. The high yields of the present one-pot reactions depend on the unique prop-
erties of Ln(OTf)3 as the Lewis acid catalysts. Although TiCl4 and TMSOTf are
known to be effective for the activation of imines [39, 44], the use of even sto-
ichiometric amounts of TiCl4 and TMSOTf instead of Ln(OTf)3 in the present
one-pot reactions gave only trace amounts of the product in both cases
(Table 15).
8. One-pot preparation of a b-lactam from an aldehyde, an amine, and a silyl
enolate has been achieved based on the present reaction (Scheme 5). The re-
action of the aldehyde, the amine, and 2 was carried out under the standard
conditions, and Hg(OCOCF3)2 was then added to the same pot. The desired
b-lactam was isolated in a 78% yield.
86 Shu Kobayashi
1 Yb(OTf) 3 10 mol% 89
Scheme 5
Thus, the one-pot synthesis of b-amino esters from aldehydes has been
achieved by using lanthanide triflate catalysis. The high efficiency using simple
starting materials and a catalytic amount of a reusable catalyst is especially note-
worthy.
5.3
Use of Acylhydrazones as Electrophiles in Mannich-Type Reactions
Hydrazones are aldehyde and ketone equivalents as well as imines. Their stabil-
ity is much higher than imines and actually hydrazones derived from aliphatic
aldehydes are often crystalline and can be isolated and stored at room tempera-
ture. However, their reactivity as electrophiles is known to be low, and there have
been many fewer reports on the reactions of hydrazones with nucleophiles than
those of imines [45, 46].
While 3-phenylpropionaldehyde phenylhydrazone did not react with ketene
silyl acetal 3b (R2=tBu) derived from methyl isobutyrate at all, 3-phenylpropion-
aldehyde acylhydrazones reacted with 3b in the presence of a catalytic amount
of Sc(OTf)3. Among the acylhydrazones tested, 4-trifluoromethylbenzoylhydra-
zone (4a, R1=CF3) gave the best yield (Table 16). It is noteworthy that the elec-
tronic effect of the benzoyl moieties influenced the yields dramatically. While
hydrazones with electron-donating groups gave lower yields, higher yields were
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 87
1 Sc(OTf)3 H t -20 37
2 Sc(OTf)3 MeO tBu -20 7
3 Sc(OTf)3 NO2 tBu -20 77
4 Sc(OTf)3 CF3 tBu -20 88
5 Sc(OTf)3 CF3 Me -20 97 (75a,b)
6 BF3•OEt 2 CF3 Me 0 4a
7 TiCl4 CF3 Me 0 13a
8 SnCl 4 CF3 Me 0 46a
a Dichloromethane was used as a solvent.
b 0 °C.
O
H Boc
N NH2 O NH O
HN O Raney Ni Boc2O
CF3 OMe OMe
OMe
5 6
71% (2 steps)
Scheme 6
at 70°C (Scheme 7). Since isomerization from 7 to 8 was observed under these
conditions (NaOMe), 7 and 8 were expected to be kinetic and thermodynamic
products, respectively. Moreover, pyrazolidinone 9 was obtained in the presence
of samarium diiodide (SmI2) [54] in THF-MeOH at 45°C.
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 89
O
H
N
n-BuLi, -78 °C N
CF3
93%
O
7
NaOMe, 70 °C (O2) N NH
5
99% O
8
SmI2, 45 °C HN NH
94% O
Scheme 7
5.4
Aqueous Mannich-Type Reaction
As mentioned in the previous sections, silyl enolates are excellent enolate com-
ponents in the Mannich-type reactions with imines. Alternatively, it was found
that vinyl ethers also reacted with imines smoothly in the presence of a catalytic
amount of Ln(OTf)3 to afford the corresponding b-amino ketones. In addition,
the reactions proceeded smoothly by the combination of aldehydes, amines, and
vinyl ethers in aqueous media [55].
A general scheme for the new Mannich-type reaction is shown below
(Scheme 8); the procedure is very simple. In the presence of 10 mol% of
Yb(OTf)3, an aldehyde, an amine, and a vinyl ether were combined in a solution
of THF-water (9:1) at room temperature to afford a b-amino ketone.
Selected examples of the present reaction are shown in Table 18. In all cases,
b-amino ketones were obtained in good yields. Several characteristic features
are noteworthy in this reaction. The procedure is very simple, consisting of sim-
ply mixing an aldehyde, an amine, a vinyl ether, and a small amount of Ln(OTf)3
in aqueous solution. The catalyst could be recovered after the reaction was com-
pleted and could be reused (first run, 93%; second run, 83%; third run, 87%, in
the reaction of phenylglyoxal•monohydrate, p-chloroaniline, and 2-methoxy-
propene). Commercially available formaldehyde and chloroacetaldehyde water
solutions were used directly and the corresponding b-amino ketones were ob-
tained in good yields. Phenylglyoxal•monohydrate, methyl glyoxylate, an
aliphatic aldehyde, and an a,b-unsaturated ketone also worked well to give the
corresponding b-amino esters in high yields. In some Mannich reactions with
preformed iminium salts and imines, it is known that yields are often low be-
cause of the instability of the imines derived from these aldehydes or trouble-
some treatments are known to be required for their use [43, 56].
90 Shu Kobayashi
Scheme 8
R1 R2 R3 Yield/%
H p-ClPh Me 92
H p-Ans Me 76
H p-Ans Ph quant.
Ph p-ClPh Me 90
Ph p-Ans Me 74
Ph(CH2 )2 p-ClPh Me 55
ClCH2 p-ClPh Me 59
PhCH=CH p-ClPh Me 73
PhCO p-ClPh Me 93
PhCO Ph Me 90
PhCO p-Ans Me 75
PhCO p-Ans Ph 85
MeO2C p-Ans Me 67
MeO2C p-Ans Ph 58
p-Ans=p-Anisidine
OMe
R2NH R2 2 _ R2 R2
2
N R3 R N OMe
H2O
NH OMe NH O
R1CHO OH
-H2O R1 R1 + R3 R1 R1 R3
H R3
Scheme 9
6
Aza Diels-Alder Reactions
6.1
Reactions of Imines with Dienes or Alkenes
The imino Diels-Alder reaction is among the most powerful synthetic tools for
constructing N-containing six-membered heterocycles, such as pyridines and
quinolines [58]. Although Lewis acids often promote these reactions, more than
stoichiometric amounts of the acids are required due to the strong coordination
of the acids to nitrogen atoms [58]. Use of Ln(OTf)3 as a catalyst was investigated
in this reaction.
R1 R1
R4
R3
N N
or
R1
Ph O Ph R2
R2 cat. Ln(OTf)3 12 13
N + R4
R4 CH3CN, rt
Ph H R3 R2 H
R1 R3 R1
10
or H
N Ph N Ph
H H
14 15
a Not determined.
b Relative configuration assignment was not made.
Ln Yield/%
Sc 94
Y 60
La 45
Pr 57
Nd 60
Sm 63
Eu 63
Gd 76
Dy 63
Ho 64
Er 97
Tm 92
Yb 85
Lu 72
6.2
Three-Components Coupling Reactions of Aldehydes, Amines, and Dienes or Alkenes
One synthetic problem in the imino Diels-Alder reactions is the imines' stability
under the influence of Lewis acids. It is desirable that the imines activated by
Lewis acids are immediately trapped by dienes or dienophiles [57]. In 1989, Sis-
ko and Weinreb reported a convenient procedure for the imino Diels-Alder re-
action of an aldehyde and a 1,3-diene with N-sulfinyl p-toluenesulfonamide via
N-sulfonyl imine produced in situ, by using a stoichiometric amount of BF3•OEt2
as a promoter [64].
Bearing in mind the usefulness and efficiency of one-pot procedures, three-
component coupling reactions between aldehydes, amines, and alkenes via imi-
ne formation and imino Diels-Alder reactions were examined by using Ln(OTf)3
as a catalyst.
In the presence of 10 mol% of ytterbium triflate and magnesium sulfate, ben-
zaldehyde was treated with aniline and 11 successively in acetonitrile at room
94 Shu Kobayashi
R2 Yb(OTf)3
(5-20 mol%) pyridine or quinoline
R1 CHO + + diene or alkene
CH3CN, rt derivatives
H 2N
Ph H 11 12a 80
(83)a)
15a 56 94/ 6
SPh 16a 70 nd b)
Cl 16c quant. nd b)
PhCO H 11 76
N COPh
Ph
19
H
R2
H H 94 96/ 4
N R1 (97)c) (96/ 4)
H
20a
OMe 20b 94 94/ 6
Cl 20c quant. 96/ 4
MeO2C H 21a 82 99/ 1
Cl 21c 84 99/ 1
SPh
Cl
Cl SPh 65 ndb)
N CO2Me
H
22
H
Cl
Hd) Cl H 90c)
N
H
23
a Sc(OTf)3 (10 mol%) was used. d Commercial formaldehyde water solution
b Not determined. was used.
c The reactions were carried out in aqueous
solution (H2O:EtOH:toluene=1:9:4).
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 95
Sc 63
Y 77
La 88
Pr 75
Nd 97
Sm 91
Eu 87
Gd 91
Dy 87
Ho 76
Er 84
Tm 84
Yb 94
Lu 80
6.3
Reaction Mechanism
10a - H+ -13.19 0.15 0.31 -0.39 -0.19 -5.75 0.63 -0.48 -0.09 0.21
10b - H+ -12.53 0.20 0.18 -0.46 -0.08 -5.58 0.62 -0.49 -0.07 0.21
10c - H+ -12.58 0.16 0.15 -0.37 -0.10 -5.77 0.63 -0.47 -0.19 0.21
a Calculated with MOPAC Ver. 6.01 using the PM3 Hamiltonian. MOPAC Ver. 7: Stewart JJP
(1989) QCPE Bull 10:9. Revised as Ver. 6.01 by Tsuneo Hirano, University of Tokyo, for HI-
TAC and UNIX machines (1989) JCPE Newslett 10:1.
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 97
NH2 OMe
(TfO)3Ln Ph-R
Ln(OTf)3
PhCHO + N
Ph H
R
H 2O
R R-Ph
NH OMe
(TfO)3Ln OH 27
- Ph
N
OMe MeOH
+
Ph
28 26
MeOH
24 25
(oxidation)
Scheme 10
NH2
OMe Yb(OTf)3 (10 mol%) PMP
NH O
PhCHO + + 26b +
MgSO4, CH3CN, rt
Ph
(8)
MeO 27b
PMP = p-MeO-Ph 34% 37%
Thus, a new type of Lewis acid, lanthanide triflates, is quite effective for the
catalytic activation of imines, and has achieved imino Diels-Alder reactions of
imines with dienes or alkenes. The unique reactivities of imines which work as
both dienophiles and azadienes under certain conditions were also revealed.
Three-component coupling reactions between aldehydes, amines, and dienes or
alkenes were successfully carried out by using Ln(OTf)3 as catalysts to afford py-
ridine and quinoline derivatives in high yields. The triflates were stable and kept
their activity even in the presence of water and amines. According to these reac-
tions, many substituted pyridines and quinolines can be prepared directly from
aldehydes, amines, and dienes or alkenes. A stepwise reaction mechanism in
these reactions was suggested from the experimental results.
7
Asymmetric Diels-Alder Reactions
OH
1.2 eq
OH (9)
2.4 eq. amine
Yb(OTf)3 “chiral Yb triflate”
MS 4A 0 °C, 30 min
O O “chiral Yb catalyst” CO N O
+ additive (20 mol%) (2S, 3R)
N O +
MS4A, CH2Cl2, 0 °C
O
CO N O
(2R, 3S)
O O
80 88/12 22.5/77.5 (55)
O O
36 81/19 19.0/81.0 (62)
O O 15.5/84.5 (69)
69 88/12
Ph 83 93/ 7 9.5/90.5 (81) b)
a Enantiomer ratios of endo adducts.
b 1,2,2,6,6-Pentamethylpiperidine was used instead of cis-1,2,6-trimethylpiperidine.
Yb(OTf)3, MS4A, and the additive were stirred in dichloromethane at 40 °C for 3 h.
obtained in 93% ee and the absolute configuration of the product was 2 S, 3R. On
the other hand, when acetyl acetone derivatives were mixed with the catalyst, re-
verse enantiofacial selectivities were observed. The endo adduct with an absolute
configuration of 2R, 3S was obtained in 81% ee when 3-phenylacetylacetone
(PAA) was used as an additive (catalyst B). In these cases, the chiral source was
the same (R)-(+)-binaphthol. Therefore, the enantioselectivities were controlled
by the achiral ligands, 3-acetyl-1,3-oxazolidin-2-one and PAA [73].
As shown in Table 26, the same selectivities were observed in the reactions of
other 3-acyl-1,3-oxazolidin-2-ones. Thus, by using the same chiral source ((R)-
(+)-binaphthol), both enantiomers of the Diels-Alder adducts between 3-acyl-
1,3-oxazolidin-2-ones and cyclopentadiene were prepared. Traditional methods
have required both enantiomers of chiral sources in order to prepare both enan-
tiomers stereoselectively [74], but the counterparts of some chiral sources are of
poor quality or are hard to obtain (for example, sugars, amino acids, alkaloids,
etc.). It is noted that the chiral catalysts with reverse enantiofacial selectivities
could be prepared by using the same chiral source and a choice of achiral ligands.
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 101
Table 26. Synthesis of both enantiomers of the Diels-Alder adducts between cyclopentadi-
ene and dienophiles by use of catalysts A and B
Dienophile Catalyst A
Yield/% endo/exo 2S,3R/2R,3S (ee (%))b)
O O (93)
77 89/11 96.5/ 3.5
N O
77 89/11 97.5/ 2.5 (95)c)
O O
O O
34 80/20 93.0/ 7.0 (86)
n N
Pr O
81 80/20 91.5/ 8.5 (83)c)
NR'3
H
O R
Yb(OTf)3 O
O site A CO N O
H
NR'3 O O O O (2S, 3R)
or
R N O N O
30
R N O
NR'3 O O
H
O site B
R O
Yb(OTf)3
CO N O
O site A
H (2R, 3S)
O O
NR'3
Ph
Scheme 11
Ln Catalyst B
Yield/% endo/exo 2S,3R /2R,3S (ee (%))a)
Lu 30 89/11 24.5/75.5 (51)
Yb 88 92/ 8 15.0/85.0 (70)
Tm 72 91/ 9 13.0/87.0 (74)
Er 59 90/10 13.0/87.0 (74)
Ho 70 84/16 21.0/79.0 (58)
Y 85 91/ 9 19.5/80.5 (61)
Gd 61 85/15 28.5/71.5 (43)
a
Enantiomer ratios of endo adducts.
10 84 86/14 96 (2S,3R )
5 84 87/13 93 (2S,3R )
3 83 87/13 92 (2S,3R )
20 99 89/11 93 (2R,3R)
10 96 90/10 97 (2R,3R)
20 95 78/22 74 (2S,3R )
10 86 78/22 75 (2S,3R )
fect of the lanthanide elements was also observed. In catalyst A, lutetium triflate
(Lu(OTf)3) was also effective in generating the endo Diels-Alder adduct in 93%
ee. The yields and selectivities diminished rapidly in accordance with the en-
largement of the ionic radii. In catalyst B, on the other hand, the best results were
obtained when thulium triflate (Tm(OTf)3) or erbium triflate (Er(OTf)3) was
employed. Deviations to either larger or smaller ionic radii resulted in decreased
selectivities, although the Diels-Alder adduct was obtained in an 85% yield with
good selectivities (endo/exo=92/8, endo isomer=61% ee) even when holmium
triflate (Ho(OTf)3) was used.
Although Sc(OTf)3 has slightly different properties compared with other lan-
thanide triflates, the chiral Sc catalyst could be prepared from Sc(OTf)3, (R)-(+)-
binaphthol, and a tertiary amine in dichloromethane [76]. The catalyst was also
found to be effective in the Diels-Alder reactions of acyl-1,3-oxazolidin-2-ones
with dienes. The amines employed in the preparation of the catalyst influenced
the enantioselectivities strongly. For example, in the Diels-Alder reaction of 3-
(2-butenoyl)-1,3-oxazolidin-2-one with cyclopentadiene (CH2Cl2, 0 °C), the
enantiomeric excesses of the endo adduct depended crucially on the amines em-
ployed; aniline, 14% ee; lutidine, 46% ee; triethylamine, 51% ee; 2,2,6,6-tetram-
ethylpiperidine, 51% ee; diisopropylethylamine, 69% ee; 2,6-dimethylpiperid-
ine, 69% ee; 1,2,2,6,6-pentamethylpiperidine, 72% ee; and cis-1,2,6-trimethyl-
piperidine, 84% ee.
Several examples of the chiral Sc(III)-catalyzed Diels-Alder reactions are
shown in Table 28. The highest enantioselectivities were observed when cis-
1,2,6-trimethylpiperidine was employed as an amine. 3-(2-Butenoyl)-, 3-cin-
namoyl-, and 3-(2-hexenoyl)-1,3-oxazolidin-2-ones reacted with cyclopentadi-
104 Shu Kobayashi
N
H
O
Ln(OTf)3
O
H
N
Scheme 12
ene smoothly in the presence of the chiral Sc catalyst to afford the corresponding
Diels-Alder adducts in high yields and high selectivities. It should be noted that
even 3 mol% of the catalyst was enough to complete the reaction and the endo
adduct was obtained in a 92% ee.
The catalyst was also found to be effective for the Diels-Alder reactions of an
acrylic acid derivative [77]. 3-Acryloyl-1,3-oxazolidin-2-one reacted with 2,3-
dimethylbutadiene to afford the corresponding Diels-Alder adduct in a 78%
yield and a 73% ee, whereas the reaction of 3-acryloyl-1,3-oxazolidin-2-one with
cyclohexadiene gave a 72% ee for the endo adduct (88% yield, endo/exo=100/0).
Similar to the chiral Yb catalyst, aging was observed in the chiral Sc catalyst.
It was also found that 30 or 3-benzoyl-1,3-oxazolidin-2-one was a good additive
for stabilization of the catalyst, but that reverse enantioselectivities by additives
were not observed. This can be explained by the coordination numbers of
Yb(III) and Sc(III); while Sc(III) has up to seven ligands, specific coordination
numbers of Yb(III) allow up to twelve ligands [69, 78].
As for the chiral ytterbium and scandium catalysts, the following structures
were postulated. The unique structure shown in Scheme 12 was indicated by
13C NMR and IR spectra. The most characteristic point of the catalysts was the
O O
O H O
Sc N N
H
TfO O
N OTf
diene
Scheme 13
2-one is effectively shielded by the amine part, and a diene approaches the dien-
ophile from the si face to afford the adduct in a high enantioselectivity.
It was also suggested that aggregation of the catalysts influenced the selectiv-
ities in the Diels-Alder reactions, and the reaction of 3-(2-butenoyl)-1,3-oxazo-
lidin-2-one with cyclopentadiene using (R)-(+)-binaphthol in lower enantio-
meric excesses was examined [84]. The results are shown in Fig. 1. Very interest-
ingly, a positive nonlinear effect was observed in the chiral Sc catalyst. In the chi-
ral Yb catalysts, on the other hand, the effect was dependent on the additives.
The extent of asymmetric induction in catalyst A did not deviate from the enan-
tiomeric excesses of (R)-(+)-binaphthol in the range 60–100% ee [85], while a
negative nonlinear effect was observed in catalyst B. These results can be as-
cribed to a difference in aggregation between the Sc catalyst, Yb catalyst A, and
Yb catalyst B.
8
Asymmetric Aza Diels-Alder Reactions
While asymmetric reactions using chiral Lewis acids have been demonstrated to
achieve several highly enantioselective carbon-carbon bond-forming processes
using catalytic amounts of chiral sources [86], chiral Lewis acid-catalyzed asym-
metric reactions of nitrogen-containing substrates are rare, probably because
most chiral Lewis acids would be trapped by the basic nitrogen atoms to block
the catalytic cycle. For example, aza Diels-Alder reactions are one of the most
basic and versatile reactions for the synthesis of nitrogen-containing heterocy-
clic compounds [58, 87]. Although asymmetric versions using chiral auxiliaries
or a stoichiometric amount of a chiral Lewis acid have been reported [88], exam-
ples using a catalytic amount of a chiral source are unprecedented.
In the previous section, lanthanide triflates were shown to be excellent cata-
lysts for achiral aza Diels-Alder reactions. While stoichiometric amounts of
Lewis acids are required in many cases, a small amount of the triflate effectively
catalyzes the reactions. On the other hand, chiral lanthanide Lewis acids have
been developed to realize highly enantioselective Diels-Alder reactions of 2-ox-
azolidin-1-one with dienes [89]. The reaction of N-benzylideneaniline with cy-
clopentadiene was first performed under the influence of 20 mol% of a chiral yt-
terbium Lewis acid prepared from ytterbium triflate (Yb(OTf)3), (R)-(+)-1,1'-
bi-naphthol (BINOL), and trimethylpiperidine (TMP). The reaction proceeded
smoothly at room temperature to afford the desired tetrahydroquinoline deriv-
ative in a 53% yield, although no chiral induction was observed. At this stage, it
was indicated that bidentate coordination between a substrate and a chiral Lewis
acid would be necessary for reasonable chiral induction. N-Benzylidene-2-hy-
droxyaniline (31a) was then prepared, and the reaction with cyclopentadiene
(32a) was examined. It was found that the reaction proceeded smoothly to afford
the corresponding 8-hydroxyquinoline derivative (33a) [90] in a high yield. The
enantiomeric excess of the cis adduct in the first trial was only 6%; however, the
selectivity increased when diazabicyclo-[5,4,0]-undec-7-ene (DBU) was used in-
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 107
stead of TMP (Table 29). It was also indicated that the phenolic hydrogen of 31a
would interact with DBU, which should interact with the hydrogen of (R)-(+)-
BINOL [91], to decrease the selectivity. Additives which interact with the phe-
nolic hydrogen of 31a were then examined. When 20 mol% of N-methylimida-
zole (MID) was used, 91% ee of the cis adduct was obtained; however, the chem-
ical yield was low. Other additives were screened and it was found that the de-
sired tetrahydroquinoline derivative was obtained in a 92% yield with high se-
lectivities (cis/trans>99/1, 71% ee), when 2,6-di-t-butyl-4-methylpyridine (DT-
BMP) was used.
Other substrates were tested, and the results are summarized in Table 30 [92].
Vinyl ethers (32b–32d) also worked well to afford the corresponding tetrahydro-
quinoline derivatives (33b–33e) in good to high yields with good to excellent di-
astereo- and enantioselectivities (entries 1–9). Use of 10 mol% of the chiral cat-
alyst also gave the adduct in high yields and selectivities (entries 2, 6). As for ad-
ditives, 2,6-di-t-butylpyridine (DTBP) gave the best result in the reaction of im-
ine 31a with ethyl vinyl ether (32b), while higher selectivities were obtained
when DTBMP or 2,6-diphenylpyridine (DPP) was used in the reaction of imine
31a with 32b. This could be explained by the slight difference in the asymmetric
environment created by Yb(OTf)3, (R)-(+)-BINOL, DBU, and the additive (see
below). While use of butyl vinyl ether (32c) decreased the selectivities (entry 7),
dihydrofurane (32d) reacted smoothly to achieve high levels of selectivity
(entries 8, 9). It was found that the imine (31c) prepared from cyclohexanecar-
108 Shu Kobayashi
O R1
O H
t Yb
Bu N N N
H
N H O
tBu
N
N R2
R3
Scheme 14
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 109
9
Catalytic Enantioselective 1,3-Dipolar Cycloadditions
1,3-Dipolar cycloadditions between nitrones and alkenes are most useful and
convenient for the preparation of isoxazolidine derivatives, which are readily
converted to 1,3-amino alcohol equivalents under mild reducing conditions
[94]. In spite of the importance of chiral amino alcohol units for the synthesis of
biologically important alkaloids, amino acids, b-lactams, and amino sugars, etc.
[95], catalytic enantioselective 1,3-dipolar cycloadditions remain relatively un-
explored [95, 96]. Catalytic enantioselective 1,3-dipolar cycloadditions of ni-
trones with alkenes using a chiral lanthanide catalyst were investigated [97, 98].
First, the 1,3-dipolar cycloaddition reaction of N-benzylidenebenzylamine N-
oxide with 3-(2-butenoyl)-1,3-oxazolidin-2-one was performed in the presence
of a chiral Yb(III) catalyst (20 mol%) prepared from Yb(OTf)3, (S)-1,1'-binaph-
thol ((S)-BINOL), and triethylamine (Et3N). The reaction proceeded smoothly
at room temperature to afford the corresponding isoxazolidine derivative in a
65% yield with high endo/exo selectivity (99/1), and a moderate ee of the endo
adduct was observed (Table 31). The enantiomeric excess was improved to 78%
when cis-1,2,6-trimethylpiperidine (TMP) was used instead of Et3N. Further-
110 Shu Kobayashi
endo exo
Table 31. Effect of amines
Amine Yield/% endo /exo ee/% b
Et 3 N 65 99/1 63
iPr2 NEt 73 >99/1 62
c
cis -1,2,6-TMP 73 99/1 78
(R )-MPEA d 92 >99/1 71
(S )-MPEA 80 97/3 35
e
(R )-MNEA 92 99/1 96
(S )-MNEA 87 99/1 62
a Chiral Yb(III)=Yb(OTf)3+(S)-BINOL+amine.
b Ee of the endo adducts.
c cis-1,2,6-Trimethylpiperidine.
more, it was found that use of chiral amines influenced the selectivity dramatical-
ly, and combination of the chirality of BINOL and the amine was crucial for the
selectivity. Namely, 71% ee of the endo adduct was obtained in the model reac-
tion using a catalyst prepared by the combination of (S)-BINOL and N-methyl-
bis[(R)-1-phenylethyl]amine ((R)-MPEA), while only 35% ee was observed by
the combination of (S)-BINOL and (S)-MPEA. Moreover, it was found that 96%
ee of the endo adduct was obtained with an excellent yield (92%) and diastereo-
selectivity (endo/exo=99/1) by the combination of (S)-BINOL and a newly-pre-
pared chiral amine, N-methyl-bis[(R)-1-(1-naphthyl)ethyl]amine ((R)-MNEA)
[99]. The chiral Yb(III) catalyst thus prepared has two independent chiralities
(hetero-chiral Yb(III) catalyst, see below) and it was found that the sense of the
chiral induction in these reactions was mainly determined by BINOL and that
the chiral amine increased or decreased the induction relatively.
Several examples of the 1,3-dipolar cycloadditions between nitrones and 3-
(2-alkenoyl)-1,3-oxazolidin-2-ones using the novel hetero-chiral Yb(III) catalyst
are shown in Table 32. In most cases, the desired isoxazolidine derivatives were
obtained in excellent yields with excellent diastereo- and enantioselectivities. It
is noted that high levels of selectivities were attained at room temperature. Ni-
trones derived from aromatic and heterocyclic aldehydes gave satisfactory re-
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 111
endo exo
Table 32. Catalytic enantioselective 1,3-dipolar cycloadditions
R1 R2 Yield/% endo/exo ee/%b
Ph CH3 92 99/1 96
p-Cl-Ph CH3 93 99/1 92
p-MeO-Ph CH3 82 95/5 90
2-furyl CH3 89 95/5 89
1-naphthyl CH3 88 98/2 85
Ph H 91 >99/1 79
Ph C 3H7 89 98/2 93
C2H5 CH3 88 53/47 96
a Chiral Yb(III)=Yb(OTf) +(S)-BINOL+(R)-MNEA.
3
b Ee of the endo adduts.
NMeR2
H
O
NMeR2
Yb(OTf)3
O
H
NMeR2
N
(R)-MNEA
Scheme 15
sults, and even in the reaction using the nitrone derived from an aliphatic alde-
hyde, the cycloaddition proceeded smoothly to give the endo adduct in an excel-
lent enantiomeric excess, albeit low endo/exo selectivity was observed. Moreo-
ver, it was found that alkenes which could be employed in the present 1,3-dipolar
cycloaddition were not limited to 3-(2-alkenoyl)-1,3-oxazolidin-2-one deriva-
tives. When N-phenylmaleimide was used as a dipolarophile, the desired isoxa-
zolidine derivative was obtained in a 70% yield with endo/exo>99/1, and the
enantiomeric excess of the endo adduct was 70% ee under the standard reaction
conditions [100, 101].
As for the structure of the hetero-chiral Yb(III) catalyst, the following struc-
ture was supported (Scheme 15). Actually, the existence of hydrogen bonds be-
tween the phenolic hydrogens of (S)-BINOL and the nitrogens of (R)-MNEA was
confirmed by the IR spectra of the catalyst [79, 102].
112 Shu Kobayashi
Bn O Bn O
N O MeOMgI N Pd/C, H2
(quant.) (65%)
Ph CON O Ph OMe
O
34
35
NH2 OR TBSO
H H
LDA (78%)
Ph Ph
CO2Me NH
O
TBSCl/imid. 36: R = H 38
(quant.) 37: R = TBS
Scheme 16
10
Conclusions
Lanthanide triflates are new types of Lewis acids different from typical Lewis ac-
ids such as AlCl3, BF3, SnCl4, etc. While most Lewis acids are decomposed or de-
activated in the presence of water, lanthanide triflates are stable and works as
Lewis acids in water solutions. Many nitrogen-containing compounds such as
imines and hydrazones are also successfully activated by using a small amount
of Ln(OTf)3. Lanthanide triflates are also excellent Lewis acid catalysts in organ-
ic solvents. A catalytic amount of Ln(OTf)3 is enough to complete reactions in
most cases. In addition, Ln(OTf)3 can be recovered after reactions are completed
and can be reused.
Recently, polymer-supported lanthanide catalysts have beem of great interest,
and these topics are discussed elsewhere. Use of lanthanide catalysts in solid-
phase organic synthesis is now well-recognized [107]. There have also been
many transformations other than carbon-carbon bond-forming reactions in or-
ganic synthesis using lanthanide triflates as catalysts, and all these will be re-
viewed in the near future.
Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions in Organic Synthesis 113
Acknowledgments. Our work in this area was partially supported by CREST, Ja-
pan Science and Technology Corporation (JST), and a Grant-in-Aid for Scientif-
ic Research from the Ministry of Education, Science, Sports, and Culture, Japan.
The author thanks and expresses his deep gratitude to his coworkers whose
names appear in the references.
11
References and Notes
1. Schinzer D (ed) (1989) Selectivities in Lewis acid promoted reactions. Kluwer Aca-
demic, Dordrecht
2. Molander GA (1992) Chem Rev 92:29
3. Baes CF Jr, Mesmer RE (1976) The hydrolysis of cations. Wiley, New York, p 129
4. Thom KF (1971) US Pat 3,615,169; CA (1972) 76:5436a
5. (a) Forsberg JH, Spaziano VT, Balasubramanian TM, Liu GK, Kinsley SA, Duckworth
CA, Poteruca JJ, Brown PS, Miller JL (1987) J Org Chem 52:1017; (b) Collins S, Hong Y
(1987) Tetrahedron Lett 28:4391; (c) Almasio MC, Arnaud-Neu F, Schwing-Weill M-J
(1983) Helv Chim Acta 66:1296; (d) Harrowfield JM, Kepert DL, Patrick JM, White AH
(1983) Aust J Chem 36:483
6. (a) Mukaiyama T, Narasaka K, Banno T (1973) Chem Lett 1011; (b) Mukaiyama T, Ban-
no K, Narasaka K (1974) J Am Chem Soc 96:7503
7. Mukaiyama T (1982) Org React 28:203
8. Kobayashi S, Murakami M, Mukaiyama T (1985) Chem Lett 1535
9. (a) Kawai M, Onaka M, Izumi Y (1986) Chem Lett 1581; (b) Kawai M, Onaka M, Izumi
Y (1988) Bull Chem Soc Jpn 61:1237
10. (a) Noyori R, Yokoyama K, Sakata J, Kuwajima I, Nakamura E, Shimizu M (1977) J Am
Chem Soc 99:1265; (b) Nakamura E, Shimizu M, Kuwajima I, Sakata J, Yokoyama K,
Noyori R (1983) J Org Chem 48:932
11. Lanthanide(III) chlorides or some organolanthanide compounds catalyzed aldol reac-
tions of ketene silyl acetals with aldehydes were reported: (a) Takai K, Heathcock CH
(1985) J Org Chem 50:3247; (b) Vougioukas AE, Kagan HB (1987) Tetrahedron Lett
28:5513; (c) Gong L, Streitwieser A (1990) J Org Chem 55:6235, (d) Mikami K, Terada
M, Nakai T (1993) J Chem Soc, Chem Commun 343 and references cited therein
12. (a) Lubineau A (1986) J Org Chem 51:2142; (b) Lubineau A, Meyer E (1988) Tetrahe-
dron 44:6065
13. For example: (a) Hajos ZG, Parrish DR (1973) J Org Chem 38:3244; (b) Stork G, Isobe
M (1975) J Am Chem Soc 97:4745; (c) Lucast DH, Wemple J (1976) Synthesis 724; (d)
Ono N, Miyake H, Fujii M, Kaji A (1983) Tetrahedron Lett 24:3477; (e) Tsuji J, Nisar M,
Minami I (1986) Tetrahedron Lett 27:2483; (f) Larsen SD, Grieco PA, Fobare WF (1986)
J Am Chem Soc 108:3512
14. Snider BB and Yamamoto K respectively developed formaldehyde-organoaluminum
complex as formaldehyde source in several reactions: (a) Snider BB, Rodini DJ, Kirk
TC, Cordova R (1982) J Am Chem Soc 104:555; (b) Snider BB (1989) In: Schinzer D (ed)
Selectivities in Lewis acid promoted reactions. Kluwer Academic Publishers, London,
pp 147–167; (c) Maruoka K, Conception AB, Hirayama N, Yamamoto H (1990) J Am
Chem Soc 112:7422; (d) Maruoka K, Concepcion AB, Murase N, Oishi M, Yamamoto H
(1993) J Am Chem Soc 115:3943
15. Cf. TMSOTf-mediated aldol-type reaction of silyl enol ethers with dialkoxymethanes
was also reported: Murata S, Suzuki M, Noyori R (1980) Tetrahedron Lett 21:2527
16. Kobayashi S (1991) Chem Lett 2187
17. Gaut H, Skoda J (1946) Bull Soc Chim Fr 13:308
114 Shu Kobayashi
38. Kleinman EF (1991) In: Trost BM (ed) Comprehensive organic synthesis, vol 2. Perga-
mon, Oxford, p 893
39. Ojima I, Inaba S-I, Yoshida K (1977) Tetrahedron Lett 3643
40. (a) Ikeda K, Achiwa K, Sekiya M (1983) Tetrahedron Lett 24:4707; (b) Mukaiyama T,
Kashiwagi K, Matsui S (1989) Chem Lett 1397; (c) Mukaiyama T, Akamatsu H, Han JS
(1990) Chem Lett 889; (d) Onaka M, Ohno R, Yanagiya N, Izumi Y (1993) Synlett 141.
For a stoichiometric use: (e) Dubois J-E, Axiotis G (1984) Tetrahedron Lett 25:2143; (f)
Colvin EW, McGarry DG (1985) J Chem Soc, Chem Commun 539; (g) Shimada S, Saigo
K, Abe M, Sudo A, Hasegawa M (1992) Chem Lett 1445. For an enantioselective ver-
sion: (h) Hattori K, Yamamoto H (1994) Tetrahedron 50:2785
41. Kobayashi S, Araki M, Ishitani H, Nagayama S, Hachiya I (1995) Synlett 233
42. Kobayashi S, Araki M, Yasuda M (1995) Tetrahedron Lett 36:5773
43. Lucchini V, Prato M, Scorrano G, Tecilla P (1988) J Org Chem 53:2251
44. Guanti G, Narisano E, Banfi L (1987) Tetrahedron Lett 28:4331
45. Selected examples using hydrazones as electrophiles or equivalents are as follows. As
far as we know, there have been no examples of Lewis acid-mediated reactions of hy-
drazones with nucleophiles: (a) Kodata I, Park J-Y, Yamamoto Y (1996) J Chem Soc,
Chem Commun 841; (b) Enders D, Ward D, Adam J, Raabe G (1996) Angew Chem, Int
Ed Engl 35:981; (c) Denmark SE, Weber T, Piotrowski DW (1987) J Am Chem Soc
109:2224; (d) Enders D, Schubert H, Nubling C (1986) Angew Chem, Int Ed Engl
25:1109; (e) Claremon DA, Lumma PK, Phillips BT (1986) J Am Chem Soc 108:8265; (f)
Takahashi H, Tomita K, Otomasu H (1979) J Chem Soc, Chem Commun 668
46. In pioneering work, rhodium-catalyzed enantioselective hydrogenation of acylhydra-
zones derived from ketones was reported: Burk MJ, Feaster JE (1992) J Am Chem Soc
114:6266
47. Oyamada H, Kobayashi S (1998) Synlett 249
48. Kobayashi S, Furuta T, Sugita K, Oyamada H (1998) Synlett (in press)
49. It has been shown that lanthanide triflates are excellent catalysts for the catalytic acti-
vation of imine and related compounds, while most Lewis acids are decomposed or
deactivated in the presence of basic nitrogen atoms. For example, (a) Kobayashi S, Bu-
sujima T, Nagayama S (1998) J Chem Soc, Chem Commun 19; (b) Kobayashi S, Na-
gayama S (1997) J Am Chem Soc 119:10,049; (c) Kobayashi S, Akiyama R, Kawamura
M, Ishitani H (1997) Chem Lett 1039; (d) Kobayashi S, Ishitani H, Ueno M (1997) Syn-
lett 115
50. Yb(OTf)3 was effective for the activation of the acylhydrazone, although the yield was
ca. 10% lower than that using Sc(OTf)3 in a preliminary experiment
51. It is also known that the imines derived from a,b-unsaturated aldehydes are often dif-
ficult to prepare due to their instability
52. Alexakis A, Lensen N, Mangeney P (1991) Synlett 625. Cf. Seebach D, Wykypiel W
(1979) Synthesis 423. See also [45]
53. The structure was confirmed by X-ray analysis of the corresponding N-benzoyl deriv-
ative
54. Cf. Overman LE, Rogers BN, Tellew JE, Trenkle WC (1997) J Am Chem Soc 119:7159
55. Kobayashi S, Ishitani H (1995) J Chem Soc, Chem Commun 1379
56. Other lanthanide triflates can also be used. In the reaction of phenylglyoxal•monohy-
drate, p-chloroaniline, and 2-methoxypropene, 90% (Sm(OTf)3), 94% (Tm(OTf)3),
and 91% (Sc(OTf)3) yields were obtained
57. Grieco et al. reported in situ generation and trapping of ammonium salts under Man-
nich-like conditions: (a) Larsen SD, Grieco PA (1985) J Am Chem Soc 107:1768; (b)
Grieco PA, Parker DT (1988) J Org Chem 53:3325 and references cited therein
58. (a) Weinreb SM (1991) In: Trost BM (ed) Comprehensive organic synthesis, vol 5. Per-
gamon, Oxford, p 401; (b) Boger DL, Weinreb SM (1987) Hetero Diels-Alder method-
ology in organic synthesis. Academic Press, San Diego, chaps 2 and 9
116 Shu Kobayashi
78. Review: Hart, FA (1987) In: Wilkinson G (ed) Comprehensive coordination chemistry,
vol 3. Pergamon, New York, p 1059
79. Fritsch J, Zundel G (1981) J Phys Chem 85:556
80. (a) Maruoka K, Yamamoto H (1989) J Am Chem Soc 111:789; (b) Bao J, Wulff WD,
Rheingold AL (1993) J Am Chem Soc 115:3814
81. Hattori K, Yamamoto H (1992) J Org Chem 57:3264
82. (a) Reetz MT, Kyung S-H, Bolm C, Zierke T (1986) Chem Ind 824; (b) Mikami K, Terada
M, Nakai T (1990) J Am Chem Soc 112:3949
83. Since the amine part can be freely chosen, the design of the catalyst was easier than
other catalysts based on (R)-(+)-binaphthol. Although some “modified” binaphthols
were reported to be effective as chiral sources, their preparations often require long
steps: Bao J, Wulff WD, Rheingold AL (1993) J Am Chem Soc 115:3814. See also [68a]
84. (a) Oguni N, Matsuda Y, Kaneko T (1988) J Am Chem Soc 110:7877; (b) Kitamura M,
Okada S, Suga S, Noyori R (1989) J Am Chem Soc 111:4028; (c) Puchot C, Samuel O,
Dunach E, Zhao S, Agami C, Kagan H (1986) J Am Chem Soc 108:2353
85. Lower enantiomeric excesses of the product were observed when (R)-(+)-binaphthol
in less than 60% ee was used. This may be ascribed to turnover of the catalyst
86. (a) Maruoka K, Yamamoto H (1993) In: Ojima I (ed) Catalytic asymmetric synthesis.
VCH, Weinheim, p 413; (b) Narasaka K (1991) Synthesis 1
87. (a) Kametani T, Kasai H (1989) Studies in Natural Product Chem 3:385; (b) Grigos VI,
Povarov LS, Mikhailov BM (1965) Izv Akad Nauk SSSR, Ser Khim 2163; CA (1966)
64:9680
88. (a) Waldmann H (1994) Synthesis 535; (b) Borrione E, Prato M, Scorrano G, Stiranello
M (1989) J Chem Soc, Perkin Trans 1 2245; (c) Ishihara K, Miyata M, Hattori K, Tada
T, Yamamoto H (1994) J Am Chem Soc 116:10,520
89. Kobayashi S, Ishitani H, Hachiya I, Araki M (1994) Tetrahedron 50:11,623
90. Some interesting biological activities have been reported in 8-hydroxyquinoline deriv-
atives. For example: (a) Rauckman BS, Tidwell MY, Johnson JV, Roth B (1989) J Med
Chem 32:1927; (b) Johnson JV, Rauckman BS, Baccanari DP, Roth B (1989) J Med Chem
32:1942; (c) Ife RJ, Brown TH, Keeling DJ, Leach CA, Meeson ML, Parsons ME, Reavill
DR, Theobald CJ, Wiggall KJ (1992) J Med Chem 35:3413; (d) Sarges R, Gallagher A,
Chambers TJ, Yeh L-A (1993) J Med Chem 36:2828; (e) Mongin F, Fourquez J-M, Rault
S, Levacher V, Godard A, Trecourt F, Queguiner G (1995) Tetrahedron Lett 36:8415
91. Kobayashi S, Ishitani H, Araki M, Hachiya I (1994) Tetrahedron Lett 35:6325
92. Ishitani H, Kobayashi S (1996) Tetrahedron Lett 37:7357
93. Inverse electron-demand asymmetric Diels-Alder reactions of 2-pyrone derivatives
were reported: (a) Posner GH, Carry J-C, Lee JK, Bull DS, Dai H (1994) Tetrahedron
Lett 35:1321; (b) Markó IE, Evans GR (1994) Tetrahedron Lett 35:2771
94. (a) Tufariello JJ (1984) In: Padwa A (ed) 1,3-Dipolar cycloaddition chemistry, vol 2.
Wiley, Chichester, p 83; (b) Torssell KBG (1988) Nitrile oxides, nitrones and nitronates
in organic synthesis. VCH, Weinheim
95. Review: Frederickson M (1997) Tetrahedron 53:403
96. Pioneering works in this field: (a) Seerden J-PG, Scholte OP, Reimer AWA, Scheeren
HW (1994) Tetrahedron Lett 35:4419; (b) Gothelf KV, Jørgensen KA (1994) J Org
Chem 59:5687; (c) Seerden J-PG, Kuypers MMM, Scheeren HW (1995) Tetrahedron
Asym 6:1441; (d) Gothelf KV, Thomsen I, Jørgensen KA (1996) J Am Chem Soc 118:59;
(e) Seebach D, Marti RE, Hintermann T (1996) Helv Chim Acta 79:1710; (f) Hori K,
Kodama H, Ohta T, Furukawa I (1996) Tetrahedron Lett 37:5947; (g) Jensen KB,
Gothelf KV, Hazell RG, Jørgensen KA (1997) J Org Chem 62:2471 and references cited
therein
97. It was recently found that lanthanide triflates are excellent catalysts in achiral 1,3-di-
polar cycloadditions between nitrones and alkenes and also in three-component cou-
pling reactions of aldehydes, hydroxylamines, and alkenes: Kobayashi S, Akiyama R,
118 Shu Kobayashi Lanthanide Triflate-Catalyzed Carbon-Carbon Bond-Forming Reactions
Kawamura M, Ishitani H (1997) Chem Lett 1039. Cf. Minakata S, Ezoe T, Ilhyong R, Ko-
matsu M, Ohshiro Y (1997) The 72nd Annual Meeting of the Chemical Society of Japan,
Tokyo, 2F3 37. See also [98]
98. Quite recently, Jøgensen et al. reported similar asymmetric 1,3-dipolar cycloadditions
using Yb(OTf)3-PyBOX; however, enantiomeric excesses obtained were up to 73%:
Sanchez-Blanco AI, Gothelf KV, Jøgensen KA (1997) Tetrahedron Lett 38:7923
99. (R)-MNEA was prepared from (R)-1-(1-naphthyl)ethylamine
100. It is believed that bidentate coordination (ex. Yb(III)-3-(2-alkenoyl)-1,3-oxazolidin-2-
one) is necessary to obtain high selectivities in many chiral lanthanide-catalyzed reac-
tions [101]. These results are very interesting and promising because it has been
shown that even monodentate coordination can achieve good selectivities by using the
hetero-chiral Yb(III) catalyst
101. [72, 73, 76, 89, 91]. See also (a) Marko IE, Evans GR (1994) Tetrahedron Lett 35:2771;
(b) Burgess K, Lim H-J, Porte AM, Sulikowski GA (1996) Angew Chem, Int Ed Engl
35:220
102. A bond pair (953 and 987 cm–1), which indicated hydrogen bonds (the OH•••N and O–
•••H+N equilibrium), was observed in the area from 930 to 1000 cm–1 in the IR spectra
of the hetero-chiral Yb(III) catalyst [103]. Direct coordination of the amine to Yb(III)
is doubtful in light of the fact that the 1,3-dipolar cycloaddition proceeded very slowly
when Yb(OTf)3 and (R)-MNEA were first combined and then (S)-BINOL was added
103. Evans DA, Morrissey MM, Dorow RL (1985) J Am Chem Soc 107:4346
104. Kametani T, Huang SP, Nakayama A, Honda T (1982) J Org Chem 47:2328
105. Sekiya M, Ikeda K, Terao Y (1981) Chem Pharm Bull 29:1747
106. As for carbapenems and penems: (a) Palomo C (1990) In: Lukacs G, Ohno M (eds) Re
cent progress in the chemical synthesis of antibiotics. Springer, Berlin Heidelberg New
York, p 565; (b) Perrone E, Franceschi G (1990) In: Lukacs G, Ohno M (eds) Recent
progress in the chemical synthesis of antibiotics. Springer, Berlin Heidelberg New
York, p 613
107. Kobayashi S (1998) Chem Soc Rev (in press)