You are on page 1of 18

4

EXTRACOORDINATION
AT SILICON

One of the most striking differences between carbon and silicon is the ready
ability of the latter to expand its coordination shell beyond four to five and six.
This chapter focuses on the ease with which coordination expansion occurs and
the properties of the resulting extracoordinate systems.'

4.1. EXTRACOORDINATION SILICON AS A LEWIS ACID

Lewis acids accept pairs of electrons. A simple example comes from the first row
of the periodic table. Boron compounds are trivalent, typically forming
compounds such as BF3 using three 0 bonds. Exposure to molecules containing
lone pairs leads to new compounds formed from dative bonding between the
Lewis acid and Lewis base (Eq. 4.1). The strength of these bonds can be seen
from the changes in boiling point upon coordination.

Gas: b.p. -1OOOC BF,


Et,O:
b.p. 35OC
- F,B OEt, Distillable Liquid: b.p. 126% (4.1)

Elements with d orbitals of appropriate energy can accommodate additional


Lewis base electron pairs, even if they possess four valence electron pairs. The
best-characterized Lewis acids are derived from elements found in Groups 4 (Ti,
Zr), 5 (V), 13 (Al), 14 (Si, Sn), 15 (Sb), and 16 (S, P)2; we shall focus on Ti, Sn,
and Si (many synthetic organic transformations of organosilanes are mediated by
Lewis acids derived from these elements, see Chapters 8 and 16). These three
elements form strong extracoordinate complexes with Lewis bases: The relative
order of Lewis acidity is Ti > Sn > Si. The energy of Lewis acid-Lewis base
bonds ranges from 42 to 126 kl mol- (10-30 kcal mol- 1).3
97
98 EXTRACOORDINATION AT SILICON

Many factors contribute to the magnitude of Lewis acidity. Hybridization to


extracoordinate silicon requires a suitable overlap of d orbitals with 3s and 3p
orbitals (sp3d, sp3d2: We equivocate about the actual orbitals involved in
extracoordination. d Orbitals are utilized for the discussion of Lewis acidity,
although increasingly calculations suggest that d orbitals are not important in the
bonding in extracoordinate systems. For a more detailed discussion of this point,
and the consideration of hypervalent and (J* orbitals and ionic effects, see
Chapter 2). The d-orbital contraction necessary for extracoordinationis provided
by strongly electronegative groups. In such cases, silicon is a reasonably strong
Lewis acid. This can been seen from gas-phase experiments in which the affinity
for fluoride of various main group compounds was determined (Table 4.1).4Note
that relatively high Lewis acidity is found with compounds bearing electro-
negative groups; any substitution of F by alkyl (or O R y leads to a significant
loss of Lewis acidity (see next section).
Table 4.1 shows that the general trend for ease of coordination expansion of
silanes is: SiX, > RSiX3 > R2SiX2> R3SiX. The facility for extracoordination
is dependent both upon the electronegativity and polarizability of the ligands on
silicon (see Chapter 5 ) and the Lewis basicity of the extracoordinating group.
SiF4 is a Lewis acid as a result of the electronegativity of fluorine. It is readily
converted into the perfectly stable, crystalline solid SiFz- (Scheme 4-1A): The
electron-withdrawing ability of fluoride, and the use of “naked fluoride” as the
incoming Lewis base, can even compensate for the presence of two aryl groups,
groups that reduce the ease of coordination expansion at silicon (Scheme 4- lB).’
By contrast, SiC162- is not stable: C1- is not sufficiently electronegative to
contract the d orbitals (Scheme 4-1C). On the other hand, the dipyridyl complex
4-1 is stable, a consequence of the better Lewis basicity of pyridine than C1-
(Scheme 4-1D).9However, the analogous MeSiC13-pyridinecomplex could not
be isolated: The Lewis acidity at silicon is too attenuated by the Me group

Table 41. Fluoride Atfinityof Some Maln Group Metal Compounds

Fluoride Affinity
Compound kJ mol-’ (kcal mol-’)
BF3 297 (71)
SiF4 285 (68)
(i-Pr)zBF 278 (66.5)
(i-Pr)3B 272 (65)
Et2BF 268 (64)
MezBF 257 (61.5)
MeSiF3 256 (61.3)
Me3B 245 (58.5)
MezSiFz 232 (55.5)
SF4 226 (54)
Adapted with permission from Ref. 4. Copyright 1977
American Chemical Society.
MORE COMPLEX PENTACOORDINATE COMPOUNDS 99

(Scheme 4-1E).1°The simple adduct between methyl ether and chlorosilane has
been characterized by X-ray crystallography: The H-Si-Cl bond angle is 98.3"
(Scheme 4-1F)."

SiF, - HO
,
H', SiFz- + SiO,
F; :
F-Si-F
F 'F 12-
KF
Ph,SiF, P Ph,SiF, K*
18-crownd

SiCI, - KCI
No reaction

MeSiCI, -
2 pyridine
No reaction
4-1

Scheme4-1

4.2. MORE COMPLEX PENTACOORDINATE COMPOUNDS

Our interest in extracoordinate systems is due less to the role of silicon as a


Lewis acid, and more to the role these compounds play in nucleophilic
substitution at silicon.12 Pentacoordinate systems can be readily prepared,
particularly when the silicon bears electron-withdrawing groups. In a surprising
example of extracoordination, in the absence of strongly electron-withdrawing
ligands, silica can be converted into pentacoordinate glycol esters simply by
refluxing in basic ethylene glycol (Eq. 4.2).13 With strongly silaphilic
nucleophiles (Lewis bases) such as imidazole, pentacoordinate systems derived
from dialkylsilanes can be prepared and isolated (Eq. 4.3).14 In exceptional
cases, pentaalkyl pentacoordinate systems 4-2 can be prepared! l5
12-
100 EXTRACOORDINATION AT SILICON

Several interesting examples of extracoordinate silanes come from Tacke and


co-workers,’6 based on the work of Frye et a1.” The addition of different
dihydroxy-organic species to aminomethylsilanes leads to crystalline penta-
coordinate structures, as exemplified in Scheme 4-2. These species are
zwitterionic: The positive charge is borne locally on the ammonium ion, rather
than on an external counterion. More electronegative li ands, such as carboxyl
oxygens, prefer to occupy the axial (apical) positions. I P

Scheme 4-2

Silatranes comprise a very special class of pentacoordinate species.19-**


Historically, these were the first stable pentacoordinate species reported (Scheme
4-3). The distal nitrogen on the triethanolamine cage forms a dative bond with
the silicon, the fifth bond. The bond length is intermediate between the van der
Waals contact distance (Table 4.2; sum of van der Waals radii of the two atoms)
and the bond length found in tetracoordinate silazanes. The nature of the
transannular Si-N bond has been the focus of study by theoretical calculationz3
and by NMR.24 Aza-analogues 4-3 have been described.25 As with many
hydrosilanes under basic conditions, cleavage of the Si-H bond (X=H)readily
occurs to yield, in this case, a dimer (Scheme 4-3; see also Chapter 7).

X
43

Scheme 4-3

Table 4.2. SI-N Bond Lengthsin DWerent Siiazanes

Si-N bond Bond length (A)


normal 1.74
van der Waals sum 3.5
in silatranes 2.19-2.34
REACTIVITY OF EXTRACOORDINATE SPECIES 101

In addition to their structural interest, silatranes have biological activity (see


Chapter 13). The nature of the ligand trans to the nitrogen is a strong determinant
in the nature of this activity. The toxicity, however, associated with some of these
compounds will likely preclude their use as drugs.
Transannular interactions, analogous to the Si-N interactions just discussed,
have been observed between sulfur and silicon. Holmes and co-workers have
prepared a series of compounds containing silicon in a seven-membered ring.
The interaction of the sulfur with silicon leads to a structure about half-way
between a tetrahedron and a trigonal bipyramid (Chart 4-1).26

Side on view

4.3. REACTIVITY OF EXTRACOORDINATE SPECIES


ELECTRONIC DISTRIBUTION

Tetracoordinate silicon atoms are usually positively polarized as a consequence


of the low electronegativity of silicon. What are the electronic consequences of
expanding the coordination shell at silicon by the addition of a pair of electrons
(neutral or negatively charged Lewis base)? Intuitively, one might expect the
silicon to become more negatively charged. However, calculations and
experiment suggest either that little change occurs or that the silicon becomes
more positively charged.
Ab initio calculations on SiH5- and related molecules suggest that silicon
atoms are more positively charged (Mulliken) in pentacoordinate structures than
in the neutral tetracoordinate analogues (e.g., S i b , Table 4.3).27-30 As is
sometimes the case with ab initio calculations, if one is prepared to spend more
money and do the calculations at a higher level of theory, different results can be
obtained. For example, Schleyer et aL31and Holmes and co-workers3*found that
very little change in positive polarization at silicon is observed on going from
tetra- to pentacoordinate: Further expansion to a hexacoordinate species leads to
an increase in the positive charge at silicon. Irrespective of the charge at silicon,
it is clear by all accounts that the ligands are more electron-rich in the
pentacoordinate than the tetracoordinate com ounds (Table 4.3).
The silicon atom is at or m~re~~.~'electrophilic in a pentacoordinate
than in a tetracoordinate system and, thus, at least as receptive to nucleophilic
attack (see below).33That is, the rate at which a tetracoordinate system 4-4 is
102 EXTRACOORDINATION AT SILICON

Table 4.3. CalculatedCharges onthe Atoms of Simple Silicon Complexes

Charge on Ligand
Chargea on
Calculation Compound central Si atom eq SP3 ax
CNDO/229 F4Si 1.214 -0.304
F5Si- 2.028 -0.399 -0.415
ab initio3' F4Si 2.46
F5Si- 2.40
F6si2- 2.65
ab initio3' F4Si 1.52
F5Si- 1.76
ab initio3' H4Si 0.63 -0.16
H5Si- 0.84 -0.29 -0.49
ab initio3' &Si 0.63
H5Si- 0.42
H6Si2- 0.45
ab initio3* &Si 0.5
H5Si- 0.58
'Mulliken charge.

converted into a pentacoordinate system 4-5 should be slower than the analogous
conversion of the pentacoordinate 4-6 to a hexacoordinate compound 4-7 (i.e.,
be,,> ktet, Scheme 4-4). This was determined to be the case in the reactions
shown in Scheme 4-5 ( k ~ - l> k ~150, kMeo-/kMe0 > Note that the
increased rate could have its origins elsewhere. For instance, the silicon could
be more accessible to nucleophiles as a pentacoordinate system.
Relative reaction rate towards nucleophiles

2-

- Nu-
k,",

4-5 44 46 4-7

Scheme4-4

A corollary of this higher reactivity at silicon in pentacoordinate systems is


that the ligands should be more nucleophilic than in the tetracoordinate
counterparts (Table 4.3). This has been shown to be the case with hydrides and
ally1 groups on silicon as nucleophiles (Chart 4-2).
REACTIONS OF PENTACOORDINATE SYSTEMS 103

K+ 18-crown6 K+ 18-crown-6 I

phJl
OMe
I -- Ph
Ph, si

I
OMe
i-PrMgBr_

ko~e- ,I,
OMe
Ph, I -
phc Si - Ph
Ph
',Si'Ph
Ph- I
OMe
- J,
CPrMgBr
kOMe
PhcSi
,Ph

K+ 18-crown-6 K+ 18-crown-6
I
Scheme 4-5

Relative nucleophilicity

H-donor HSi(OMe), < HSi(OR),

-
allyl donor F
si/3
\/ c b S i F ,
/ /
Chart 4 - 2 Relative Reactivity of Tetra- and Pentacoordinate Silanes

4.4. REACTIONS OF PENTACOORDINATE SYSTEMS

It is very likely that many nucleophilic substitutionreactions at silicon take place


via pentacoordinate systems.35 In the following discussion, we compare the
reactivity of five-coordinate systems that have been isolated with their four-
coordinate counterparts. In the following chapter, we examine the mechanism of
nucleophilic substitution at silicon, elaborating on the role of pentacoordinate
intermediates and transition states.

4.4.1. Extracoordinate Hydrosilanes: Reduction


Hydrosilanes are mild reducing agents (see Chapter 7). Normally, they require a
catalyst to initiate reaction. Corriu and co-workers compared the reactivity of a
naphthylphenylsilane4-8 with a pentacoordinate analogue: The amino group in
4-9 can form a dative bond to silicon similar to the transannular structure found
in the silatranes (Scheme 4-3).36 Under identical reaction conditions, with
alcohol as nucleophile, the four-coordinate system 4-8 showed no reactivity,
while 4-9 underwent efficient alcoholysis, yielding two equivalents of hydrogen
(see also Chapter 7). The latter system also reduced aldehydes to alkoxysilanes
without the need for additional catalysts, demonstrating the higher reactivity of
the pentacoordinate system when compared with the tetracoordinate compound
(Scheme 4-6).37
104 EXTRACOORDINATIONAT SILICON

d‘
- ROH No Reaction

+ 2H2

Scheme 4-6

Normally, an aldehyde is much more easily reduced than a carboxylic acid


(Scheme 4-7A). However, during reduction by a pentacoordinatehydrosilane, an
oxysilane complex is formed that does not undergo further reduction to alcohol:
Hydrolysis liberates the aldehyde and a siloxane (for example, EtCOzH+
EtCHO in 85% yield; Scheme 4-7B).38

A RC0,H -
1/4 LiAIH,
[RCHO] + RCO,H RCH,OH + RC0,H

- i) RC0,H

ii) HO
,
+ RCHO

Scheme 4-7

A different class of pentacoordinate system arises from the reaction of


trifunctional silanes with catechol and catecholates. The reaction between
trichlorosilane and dilithium catecholate leads to the formation of a reactive
pentacoordinate system 4-10 (Scheme The compound is a powerful
reducing agent, needing no further activation to reduce aldehydes (Scheme 4-
8B) or ketones (Scheme 4-8C,D). Related reducing agents [HSi(OMe)4-],
prepared from alkoxysilanes and alkoxides, behave in a similar manner,
enantioselectively in some ~ a s e s . 4 ~ ’ ~ ~

4.4.2. Extracoordinate Allylsilanes: Allylation


Extracoordinate allylsilanes can be formed by a variety of means from
allyltrifunctional silanes. These compounds are powerful allyl-nucleophiles.The
REACTIONS OF PENTACOORDINATE SYSTEMS 105

a O LOLii

o'.%
Scheme4-8

reaction with aldehydes leads to the formation of a new C-C bond (Eq. 4.4).42
Prochiral ally1 groups [e.g., ( E )- or (2)-crotyl groups] undergo these Grignard-
like reactions stereoselectively (Scheme 4-9A,43 B44;see also Chapter 16).
r 1-

KF (or NaOMe)
Me,"
PhCHO * HoY
(4.4)
Ph

A
SiF,
Ph -[Abh]-ph+
PhCHO
CsF OH Ph

bPh
Si(0Et)3 PhCHO
*
Ph
\(--..
OH
4- P h
OH
A P h

9O:lO

&OH OH
Scheme 4-9

The high regio- and stereoselectivity of the allylation strongly suggests that
the reaction proceeds via a six-membered cyclic transition state having a chair
conformation. Such a mechanism is consistent with the suggestion given above
106 EXTRACOORDINATION AT SILICON

that pentacoordinate compounds are more nucleophilic (ligands) and more


electrophilic (silicon) than their tetracoordinate analogues. Thus the suggested
reaction pathway involves nucleophilic conversion of the tetracoordinate silane
4-11 to the pentacoordinate compound 4-12. A further nucleophilic attack by the
aldehyde (carbonyl lone pair) gives the hexacoordinate silicate 4-13. Finally, the
cyclic, aldol-like transition state leads stereoselectively to the product 4-14,
which is liberated during hydrolysis (Scheme 4-10).45
Similar reactions can occur with less electrophilic silanes in the presence of
silaphilic nucleophiles (see Chapter 5). In situ silylation (Chapter 12) of (0-
crotyl chloride or hydrosilylation of butadiene leads to allyltrichlorosilanes.
With DMF (Me2NCHO) as the activating molecule, via extracoordinate
intermediates (see Chapter 5), allylation of benzaldehyde cleanly occurs with
high stereoselectivity in one pot (Scheme 4-11)?6 Another neutral reaction was
reported by Kira et al.47Utilizing a cycloheptatrieneone,which readily forms the
aromatic tropylium ion, they were able to prepare pentacoordinate species, with
subsequent ally1 transfer, under very mild conditions without the need of
catalysts. Thus conversion of tropanone to a putative hexacoordinate species
leads to the monoallylated compound 4-15 (Scheme 4-12).

4.5. HEXACOORDINATE COMPOUNDS

Hexacoordinate compounds were invoked as reaction intermediates in the


mechanisms shown in Schemes 4-10and 4-12. The readiness to allude to such

4-11 L 4-12
Et,NH'

P R2 h R' b

Et,NH' 4-14

Scheme 4-10

-cl N
(i) Et,N, CuCl
Ph HSiCI,
(ii) PhCHO, DMF (i) Et,N,
(ii) PhCHO.
CuCl
DMF* P h h

97:3 >99:1
Scheme 4-11
-
HEXACOORDINATE COMPOUNDS 107

+
- 2 eq. NEt,

qSiMeCt2
R

Scheme 4-12

compounds is the result of many reports of stable, hexacoordinate compounds,


including M2+SiFZ- and the zwitterionic compound 4-1648(Chart 4-3). One of
the most aesthetically beautiful structures of this type is the phthalocyanine
derivative 4-1749: Stacked assemblies of such compounds are electrically
cond~ctive.~'Catecholates have a special ability to stabilize extracoordinate
silicon, as shown in the pentacoordinate examples above. This is similarly true
for hexacoordinate silicon. Wolff and Weiss, for instance, have prepared the
polycyclic catecholate 4-18 (Eq. 4.5).51
/CH,

Chart4-3

+
&I;
Si(OEt),

Y
-
3

A related, extremely interesting, hexacoordinate system was isolated by


Corriu and c o - ~ o r k e r sThey
. ~ ~ demonstrated that Si02 could serve as a starting
material for the catecholate 4-19. Compound 4-19 can be converted into
functional alkylsilanes by the use of organometallic reagents (Scheme 4-13). In
an effort to demonstrate that these reactions could provide an alternative to the
direct process for the synthesis of functional alkylsilanes, further conversion of
the alkylsilylcatecholates was undertaken (Scheme 4-14). Unfortunately,
108 EXTRACOORDINATION AT SILICON

+ base
OH

2-

Bu,SiO
9
HO
(i) RU
(ii) K20
OSiR,

R',M

R ' = & N
Scheme 4-13
40"
t
R,SiH
R = Et (53%), n-BU (68%)

LiAIH,

I
R,SiOMe R3SiCI

COCI, MeOLi /Et (6O%), n-Bu (84%)

R,SiBr
R = n-Bu (83%)
- HBr
XMgO
- RMgBr
R,SiR
R=Et, R=Ph (SO%),

\
n-BuCC (49%), PhCC (67%)

k R
3
: b30+
n-Bu, R = Me (69%)

MgBr

R,SiMPh,
R=Et, M=Si (68%), Ge (58%)

Ho
D R=Et (72%),n-Bu (82%) R=Et (64%)
SIR3

MeOLi
I
R,SiOMe
MeOH

R = n-Bu (93%)
Scherne4-14

dialkylsilanes, which may be converted into silicones (see Chapter 9), could not
be prepared in high yield.
A synthetically useful hexacoordinate species is the alkylpentafluorosilicate
4-20,53although the utility of these compounds is tempered somewhat by their
HEXACOORDINATE COMPOUNDS 109

low solubility. Kumada and co-workers outlined the chemistry that these species
undergo in the early 1980s. The addition of fluoride to an alkyltrichlorosilane
leads to halide exchange and coordination expansion. The nucleophilicity of the
alkyl group in the pentafluorosilicate was found to be much higher than in the
starting alkyltrichlorosilane.Electrophilic attack with halogens and NBS,on the
Si-C bond of the pentacoordinate system, leads to the alkyl halides with
inversion at carbon. By contrast, reaction with mCPBA leads to the alcohol with
retention of configuration at carbon (Scheme 4-15)?4*55In the presence of
copper catalysts (CuX2) the halide is formed, but with racemization.
These results have been explained in terms of two competing reactions:
electron transfer with hexacoordinate systems56and an associative mechanism
with pentacoordinate species. In polar solvents, where one of the fluoride ligands
in RSiF52- can dissociate to form the pentacoordinate system, subsequent
association by mCPBA (in analogy to the carbonyl association in Scheme 4-10)
leads to a six-coordinate intermediate from which oxidation occurs at carbon. In
contrast, in less polar solvents the electron-rich RSiF5’- undergoes electron
transfer processes.
Alkenylpentafluorosilicates undergo an even broader range of reactions,
including oxidation and palladation (Scheme 4-16).57The palladation reaction is

HSiCl
R4
3_ Rd
H,PtCI,

h0...
CI

MPBA

R-OH R 4 X
0
X, = Br2,CI,. I,. NBS
IBr
Scheme 4-15

Scheme 4-16
110 EXTRACOORDINATIONAT SILICON

e.g. Z=CO,Me. CN

Scheme 4-17

especially useful5* because of the organometallic coupling reactions in which


the alkenylpalladium species participates (Scheme 4-17).59

4.6. HIGHER EXTRACOORDINATE SPECIES

Corriu and co-workers have demonstrated that it is possible to extend


extracoordination beyond octahedral systems. A heptacoordinate transition
state was inferred from the dynamic behavior of pentacoordinate systems that
bear two distal nitrogen ligands 4-21 (5+2 coordination).6' The stable
hexacoordinate system 4-22, for which a crystal structure was obtained, was
proposed to undergo exchange of the two N-ligands via intramolecular
nucleophilic substitution (Scheme 4- 18).
+
Using distal nitrogen ligands, it is also possible to see a 4 4 coordination
4-23, where the nitrogen ligands take up the distal coordination sphere (Scheme
4- 19). Appropriate functionalization allows the formation and observation of
+
formal silicon cations (3 + 4 coordination) and dications (2 4 coordina-

-Q
\'
/N Si-N:
. / \

8
1 4-22 2
111

2+

- \

4-23
N-Br =NBS

0
Scheme4-19

4.7. REFERENCES

1. For a review, see: Chuit, C.; Comu, R. J. P.; Reye, C.; Young, J. C. Chem. Rev. 1993,
93, 1371.
2. Selectivities in Lewis Acid Promoted Reactions, Schinzer, D., Ed., Kluwer Academic
Publishers: Dordrecht, 1989.
3. (a) Dubry, J.-L. J. Ann. Sci. Univ. Besangon 1976,29. (b) Jolibois, H. Ann. Sci. Univ.
Besangon 1976, 3.
4. (a) Murphy, M. K.; Beauchamp, J. L. J. Am. Chem. SOC.1977,99,4992. (b) Larson,
J. W.; McMahon, T. B. J. Am. Chem. SOC.1995, 107, 766.
5. For thermodynamic values for pentacoordinate Lewis acid-Lewis base formation,
see: Graddon, D. P.; Rana, B. A. J. Organomet. Chem. 1977, 136, 315.
6. For a review on Lewis acid-Lewis base complexes of silicon, see: Tandura, S . N.;
Voronkov, M. G.; Alekseev, N. V. Top. Curr: Chem. 1986, 131, 99.
7. Gel’mbol’dt, V. 0.;Ennan, A. A. Koord. Khim. 1983, 9, 579.
8. Damrauer, R.; Danahey, S. E. Organometallics 1986, 5 , 1490.
9. Wannagat, U.; Schwarz, R. Z. Anorg. Allgem. Chem. 1954, 277, 73.
10. Graddon, D. P.; Rana, B. A. J. Organomet. Chem. 1977, 140, 21.
11. Blake, A. J.; Cradock, S.; Ebsworth, E. A. V.; Franklin, K. C. Angew. Chem., Int. Ed.
Engl. 1990, 29, 76.
12. Kost, D.; Kalikhman, I. Hypervalent Silicon Compounds, In The Chemistry of
Organic Silicon Compounds, Rappoport, Z.; Apeloig, Y., Eds., Wiley: Chichester,
UK, 1998, Vol. 2, Chap. 23, p. 1339.
13. (a) Laine, R. M.; Blohowiak, K. Y.; Robinson, T. R.; Hoppe, M. L.; Nardi, P.; Kampf,
J.; Uhmn, J. Nature 1991,353, 642. (b) Kansal, P.; Laine, R. M. J. Am. Ceram. SOC.
1994, 77, 875.
14. (a) Helmer, B. J.; West, R.; Comu, R. J. P.; Poirer, M.; Royo, G.; de SaxcC, A.
J. Organomet. Chem. 1983,251,295. (b) Bassindale, A. R.; Stout, T. J. Chem. SOC.,
Chem. Commun. 1984, 1387.
15. de Keijzer, A. H. J. F.; de Kanter, F. J. J.; Schakel, M.; Schmitz, R. F.; Klumpp, G. W.
Angew. Chem., Int. Ed. Engl. 1996, 35, 1127.
n2 EXTRACOORDINATIONAT SILICON

16. For some selected examples, see: (a) Miihleisen, M.; Tacke, R. Organometallics
1994, 13, 3740. (b) Tacke, R.; Miihleisen, M. Angew. Chem. 1994, 106, 1431. (c)
Tacke, R.; Muhleisen, M. Inorg. Chem. 1994,33,4191. ( d ) Tacke, R.; Muhleisen, M.;
Jones, P. G. Angew. Chem., Int. Ed. Engl. 1994,33, 1186. (e) Muhleisen, M.; Tacke,
R. Chem. Ber. 1994,127,1615. ( f ) Tacke, R.; Becht, J.; Lopez-Mras, A.; Sperlich, J.
J. Organomet. Chem. 1993,446, 1. (g) Tacke, R.; Dannappel, 0. New Zwitterionic
Silicon-Oxygen Compounds Containing Pentacoordinate Silicon Atoms: Experi-
mental and Theoretical Studies (from a Workshop at the University of Bielefeld,
1995), Corriu, R.; Jutzi, P., Eds., Vieweg: Braunschweig, 1996, p. 75. (h) Tacke, R.;
Dannappel, 0.; Miihleisen, M. Syntheses, Structures and Properties of Molecular
X’Si-Silicates ContainingBidentate 1,2-Diolat0(2-)Ligands Derived from a-Hydro-
xycarboxylic Acids, Acetohydroximic Acid, and Oxalic Acid: New Results in the
Chemistry of PentacoordinateSilicon, In Organosilicon Chemistry: From Molecules
to Materials II (Proceedingsof the Munich Silicon Days, 1994),Auner, N.; Weiss, J.,
Eds., VCH: Weinheim, 1996, pp. 427, 447, 453 (three separate papers).
17. (a) Frye, C. L. J. Am. Chem. SOC.1970,92, 1205. See also: (b) Brook, M. A.; Chau,
D.; Roth, M. J.; Yu, W.; Penny, H. Organometallics 1994, 13, 750. (c) Roth, M. J.;
Brook, M. A.; Penny, H. B. J. Organomet. Chem. 1996, 521,65.
18. Tacke, R.; Lopez-Mras, A.; Jones, P. G. Organometallics 1994, 13, 1617.
19. (a) Pestunovich, V.; Kuppichenko, S.; Voronkov, M. Silatranes and Their Tricyclic
Analogues, In The Chemistry of Organic Silicon Compounds, Rappoport, Z.;
Apeloig, Y., Eds., Wiley: Chichester, UK, 1998, Vol. 2, Chap. 24, p. 1447. (b)
Voronkov, M. G. Pure Appl. Chem. 1966,13,35. (c) Voronkov, M. G.; Yarosh, 0.G.;
Shchukina, L. V.; Tsetlina, E. 0.; Tandura, S. N.; Korotaeva, I. M. Zh. Obshch. Khim.
1979, 49, 614.
20. Voronkov, M. G. Top. Curr. Chem. 1979, 84, 77.
21. Frye, C. L.; Vincent, G. A.; Finzel, W. A. J. Am. Chem. SOC.1971, 93, 6805.
22. Verkade, J. G. Acc. Chem. Res. 1993, 26, 483.
23. Iwamiya, J. H.; Maciel, G. E. J. Am. Chem. SOC.1993, 115, 6835.
24. Schmidt, M. W.; Windus, T. L.; Gordon, M. S. J. Am. Chem. SOC.1995, 117,7480.
25. (a) Wan, Y.; Verkade, J. G. J. Am. Chem. SOC.1995,117, 141. (b) Wan, Y.; Verkade,
J. G. Organometallics 1996, 15, 5769.
26. Day, R. 0.; Prakasha, T. K.;Holmes, R. R.; Eckert, H. Organometallics 1994, 13,
1285.
27. Gordon, M. S.; Windus, T. L.; Burggraf, L. W.; Davis, L. P. J. Am. Chem. SOC.1990,
112,7167.
28. Ref. 6, p. 119.
29. Kleboth, K.; Rode, B. M. Monatsh. Chem. 1974, 105, 815.
30. Baybutt, P. Mol. Phys. 1975, 29, 389.
31. Maerker, C.; Kapp, J.; Schleyer, P. v. R. The Nature of Organosilicon Cations and
Their Interactions, In Organosilicon Chemistry: From Molecules to Materials II
(Proceedings of the Munich Silicon Days, 1994), Auner, N.; Weiss, J., Eds., VCH:
Weinheim, 1996, p. 329.
32. Dieters, J. A.; Holmes, R. R. J. Am. Chem. SOC.1990, 112, 7197.
33. Boudin, A.; Cerveau, G.; Chuit, C.; Coniu, R. J. P.;Reye, C. Angew. Chem., Int. Ed.
Engl. 1986, 25, 473.
REFERENCES 113

34. (a) BrCfort, J.-L.; Corriu, R. J. P.; GuCrin, C.; Henner, B. J. L.; Wong Chi Man, W. W.
C. Organometallics 1990, 9,2080. (b) Corriu, R. J. P.; Guerin, C.; Henner, B. J. L.;
Wong Chi Man, W. W. C. Organometallics 1988,7,237. (c) Comu, R. Reactivity of
Penta- and Hexacoordinated Silicon Species, In Organosilicon Chemistry: From
Molecules to Materials I (Proceedings of the Munich Silicon Days, 1992),Auner, N.;
Weiss, J., Eds., VCH: Weinheim, 1994, p. 157.
35. For a review on stereochemical aspects of extracoordinate systems, see: Comu,
R. J. P.; GuCrin, C.; Moreau, J. J. E. Stereochemistry at Silicon In Topics in
Stereochemistry, Eliel, E. L.; Wilen, S. H.; Allinger, N. L., Eds., Wiley: New York,
1984, Vol. 15, p. 43.
36. Boyer, J.; Brelibre, C.; Corriu, R. J. P.; Kpoton, A.; Poirier, M.; Royo, G.
J. Organomet. Chem. 1986, 311, C39.
37. For related work, see: Brelibre, C.; Comu, R. J. P.; Royo, G.; Wong Chi Man,
W. W. C.; Zwecker, J. Organornetallics 1990, 9, 2633.
38. (a) Corriu, R. J. P.; Lanneau, G. F.; Perrot, M. Tetrahedron Lett. 1987,28, 3941. (b)
Corriu, R. J. P.; Lanneau, G. F.; Perrot, M. Tetrahedron Lett. 1988, 29, 1271.
39. Kira, M.; Sato, K.; Sakurai, H. J. Org. Chem. 1987, 52, 948.
40. Kohra, S.; Hayashida, H.; Tominaga, Y.; Hosomi, A. Tetrahedron Lett. 1988,29, 89.
41. (a) Hosomi, A.; Hayashida, H.; Kohra, S.; Tominaga, Y. J. Chem. SOC.,Chem.
Commun. 1986, 1411. (b) Comu, R. J. P.; GuCrin, C.; Henner, B.; Wang, Q.
Organometallics 1991,10,2297.(c) Becker, B.; Corriu, R. J. P.; Guerin, C.; Henner,
B.; Wang, Q. J. Organomet. Chem. 1989, 359, C33.
42. (a) Cerveau, G.; Chuit, C.; Corriu, R. J. P.; Reye, C. J. Organomet. Chem. 1987,328,
C17. (b) Kira, M.; Sato, K.; Sakurai, H. J. Am. Chem. SOC.1988, 100, 4599.
43. (a) Kira, M.; Kobayashi, M.; Sakurai, H. Tetrahedron Lett. 1987,28,4081.(b) Kira,
M.; Hino, T.; Sakurai, H. Tetrahedron Lett. 1989, 30, 1099. For a review, see: (c)
Sakurai, H. Synlett 1989, 1.
44. (a) Hayashi, T.; Matsumoto, Y.; Kiyoi, T.; Ito, Y.;Kohra, S.; Tominaga, Y.; Hosomi,
A. Tetrahedron Lett. 1988, 29, 5667. (b) Hosomi, A.; Kohra, S.; Tominaga, Y.
J. Chem. SOC., Chem. Commun. 1987, 1517.
45. (a) Kira, M.; Sato, K.; Sekimoto, K.; Gewald, R.; Sakurai, H. Chem. Lett. 1995,281.
(b) Sato, K.; Kira, M.; Sakurai, H. J. Am. Chem. SOC. 1989,111,6429.(c) Hosomi,
A.; Kohra, S.; Ogata, K.; Yanagi, T.; Tominaga, Y. J. Org. Chem. 1990, 55, 2415.
46. Kobayashi, S.; Nishio, K. J. Org. Chem. 1994, 59, 6620.
47. Kira, M.; Zhang, L. C.; Kabuto, C.; Sakurai, H. Organometallics 1996, 15, 5335.
48. Tacke, R.; Miihleisen, M. Angew. Chem., Int. Ed. Engl. 1994, 33, 1359.
49. (a) Pommier, J. C. Rev. Silicon, Germanium, Tin,Lead Cmpd. 1979,4,91. (b) Zheng,
J.-Y.; Konishi, K.; Aida, T. Chem. Lett. 1998, 453. (c) Kane, K. M.; Lorenz, C. R.;
Heilman, D. M.; Lemke, F. R. Inorg. Chem. 1998,37,669. (d) Kane, K. M.; Lemke,
F. R.; Petersen, J. L. Inorg. Chem. 1995, 34, 4085.
50. Marks, T. Science 1985, 227, 881.
51. Wolff, B.; Weiss, A. Angew. Chem., Int. Ed. Engl. 1986, 25, 162.
52. Boudin, A.; Cerveau, G.; Chuit, C.; Comu, R. J. P.; Reye, C. Organometallics 1987,
7, 1165.
53. Miiller, R. 2. Chem. 1984, 24, 41.
114 EXTRACOORDINATION AT SILICON

54. (a) Tamao, K.; Yoshida, J.-I.; Takahashi, M.; Yamamoto, H.; Kakui, T.; Matsumoto,
H.; Kurita, A.; Kumada, M. J. Am. Chem. SOC. 1978, 100, 290. (b) Yoshida, J.-I.;
Tamao, K.; Kakui, T.; Kurita, A.; Murata, M.; Yamada, K.; Kumada, M. Organo-
metallics 1982,1,369. (c) Tamao, K.; Yoshida, J.; Akita, M.; Sugihara, Y.; Iwahara,
T.; Kumada, M. Bull. Chem. SOC. Japan 1982, 55, 255.
55. (a) Tamao, K.; Yoshida, J.-I.; Yamamoto, H.; Kakui, T.; Matsumoto, H.; Takahashi,
M.; Kurita, A.; Murata, M.; Kumada, M. Organometallics 1982,1,355. (b) Tamao,
K.; Yoshida, Y.; Murata, M.; Kumada, M. J. Am. Chem. SOC.1980, 102, 3267.
56. Yoshida, J.; Tamao, K.; Kumada, M.; Kawamura, T. J. Am. Chem. SOC. 1980, 102,
3269.
57. (a) Kumada, M.; Tamao, K.; Yoshida, J.-I. J. Organomet. Chem. 1982,239, 115. (b)
Tamao, K.; Akita, M.; Kumada, M. J. Organomet. Chem. 1983,254, 13. (c) Tamao,
K.; Kakui, T.; Akita, M.; Iwahara, T.; Kanatani, R.; Yoshida, J.; Kumada, M.
Tetrahedron 1983, 39, 983.
58. (a) Maitlis, P. M. The Organic Chemistry of Palladium, Academic: New York, Vol. 2,
1971. (b) Tsuji, J. Organic Synthesis with Palladium Compounds, Springer: Berlin,
1980.
59. Yoshida, J.; Tamao, K.; Yamamoto, H.; Kakui, T.; Uchida, T.; Kumada, M.
Organometallics 1982, 1, 542.
60. Carre, F.; Chuit, C.; Corriu, R. J. P.; Fanta, A.; Medhi, A.; Reye, C. Organometallics
1995, 14, 194.
6 1. CarrB, F.; Chuit, C.; Comu, R. J. P.; Mehdi, A.; ReyC, C. Angew. Chem.,Int. Ed. Engl.
1994,33, 1097.
62. Carre, F.; Chuit, C.; Comu, R. J. P.; Mehdi, A.; Reye, C. Organometallics 1995,14,
2754.

You might also like