You are on page 1of 27

REACTION MECHANISMS

FOR NUCLEOPHILIC
SUBSTITUTION AT SILICON

Nucleophilic substitution is the most common reaction at silicon.' The early


investigations into the nature of nucleophilic substitution at silicon were
predicated on the belief that they would be similar to nucleophilic substitution at
carbon. In organic chemistry, the vast majority of nucleophilic substitutions
occur via one of two mechanisms, the s N 1 and sN2 reactions.

5.1. REACTION MECHANISMS IN CARBON CHEMISTRY

The sN1 reaction (substitution, nucleophilic, unimolecular in the rate-


determining step) can occur when non-nucleophilic, non-basic polar solvents
are used as the reaction medium. Because the rate-determining step only
involves bond cleavage, and thus has an intrinsically high barrier, the s N 1
reaction is generally observed only when the intermediate cation 5-1 is stabilized
(tertiary, allylic, benzylic, etc.). As 5-1 is a planar carbocation, the stereo-
chemical integrity of the system is lost (Scheme 5-1). Thus an sN1 reaction is
characterized by racemization (reaction of a pure enantiomer would result in a
racemic product) and a first-order rate law.
The most common reaction mechanism for substitution at a carbon atom is
the sN2 reaction (substitution, nucleophilic, bimolecular in the rate-determining
step). The reaction occurs in a single step with overall inversion of stereo-
chemistry. At the transition state, the bonds to both the leaving group and
incoming nucleophile are lengthened (Scheme 5-2).2
In principle, therefore, kinetic and/or stereochemical investigations should
allow the direct determination of the mechanism of nucleophilic substitution at
silicon, sN1 or SN2. However, several other factors must be considered. For
instance, reactions involving stable silylium ions in condensed phase are quite
uncommon, as we have seen (see Chapter 3). Thus the SN1 reaction is not
115
116 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

The S,1 Reaction: Unimolecular


Two Step

Reaction occurs with


+ LG:
Racemization

Nu: = Nucleophile
LG: Leaving Group Rate = k, l’fLG]
Scheme 5-1

The S,2 Reaction: Bimolecular


One Step

R3 L
Reaction occurs with
Stereochemical Inversion

LG: = Leaving Group


Nu: = Nucleophile Rate =

Scheme 5-2

expected to be frequently, if ever, observed. In addition, because of its greater


size than carbon, more than four ligands fit around silicon, and because of the d
orbitals (or other orbitals of similar symmetry, e.g., c*,three-center; see Chapter
23), stable five- and six-coordinate species can be prepared if the ligands on
silicon are sufficiently electron withdrawing (Chapter 4). Therefore, the
chemical behavior of extracoordinate species must also be considered when
undertaking mechanistic investigations.
Having noted that one must not limit the mechanistic considerations to the
s N 1 and sN2 reactions, it is still true that the clearest interpretation of the
mechanism will arise if: (1) the kinetics of the process are established; and (2)
enantiopure silanes are used so that the enantioselectivity of the process
(retention, inversion, racemization) can be determined.

5.2. THE STUDY OF STEREOCHEMISTRY, MECHANISM, AND SILICON:


ENANTIOPURE SILANES

Sommer and co-workers recognized the advantages of studying the mechanism


of nucleophilic substitution at silicon with chiral silanes and were the first to
THE STUDY OF STEREOCHEMISTRY. MECHANISM, AND SILICON: 117

prepare enantiopure functional silanes4 (although enantiopure tetraalkylsilanes


had been isolated in pioneering work by Kipping and co-workers5 and later by
Eaborn and Pitt6). Much work has since been done on the preparation of
enantiopure silanes.1,7-9 With such compounds, changes (or not) in stereo-
chemistry during nucleophilic substitution can be combined with kinetic data in
order to assess likely reaction mechanisms.lo
The synthesis of enantiopure methylnaphthylphenylsilanes began with the
readily available dimethoxymethylphenylsilane,which, using a Grignard reac-
tion, was converted to a methoxymethylnaphthylphenylsilaneracemic mixture.
Transesterification with menthol and separation by crystallization led to the two
pure enantiomers, 5-2 and 5-3.These could be converted by well-established
techniques into enantiopure methylnaphthylphenylsilanes (R3Si*X, X = alkyl,
aryl, halogen, hydrogen, OR, OAc, etc.; Scheme 5-3).

Scheme 5-3

Sommer and his colleagues, in a long series of seminal papers collated in a


very readable book," systematically examined the reaction outcomes of a series
of chiral silanes with different nucleophiles. Some of these data are collected in
Table 5.1. The table is divided by the nature of the leaving group: halogen,
hydrogen, carbon-based, and oxygen based.

y-y + HMgCl

5.2.1. Inversion: The SN2 Reaction


An examination of the data in Table 5.1 indicates that inversion is associated
with compounds that possess good leaving groups (groups for which the
conjugate acid has a relatively low pK, such as C1-, Br-, I-, AcO-, TsO-,
etc.). Sommer believed the mechanism of these reactions was the SN2reaction.
118 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

Table 5.1. Stereochemical Outcome of Nucleophilic Reactions at Silicon

R,sI’-x - Nu-

(or NuH)
R,S<-NU + x- (or HX)

Stereochemical
Entry Starting material Nucleophile Solvent Product outcomea
1 R3Si*-X (X=Cl, Br, I) H20 EtzO R3Si*-OH Inv
2 EtLi THF R3Si*-Et Inv
3 AcOK PhH R3Si*-0Ac Inv
4 R3Si*-F L i A l b (“H-”) EtzO R3Si*-H Inv
5 L i A l b (“H-”) C5HI2 - NR
6 MeOH - NR
7 t-BuOH R3Si*-F Rac
8 R’Li EtzO R3Si*-R’ Ret
9 R3Si*-OAce MeOH R3Si*-OMe Inv
10 L i A l h (“H-”) Et2O R3Si*-H Inv
11 KOH xylene R3Si*-OK Inv
12 R3Si*-OMe HMgCld Et2O R3Si*-H Ret
13 LiA1H4 (“H-”) EtzO R$i* -H Ret
14 KOH xylene R3Si*-OK Ret
15 MeOH/H+ MeOH R3Si*-OMe Rac
16 R3Si*OSi*R3 KOH R3Si*-OK Ret
17 R3SP-H KOH xylene R3Si*-OK Ret
18 C1Z CCl,” R3Si*-C1 Ret
19 mCPBA‘ cc14 R3Si*-OH Ret
20 R3Si*-Ac’ R ‘0- see Scheme R3Si*-OR’ Ret
5 -4
aInv = Inversion, NR = no reaction, Rac = Racemization, Ret = Retention.
bWhile the chlorination may be, in fact, a radical reaction (see Chapter 4), consider an ionic
mechanism involving nucleophilic attack of C1- at silicon and C1’ at H (5-4): Analogous studies
have shown bromination occurs via electrophilic attack rather than a radical process.”
meta-Chloroperbenzoic acid.
dSee Eq. 5.1.
Similar results were observed for tosylates (-OS02C6&CH3).

-
-
R,Si’O

“3’ R,Si
H
R,SI’

Ph
PhLi R,S*I’

Ph
Ph
0-
Ph-C-
\

1
RETENTION

Ph
54
Brook Rearrangement

Scheme 5-4

Recall, however, that silicon can much more easily expand its coordination shell
than carbon, and, as a consequence of low-lying d or three-center orbital^,^
stable pentacoordinate species can be isolated. Thus the inversion process need
not be a true SN2 reaction: The pentacoordinate species 5-5 can be a transition
state or an intermediate (Scheme 5-5).
THE STUDY OF STEREOCHEMISTRY, MECHANISM, AND SILICON: 119

The racemization result for MeOH with acidic catalysis was similarly thought
to follow the SN2-type mechanism. However, as the leaving group and
nucleophile are the same species, the reaction is actually an equilibration that
results in loss of optical purity (Scheme 5-6).12
goo

,R
3
R'
i ,R2
Si
I
X
- Nu-
NU7

R3-~7..*R2~,200

x
I -R
5-5
,
- -X-
R3/
Nu
I
Si: \ 2
i
R'
R

Inversion
X = CI.Br, I, OTs, 0
K R

Scheme 5-5

Me, +,H
Me Me 0 H\ +M
,e OMe
Naph\ f P ,h Naph\ APh I -.+,Me H+
R Si & Rsi Naph-Sibph 1s Si
I H'
I+ '*
Naph'
Si
p
i' h
Naph' i'ph
Me
OMe H/O\Me Me Me

Symmetrical Intermediate

Scheme56

5.2.2. Retention: The SNi Mechanism


The remaining data (Table 5.1) show that retention is a common stereochemical
outcome of nucleophilic reactions at silicon. It is observed when leaving groups
are involved that form very strong bonds to silicon (F, OR) or are not highly
polarized (H, C). In carbon chemistry, retention would normally imply a double
inversion process (sequential sN2 reactions). However, there is no evidence to
suggest this is a common mechanism in the case of silicon. The alternative
mechanism for carbon, sN1, is accompanied by racemization, as noted above
(Scheme 5-1). However, given that silylium ions (R3Si+) are exceptionally rare
in condensed phases (see Chapter 3), further consideration will not be given to
the s N 1 reaction mechanism for silicon.
Sommer proposed a mechanism not commonly observed with carbon to
account for retention, the SNireaction (substitution nucleophilic internal13).This
involves dissociation in an SN1-typereaction followed by same side attack of the
nucleophile (for instance, in the conversion of ROH + ROSOCl 4 RC1 with
retention) or edge-on attack by the incoming nucleophile. For example,
nucleophilic attack into an equatorial position (Scheme 5-7A), or more likely via
a four-centered intermediate 5-6, stabilized by the coordination to the gegenion
E of both the incoming nucleophile and outgoing leaving group (Scheme 5-7B),
leads to same side substitution (Sommer also proposes a square pyramidal
mechanism for this pro~ess'~). The SNi mechanism was proposed to be more
120 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

viable for silicon than carbon because: (1) The longer bond lengths to silicon,
compared to carbon, leave more room for the same side approach of the
nucleophile; and (2) the presence of low-lying orbitals on silicon allows the
formation of stable extracoordinate intermediates (however, see below for an
alternative e~planation).~ Some support for the intermediacy of a four-
membered ring comes from the study of metalated aminofluorosilanes, many
of which aggregate in four-ring patterns (Chart 5-1).15

R' - R2 7'
R3\ i R,2
Si
Y
* R3-SSj-..y
I
.
h R' R3\ Ai / R2
+ x-
A I
X X
I
Y
Retention

R' R2 R'

X 5-6 X--.
I
Y
Retention

R,Si CI R Si _______ CI R3S!- Cl R,Si -CI


e.g. I + I
H ______ CI
I I - : : -
8
+
,

H CI I ,

H - CI H-CI

Scheme 5-7

-Si
1 '-: N-
Li- F
I
-Si-N B/Li,N,R
F
I
-Si-N
,R
I l l 1 I I I -Si-N I I
F- Li- N- Si - F-Li Si- I I . F -Li-THF
I 'F' \ I
RI t TH F TH F

Chart 5-1

The SNi mechanism would be expected to be important when the leaving


group is not well solvated (Table 5.1, entry 5 versus entry 4),when the silicon-
leaving group bond is strong (Table 5.1, entries 8, 12-16) or when the silicon-
leaving group bond is not highly polarized (Table 5.1, entries 17-20). In all these
cases, additional activation to nucleophilic substitution can be provided by the
coordination to the gegenion E."
More recent investigations have suggested that the actual mechanism of
nucleophilic substitution is somewhat more complex than the vision enunciated
by Sommer. Although he presents a cogent argument for the sN2 reaction at
silicon with good leaving groups, Sommer discounts the two-step process
involving formation of a pentacoordinate intermediate followed by its fast
decomposition (Eq. 5.2).l 6 As will be discussed below, this mechanism has some
merit in light of more recent investigations. Similarly, the reasonable SNi
proposal is inconsistent with some mechanistic results that were later obtained.
THE ROLE OF EXTRACOORDINATE SILICON IN NUCLEOPHILIC SUBSTITUTION 121

In the next section, we shall outline some of these investigations from an


inorganic perspective.

5.3. THE ROLE OF EXTRACOORDINATE SILICON IN


NUCLEOPHILIC SUBSTITUTION

We have seen (see Chapter 4) that stable extracoordinate silicon compounds are
relatively easy to prepare. Extracoordinate structures are also implicated as
intermediates in nucleophilic substitution reactions irrespective of the stereo-
chemical outcome. Bassindale and co-workers traversed between the two
extreme resonance structures of 2-pyridone, 5-7 and 5-8, to probe the nature of
these pentacoordinate intermediate^'^: Other amides also form extracoordinate
structures,l 8 as do electron-deficient silyl esters.l9 Which of the three possible
compounds 5-9,5-10, and 5-11 is present in solution can be established from the
13CNMR (which discriminates between the aromatic and amide structures) and
29Si NMR spectra (which establishes the degree of extracoordination at
silicon).20The system was studied by varying the leaving group on silicon from
excellent CF3S03- (TfO-) to poor (-OR), and by varying the nucleophilicity of
the oxygen nucleophile using the electronic characteristics of the ring group R
(Scheme 5-8).

I 5-7
0
- 0- I 5-80

5-9 I SiMe,X
R&
5-10
N O&
SiMe,X
f? o x-
5-11 LiMe,

Scheme54

It was shown that as R becomes more electron supplying, and the oxygen
consequently more nucleophilic, the equilibria shift to the right. The equilibria
also shift to the right as the leaving group ability increases X = CF3S03> Br >
C1> OR. Discreet examples of 5-9, 5-10, and 5-11 could be observed with an
appropriate choice of R. These results suggest that the general mechanism of
nucleophilic substitution at silicon involves pentacoordinate intermediates or
transition states.
122 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

5.4. SILICON AS AN INORGANIC ELEMENT (RATHER THAN AS A CARBON


ANALOGUE): UNDERSTANDING THE STEREOCHEMICAL OUTCOME

The well-studied chemistry of phosphorous, the adjacent element on the periodic


table, serves as a useful model for the behavior of extracoordinate silicon.
Inorganic chemistry teaches that, in an sp3d hybridized, trigonal bipyramidal,
pentacoordinate system, bonding to the equatorial ligands is normally different
from the axial (apical) ligands.21*22 In the case of PC15, for instance, the axial
P-C1 bonds are 14% longer than the equatorial bonds (1.20 versus 1.05w).23
The sp3d pentacoordinate system can be considered to be a composite of an
equatorial sp2 system plus an axial pd hybrid.24 The sp2hybrid forms stronger
and shorter bonds than the axial pd hybrid. VSEPR (Valence Shell Electron Pair
Repulsion) theory associates the difference in bond lengths with the optimization
of space occupied by the five ligands: There is more space in the equatorial
positions (2 x 120°, 1 x 90” separation) than in the axial positions (3 x 90”
~eparation).~’
In mixed ligand systems, it is observed that electronegative groups prefer to
occupy the apical sites. This observation known as “Bent’s rule.”26It has been
suggested that strong s orbital-rich covalent bonds require a larger volume in
which to bond than more highly electronegative substituents, which therefore
take the p orbital-rich axial positions.
Nucleophilic substitution at a tetrahedral (or square planar) metal complex
can be considered with these concepts in mind. Addition of a nucleophile leads
to coordination expansion, generally (with main group elements) resulting in a
trigonal bipyramidal system. The usually electronegative nucleophile and
leaving group will prefer to occupy the apical positions. Note that the
apicophilicit of a given leaving group is affected by the nature of the incoming
nucleophile! Subsequent loss of the leaving group from the apical position,
which has a longer, weaker bond than the equatorial position, leads to
regeneration of the tetrahedral configuration with inversion.28 As the
apicophilicity of a nucleophile increases, its position in the equatorial position
is destabilized. This tends to favor inversion even with relatively poor leaving
groups.29This mechanism for nucleophilic substitution is indistinguishablefrom
the SN2 reaction in carbon chemistry except that the pentacoordinate species
may be a transition state (like the organic case) or a more stable reaction
intermediate (Scheme 5-5). This mechanism therefore provides an alternative
explanation for the nucleophilic substitutions at silicon that occur with inversion
(Table 5.1). It does not account for the retention mechanism.

5.4.1. Pseudorotation
The minimum repulsion valence-shell interactions between five identical ligands
located around the central (ten-electron) silicon atom corresponds to the trigonal
bipyramid (tbp), in which the two axial bonds are longer than the three
equatorial, as The tbp structure is highly fluxional3’: At normal
SILICON AS AN INORGANIC ELEMENT 123

temperatures, for example, it is not possible to distinguish the axial and


equatorial fluorides in SiF5- .32 The most commonly invoked mechanism to
explain the stereodynamic r o c e ~ s e associated
s~~ with pentacoordinate species34
is Berry pseudorotati~n.~~~ B
By this process, ligand exchange between axial and
equatorial ligands can take place. Two pseudorotations are relevant to our
discussion. Low-energy barriers [25-50 kJ mol-' (6-12 kcal mol- 1)37 or less]38
exist between the tbp and square pyramid (spy) isomers. Interconversion
between the two leads to an exchange of axial and equatorial ligands (Scheme
5-9). The second, higher-energy pathway is analogous to the motion of a
turnstile: Two ligands are rotated in one direction, while the other three rotate in
the opposite direction (5-12 5-13 or 5-14 t 5-15, Scheme 5-10).
--f

SPY
Scheme5-9

180' 5-12 5-13


Rotation
- -
Y R3
..,
R 3 - SIi L R
x~ ] Pseudo I
Y - l < R 2 ]R'

I Rotation
R2

5-14 5-15
Scheme 5 -10

A viable mechanism for nucleophilic substitution at silicon with retention


involves pseudorotation. In the case of non-electronegative leaving groups Y,
there is no strong driving force for them to assume the apical position in a five-
coordinate intermediate 5-16. Therefore, upon coordination expansion, the
nucleophile and another group than the leaving group may occupy the apical
positions 5-17. Through pseudorotation, the nucleophile and leaving group can
swap positions on the trigonal bipyramid, giving 5-18. The leaving group can
now leave from the apical position, but from the sume side of the molecule from
124 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

which the nucleophile attacked, leading to a product formed with retention of the
original configuration (Scheme 5-11).

R<:
R’
Si
I
Y
CR2
-
NU-
Nu

Y has low electronegativity

R‘
Retention
R3’,
s’h R 2
Nu

Scheme 5-11

Several researchers have examined the question of pseudorotation at silicon.


Comu and co-workers have prepared a large number of derivatives that are
pentacoordinate by virtue of a distal coordinating nitrogen ligand, as exemplified
by 5-19 (Chart 5-2).39*40Martin and co-workers have examined the pseudorota-
tion of pentacoordinate anions, which have five o bonds to silicon, 5-20 (Chart
5-2).41-44 Pseudorotation can be followed by NMR because: (1) the apical
and equatorial ligands have different chemical shifts in the NMR; and (2) the
exchange process, between apical and equatorial ligands, frequently occurs at a
rate compatible with the NMR time scale.
Y

Chart 5-2

In the pseudorotation process, equatorial and axial ligands on a trigonal


bipyramid equilibrate. However, to establish pseudorotation as operative in
nucleophilic substitutions, one must discount a dissociative process through
which ligands can also exchange positions (dissociation, 5-21 -+ 5-22;
pseudorotation, 5-21 -+ 5-23, Scheme 5-12). A detailed discussion is provided
for 5-21, which is generally applicable.
During a pseudorotation process, the silicon atom remains bonded to all five
substituents, whereas the Si-N bond cleaves in a dissociation process. After
cleavage 5-21 -+ 5-22: (1) the diastereotopic N-Me groups become equivalent
in the I3C NMR through pyramidalization at nitrogen; and (2) simultaneously
the three Si-F groups, which lose their apical / equatorial position, become
SILICON AS AN INORGANIC ELEMENT 125

qk:,,~ ..+?
F F
I .+ F"F' F1
I

q-i,,, Mea
Meb ~ - ; . . , U I M
' ea Meb

dissociation pseudorotation
5-22
AGc = 49.4 kJ rnol-1
5-21
c:
AG = 39.3 kJ rnol-1
5-23

= 11.8 kcal rnol-I = 9.4 kcal mol-1


Scheme 5-12

equivalent tetrahedral substituents in the 19F NMR. In a pseudorotation process


5-21 + 5-23, by contrast, the three Si-F groups should become equivalent by
exchange in the 19F NMR, while the N-Me groups remain distinguishable/
diastereotopic.
It was shown with silanes of type 5-21 that both dissociation and pseudo-
rotation processes occur: Each process has a different energy of activation. For
this compound, the pseudorotation process was found to have a lower activation
energy than dissociation. By contrast, Martin and co-workers have shown that
anionic siliconates 5-20 undergo pseudorotationexclusively (Y = n-Bu, Ar,OAr,
F)!l The barrier to pseudorotation was observed to drop with increased
electronegativity of the group Y: During pseudorotation, the Y group moves to
the apical position, which puts more negative charge on Y (the barrier decreases
as Y can accept more electron density). The AG# values for pseudorotation and
dissociation, for a variety of extracoordinate silanes, are shown in Table 5.2.
As a consequence of these many studies, it is possible to generalize the ease
with which pentacoordinate silicon compounds undergo pseudorotation. The
barriers to pseudorotation have been shown to depend on the number of
electronegative substituents (Chart 5-3,functional groups, X = halogen, alkoxy,
amine, etc.). The most stable of the pseudorotamers have the electronegative
groups in the apical position 5-24 < 5-25 M 5-26 (Scheme 5-13).
AGf > 84 kJ rnol-l AG# = 50 kJ rno1-l AG# < 29 kJ mol-'
(>20 kcal mol-l) (=9-12 kcal rnol-I) (< 7 kcal mot-')
Chart 5-3

R R R

R-Si,
I
I
.-.,%$ X

Y
- pseudo

rotation
- R-Si,
I -,.(. R
I X
~
pseudo

rotation
- R-Si,
I-..,- R
I Y
R Y X
5-24 5-25 5-26

Scheme 5-13

5.4.2. Factors Affecting the Stereochemistry of Substitution


These studies on pseudorotation allow the development of a generalized reaction
model for nucleophillc substitution at silicon, extended from the proposals of
126 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

Table 5.2. Stereochemlcal Data for the Substitution of 1-NaphPhMeSIX as a Function of the Leaving
Group X

AG# AG+ AG# AG#


H mol-'H mol-' H mol - H mol - ' '
(kcal mol- ') (kcal mol-') (kcalmol-') (kcalmol-')
Compound pseudorotation dissociation Compound pseudorotation dissociation

N+Sl -R'
$ 1 .'\
R3 RZ

R'R2R3 R'R2R3
H2,1-Naph >42 (10) 41 (9.7)45 PhMeH 92 (22) >92 ( 2 2 p
H3 << 29 (7) 42 (10.5)45 PhMeCl 84 (20) > 84 (20p9
PhHCl >92 (22) > 92 ( 2 2 ~ 9
PhMeOMe 92 (22) > 92

*..'I d\
R3 Rz

R'R2R3 PhMeF 96 (23) > 96 (> 23)39


H2Ph >42 (10) 43 (10.2)45 PhC12 46 (1 1139
H
3 << 29 (7) 44 (10.5)45 Ph(OMe):! 38 (9)39
MeC12 38 (9)39
<29 (7p9

Y F3 50 (12p9
n-Bu 120 (28.6)41
Ph2SiF3- 49 (11.7)&
PhMeSiF3- 41 (9.9)46
PhSiF4- < 29 (7)46

Adapted with permission from Ref. 36. Copyright John Wiley and Sons.
SILICON AS AN INORGANIC ELEMENT 127

Sommer. In particular, pseudorotation allows an explanation of substitution with


retention. The factors that control the observed nucleophilic reactivity patterns
are presented below.
Polarizability of the Leaving Group and Nucleophile: Better leaving groups,
those with conjugate acids of low pK,, are typically replaced with inversion. This
is a consequence of apicophilicity, which parallels electronegativity. Fluoride,
however, does not always undergo substitution with inversion. Thus, to the effect
of electronegativity, must be added leaving group polarizability: More
polarizable leaving groups/nucleophiles are also more a p i c ~ p h i l i cand
~ ~ thus
favor inversion in the substitution process7 (OR > C,Z > H, Chart 5-4). The less
polarizable or harder the leaving groups/nucleophile (the more charge is
concentrated), the more retention is observed in the substitution process. H
ligands, which are among the least polarizable ligands and generally undergo
substitution with retention, have been shown to prefer the equatorial position,
consistent with this pr0posal.4~
Leaving group ability from silicon
Stereochemical outcome
-
Inversion -+
-
Br CI > SR F > OMe > H
Retention
Chart5-4

The effect of polarizability extends to the nucleophile as well. As the polari-


zability is increased, the tendency for substitution by inversion also increase^.^'
For example, while an alkyllithium such as BuLi substitutes with retention, the
conjugated anion PhCH2Li substitutes with inversion (Scheme 5-14).51

R,SiR’
Retention
-R’ = alkyl
R,SiX + R‘ Li
R’ = CH$r
R,SiCH$r
Inversion
Scheme 5-14

The Nature of Metal Counterions: The choice of metal counterion can affect
the substitution process by changing the polarizability of both nucleophile and
electrophile. Thus the effect of ligand polarizability can be amplified with soft,
polarizable metals, which can interact with the leaving group / nucleophile,
increasing the apicophilicity and favoring inversion (Chart 5-5). One example
that demonstrates this point is the reaction shown in Scheme 5-15. Substitution
with a nucleophile bound to a hard counterion, Li, leads to substitution with
retention. However, the addition of a softer metal to the reaction mixture
(MgBr2) leads to a changeover in the stereochemical outcome (Scheme 5-15).52
The importance of the leaving group and counterion on the stereochemistry of
substitution with Sommer’s silanes may be deduced from the data in Tables 5.3
and 5.4.
Li’ > Na’ > K+
Harder Softer
Retention Favored Inversion Favored
Chart 5-5: Nucleophile Counterions
-
128 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

R'Li + R,SiX R,SiR' Retention

R'Li + -MgBr,

Scheme 5-15
R,SiR' Inversion

Table 5.3. Stereochemical Data lor the Substitution of 1-NaphPhMeSiXas a Function 01the Leaving
Group X
Naph Naph
Ph\,
Si
*Me Nu_ Ph\,
Si
,Me
+ X-
I I
X Nu

Predominant stereochemistry for nucleophile


Leaving group (X) in
1-NaphPhMeSiX KOR RLi RMgBr LiAlh

Adapted with permission from Ref. 36. Copyright John Wiley and Sons.
a a-NaphPh(CH,=CH)SiX was used.

Table 5.4. Stereochemical Data lor the Substitution of 1-NaphPhR'SiX as a Function of the Nucleo-
phile

Predominant stereochemistry for leaving group'

1-NaphPhR'SiOR 1-NaphPhR'SiF 1-NaphPhR' SiSR'


C6H5CH2Li 1~~58.63 1~~57.58 1~~54.65

C6H5CH2MgBr I~~~~ I~~~~ 1 ~ ~ 5 ~


LiAlh ~~~55.56~62 1~~54~65
RLi Ret51,54,63 Ret51,57,58 ~ ~ ~ 5 4 . 6 5

R = alkyl
Adapted with permission from Ref. 36. Copyright John Wiley and Sons.

Solvent Effects: As the solvent polarity increases and/or hydrogen bonding or


metal chelation is facilitated 5-27 (Chart 5-6), inversion is favored over retention
(Table 5.1, entries 4 and 5).67 Interactions between the leaving group and
solvent / gegenions/hydrogen bonding molecules increase the leaving group
ability and apicophilicity of the group through electronic assistance (Table 5.1,
entry 15).
SILICON AS AN INORGANIC ELEMENT 129

Effect of Silune Structure: Constraining the silane in a ring can have a


dramatic effect on the reactivity. Two different cases are presented below:
endocyclic and exocyclic leaving groups. The following discussion focuses on
four-membered rings: The same trend, but with lower selectivity, is associated
with five-membered rings.
Organic rings containing fewer than six members, particularly four-
membered rings, have considerable ring strain. This is a consequence of the
hybridization, which favors C-Si-C bond angles of 109", and the ring
geometry, which constrains the angle to 90". Any atomic motion or rehybri-
dization that attempts to increase the bond angle is energetically expensive, as it
induces more ring strain. By contrast, motion / hybridization leading to a smaller
bond angle is energetically favored.
For silanes with exocyclic leaving groups, inversion is the increasingly
favored stereochemical outcome of nucleophilic substitution reactions with
increasing ring size (Chart 5-7). The reasons for this trend may be understood
when looking at the reaction pathways for retention and inversion, respectively
(Scheme 5-16). The inversion process requires rehybridization of the Si atom in

Retention * Inversion
Chart 5-7

Nu I
- Nu-
higher
barrier
- Nu-
lower
barrier Nu
5-28 5-29

Inversion Retention
Scheme 5-16
130 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

the ring, with a concomitant increase in ring strain as the ring tries to
accommodate the 120" angle of the two equatorial positions; both the leaving
group and incoming nucleophile are in apical positions 5-28. By contrast, the
retention pathway allows the possibility of coordination expansion, with the ring
occupying one axial and one equatorial position 5-29. In this case, both the
angles in the ring and around the silicon are geometrically optimized at 90",
relieving ring strain. The lower barrier to the transition state favors this pathway.
These trends are superimposed on those already mentioned for nucleophilic
substitution, as may be seen from the data in Table 5.5.
The activation to nucleophilic substitution, provided by a four membered
ring, has been independently utilized in organic synthesis by Myers and co-
w o r k e r ~and
~ ~Denmark et Nucleophilic attack by an aldehyde on the silyl
enol ether (see Chapter 8), to form a pentacoordinate intermediate, relieves ring
strain (Scheme 5-17). The pentacoordinate species then undergoes a Cope
rearrangement-type aldol reaction to give the silylated aldol product without the
need for Lewis acids or other catalysts such asjluoride. Similarly, the allylation
of aldehydes can be performed diastereoselectively without the usually needed
assistance of Lewis acids or fluoride (see Chapter 16).76

Table 5.5. Stereochemical Outcome of Nucieophilic Substitution as a Function of Ring Size

Compound Angle X=Cl X=F

ii-X

4 90" Ret

93-96" Inv

105" Inv

109" Inv

Reproduced with permission from Ref. 36. Copyright John Wiley and Sons.

Scheme 5-17
SILICON AS AN INORGANIC ELEMENT 131

If the leaving group is endocyclic in a silacyclobutane,by contrast, ring strain


favors nucleophilic substitutionwith inversion. In the case of the four-membered
ring 5-30, the pentacoordinate species 5-31 leading to inversion will be
energetically more accessible than that for retention 5-32, for the reasons
enunciated above (Chart 5-8). The stereochemical reaction outcomes for silanes
with endocyclic leaving groups are shown in Table 5.6.

\ I \ /

Inversion
I 5-30

Chart 5-8
*
0
Retention

5.4.3. Why Not the SNi Mechanism?


We have seen that fluoride is a leaving group that can undergo substitution with
both inversion and retention. Displacement of this leaving group thus serves as a
useful test of mechanistic proposals.
More nucleophilic solvents DME (MeOCH2CH20Me)>THF > Et20increase
the substitution reaction rate with Grignard rea ents. The addition of chelating
agents (e.g., TMEDA, Me2NCH2CH2NMe2)52v8'was similarly found to accele-
rate the reaction. More important for the mechanistic interpretation is that polar
solvents and cryptands also increase the degree of selectivity for retention (Table
5.7).52*59Alkyllithium reagents substitute fluorosilanes primarily with retention
in ether solvents where the metal is less aggregated or, in the case of cryptands,
sequestered. In contrast, non-polar solvents such as benzene, in which the
alkyllithium species are more aggregated, favor inversion in these reactions.

Table 5.6. Stereochemical Outcome of Substitution with Various Nucleophlles on Cyclic Silicon
Compounds with the LeavingGroup Endocyclic

Nucleophile
Substrate LiAlH4 p-AnCH2Li H2C =CHCH2Li

Ret Inv

Ret Ret

Ph,
a-Naph
,\Si
,O,
Ret Inv

Adapted with permission from Ref. 36. Copyright John Wiley and Sons.
132 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

Table 5.7. Stereochemistryof Substitution as a Function of Solvent

Solvent Selectivity Stereochemistry


EtzO 265 Inv
THF 66 Inv
DME 59 Ret
Adapted with permission from Ref. 33. Copyright John Wiley and Sons.

Corriu interpreted these results to discount the SNimechanism: If the metal is


coordinated by the solvent / a cryptand, electrophilic assistance by the metal,
which is necessary to facilitate the departure of the leaving group from the
equatorial position, is not available. Thus, if the SNimechanism were operating,
the presence of chelators would be expected to lead to a changeover from
retention to inversion, not to favor retention, as is observed. It should be noted
that this remains an issue of some controversy. For a more detailed discussion of
these points, the reader is referred to Refs. 33 and 36.

5.4.4. Retention versus Inversion


Many factors control the steric outcome of nucleophilic substitution at silicon.
We have generalized the reaction conditions that favor retention and inversion
above. However, a caveat to these typical results must be made. There are many
subtleties that can make the mechanistic interpretation very complex. As an
example, Bassindale and co-workers examined the nucleophilic exchange
reactions of pentacoordinate silanes with N-methylimidazole (NMI) as both
leaving group, NMI, and nucleophile, NMI* (Scheme 5-18).81 In this system,
both retention and inversion, associative (hexacoordinate) and dissociative
(tetracoordinate) processes were proposed to be occurring.
The inversion process was shown to be zero order in the nucleophile, NMI.
This implies that the rate-determining step is dissociation of 5-33 to give the
tetracoordinate intermediate 5-34: A high positive entropy of activation for the
process was determined, 64J K-' mol-' (15 cal K-' mol-'). By contrast, the
retention process (more than one may be occurring) was postulated to involve
association of NMI to give a hexacoordinate intermediate 5-35 prior to
dissociation to the product of retention 5-36.
THE MECHANISM OF SUBSTITUTION IN THE PRESENCE OF SllAPHlLlC CATALYSTS 133

I
5-35

Ph Ph
\

5-36
Inversion Retention

Scheme 5-18

5.5. THE MECHANISM OF SUBSTITUTION IN THE PRESENCE OF


SlLAPHlLlC CATALYSTS

In Chapter 4, the notion was presented that pentacoordinate siliconates: (1) are
more reactive towards nucleophilic attack at silicon; and (2) possess more
nucleophilic ligands than their tetracoordinate counterparts. Comu and co-
workers,82in particular, have addressed the validity of this idea by examining the
effect of coordination number at silicon on reactivity towards n u c l e ~ p h i l e s . ~ ~ * ~ ~
It was shown by his groupg4 among others,85including Allen et aLg6and Frye
and c o - w ~ r k e r s ,that
~ ~ nucleophilic substitution occurs more rapidly in the
presence of catalytic amounts of appropriate “silaphiles” or “silicophiles.”
These are molecules that have a special affinity for silicon, usually possessing
electron-rich oxygen or nitrogen ligands,88 such as: hexamethylphosphoric
triamide [(Me2N)3P=0, HMPA], Ph3P0, RC02-, N,N-dimethylaminopyridine
(DMAP),87imidazole, N-methylimidazole (NMI), N,N-dimethylpropyleneurea
(DMPU), N,N-dimethylethyleneurea(DMEU), N-methylpyrrolidinone,(NMP),
1-methyl-2-pyridone (NMPO), dimethylformamide (DMF),89 and dimethyl
sulfoxide (DMS0).82 One of the most efficacious catalysts of this type is
fluoride, which, as already noted (Chapter 2), has special affinity for silicon.
The rate law for nucleophilic substitution reaction is affected by the presence
of these silaphilic catalysts. For instance, the hydrolysis and alcoholysis of
chlorosilanes in aprotic solvents was found to follow a third-order rate law: The
weak silaphilic catalyst C1- is present in the rate law (Eq. 5.3).86The normal
substitution rate law for the SN2 reaction is second order (Eq. 5.2). Other
reactions have also been shown to be first order in added silaphilic catalyst
(Table 5.8).82*86*87The addition of silaphilic nucleophile catalysts has an
134 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

important additional consequence. The stereochemical outcome of the nucleo-


philic substitution often changes from inversion to retention in the presence of
such nucleophiles,** although this depends on the facility with which the
extracoordinate intermediates undergo pseudorotation. Silicophilic nucleophiles
can also lead to racemization of organosilanes, although the order of the
nucleophile in the rate law for racemization is higher than one in these cases
(Table 5.8).90,91

rate =k [R3SiCl][ROH][Cl-] (5.3)


The mechanism most consistent with these observations involves fast addition
of the silicophilic nucleophile to the tetracoordinate species to give 5-37. This
pentacoordinate species is much more reactive than the tetracoordinate starting
material. Slow addition of the actual nucleophile to 5-37 leads to the
hexacoordinate intermediate 5-38. Collapse of the intermediate via loss of the
leaving group and silaphilic catalyst leads to the overall substitution process with
retention (Scheme 5-19).
Alternatively, 5-37 may undergo attack by a second nucleophile to give 5-39.
Decomposition of the complex, because of the plane of symmetry, may take
place with inversion or retention, leading to racemization.
The effectiveness of catalysts that increase the rate of nucleophilic
substitution at silicon was found to follow the order shown in Chart 5-9.93
This order follows the order of nucleophilicity, to the degree that the Taft p-
constant is an indicator of nu~leophilicity.~~ The utilization of these compounds
to affect otherwise slow reactions is an important tool in organosilicon chem-
istry. For example, the whole field of silane protecting groups (see Chapter 8),

Table 5.8. Effect of Silaphilic Nucleophiles on Substitution and Racemization


~

Nucleophile
Substrate Nucleophile Concentration (M) Solvent order
Substitution by H 2 0
Ph3SiC1 DMSO 0.3-1.4 x anisole 0.7882
Ph3SiC1 HMPA 0.5-3 x 1 0 - ~ anisole 1.0582
Ph3SiC1 HMPA 0.5-2.0 x THF 0.81 82

Racemization
-

5
45.
Si4x

fix
HMPA 2 x 1 0 - ~ - 2 x 1 0 - ~ CCL 1.249'

cr
t-Bu-a-NaphPhSiBr
HMPA

HMPA
3 x 10-4-5 x 1 0 - ~ CCL

2 x 1 0 - ~ - 2 x 1 0 - ~ cc14
1.S9'

2.1g9*
(PhCHMe)SiMeZCl NMI 0.1-0.6 silane 3.37"
Adapted with permission from Ref. 36. Copyright John Wiley and Sons.
THE MECHANISM OF SUBSTITUTION IN THE PRESENCE OF SlLAPHlLlC CATALYSTS 135

intermediatesor
transition states

Nu- 11 Slow

X X

-
R2
I
R'wS~,
N IiR3
R' o,,.. ST- Inversion
*
R?X~
.,~,I R3

1 R3 Fast R2' I I
R2 NU 5-37 Nu
N
; silaphilic nucleophile
Nu- 11
Slow
Racernization
4-

Nu
R'o..,,,
R2'(
; :s
I .JX
R
-
NU 5-39 Nu
Scheme 5-19

DMAP HMPA

DMPU
Chart5-9

initially described b Pierce,95 but made practical by Corey and co-workers


(with t-BuMe$iCl)!6 depends upon the use of imidazole in the solvent DMF
(which is also silicophilic) to effect the alcoholysis of hindered chlorosilanes:
Simple bases such as triethylamine are not practical for this role (Scheme 5-20).
5.5.1. Fluoride-Catalyzed Substitutions
Fluoride is frequently used to catalyze nucleophilic substitution at silicon. For
example, allylsilanes (Chapter 16), silyl enol ethers, and alkoxysilanes (Chapter
8) are effectively activated to reaction with electrophiles in the presence of
136 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

Exceptionally
Slow Reaction
NEt,
p:
THF

Scheme 5-20

nucleophiles such as fluoride. Presumably, this mechanism is similar to that


shown for hydrolysis in Scheme 5-21. Intuitively, given the strength of the Si-F
bond, such a proposal seems surprising: Once formed, one would anticipate that
it would be difficult to break the Si-F bond in the collapse of the extracoordinate
intermediate 5-40-5-41. It is indeed likely that some of these fluoro-
substituted (R3SiF) species form. However, these reactions must be in
equilibrium, as shown in Scheme 5-19, and the thermodynamics of the overall
system favor conversion of Si-OR + Si-OH. It was noted in Chapter 2 that in
spite of the strong Si-F bond, fluorosilanes are kinetically active. This is one
example of many in which fluoride catalyzes the nucleophilic substitution of
silicon, presumably by forming a pentacoordinate inte~mediate.~~

5-40 5-41
Scheme 5-21

5.5.7.7. Oxidation of Silicon-Carbon Bonds The processes described above


are further exemplified by the fluoride-catalyzed oxidation of silanes by
hydrogen peroxide. Tamao et in particular, have shown that the oxidation is
accompanied by retention of stereochemistry at both c a r b ~ n ~and ~ . silicon.
'~ lo
The mechanism is proposed to follow the typical expansion of coordination at
silicon, initiated by fluoride, followed by intramolecular transfer of an Si-alkyl
group from silicon to oxygen (Scheme 5-22). More examples of these reactions
may be found in Chapter 16 (Oxidative Cleavage of Silyl Groups).

Scheme 5-22
REFERENCES 137

5.6. REFERENCES

1. For a review on stereochemicalaspects of nucleophilic substitution, extracoordinate


systems and stereochemistry, see: Corriu, 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.
2. March, J. Advanced Organic Chemistry, 4~ Ed., Wiley: New York, 1992,
pp. 293-305.
3. For the purposes of this discussion, the relative merits of the different orbital
pictures for extracoordinate silicon will be ignored: d , three-center, and o* orbitals
will be considered simultaneously (see also Chapter 2).
4. Sommer, L. H.; Frye, C. L. J. Am. Chem. SOC.1959, 81, 1013.
5. (a) Kipping, F. S. J. Chem. SOC. 1907, 209. (b) Kipping, F. S. J. Chem. SOC.1908,
457. (c) Luff, B. D. W.; Kipping, F. S. J. Chem. SOC. 1908,2091. (d) Challenger, F.;
Kipping, F. S. J. Chem. SOC. 1910, 755.
6. Eaborn, C.; Pitt, C. Chem. Ind. 1958, 830.
7. (a) Corriu, R. J. P.; Guerin, C. J. Organomet. Chem. 1980, 198, 231. (b) Corriu,
R. J. P.; Guerin, C. Adv. Organomet. Chem. 1982,20, 265.
8. For a review, see: Maryanoff, C. A.; Maryanoff, B. E. Synthesis and Utilization of
Compounds with Chiral Silicon Centers, In Asymmetric Synthesis, Momson, J. D.,
Ed., Academic: New York, 1984, Vol. 4, p. 355.
9. Enzymatic methods can be used (a) Djerourou, A.-H.; Blanco, L. Tetrahedron Lett.
1991,32,6325. (b) Fukui, T.; Kawamoto, T.; Tanaka, A. Tetrahedron:Asymmetry
1994, 5, 73.
10. The reader is referred to the excellent book by Sommer for a clear summary of the
preparation and study of reaction mechanisms of chiral silanes: Sommer, L.
Stereochemistry, Mechanism & Silicon, McGraw Hill, 1965.
11. MareS, F.; Hetflejg, J.; Chvalovslj, V. Coll. Czech. Chem. Commun. 1970,352831.
12. Ref. 10. pp. 84-91, 97-100.
13. Ref. 2, pp. 326-327.
14. Ref. 10, p. 63.
15. Pieper, U.; Walter, S.; Klingebiel, U.; Stalke, D. Angew. Chem., Int. Ed. Engl. 1990,
29, 209.
16. Ref. 10, pp. 91-93.
17. (a) Bassindale, A. R.; Borbaruah, M. J. Chem. SOC.,Chem. Commun. 1991, 1499.
(b) Bassindale, A. R.; Borbaruah, M. J. Chem. SOC.,Chem. Commun. 1991, 1501.
(c) Bassindale, A. R. Coordination and Reactivity in Organosilicon Chemistry, In
Progress in Organosilicon Chemistry (Proceedings of the 1 dhInternational Sym-
posium on Organosilicon Chemistry, Poznan, Poland, 1993), Marceniec, B.;
Chojnowski, J., Eds., Gordon & Breach: Basel, 1995, p. 191.
18. (a) Yoder, C. H.; Ryan, C. M.; Martin, G. F.; Ho, P. S. J. Organomet. Chem. 1980,
190, 1. (b) Macharashvili, A. A.; Shklover, V. E.; Struchkov, Y. T.; Oleneva, G. I.;
Kramarova, E. P.; Shipov, A. G.; Baukov, Y. I. J. Chem. SOC.,Chem. Commun. 1988,
683. (c) Boer, F. P.; van Remoortere, F. P. J. Am. Chem. SOC.1970, 92, 801. (d)
Albanov, A. I.; Gubanova, L. I.; Larin, M. F.; Pestunovich, V. A.; Voronkov, M. G.
138 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

J. Organomet. Chem. 1983,244, 5 . (e) Onan, K. D.; McPhail, A. T.; Yoder, C. H.;
Hillyard, R. W. J. Chem. SOC., Chem. Commun. 1978, 209.
19. Voronkov, M. G.; Frolov, Y. L.; DYakov, V. M.; Chipanina, N. N.; Gubanova, G. I.;
Gavrilova, G. A.; Klyba, L. V.; Aksamentova, T. N. J. Organomet. Chem. 1980,201,
165.
20. Coordination expansion at silicon from 4 to 5 is accompanied by an upfield shift of
about 30 ppm in the 29Si NMR; see Chapter 1.
21. The terms axial and apical will be used interchangeably. As a general rule,
thermodynamically more stable complexes are formed when electronegative
ligands are in the axidapical position rather than the equatorial position. These
ligands are thus “apicophilic.”
22. Tandura, S. N.; Voronkov, M. G.; Alekseev, N. V. Top. Cum Chem. 1986,131,99.
23. Gillespie, R. J.; Hargittai, I. The VSEPR Model of Molecular Geometry, Allyn and
Bacon, Boston, 1991, p. 56.
24. Huheey, J. E. Inorganic Chemistry, 31d Ed., Harper and Row, 1983, pp. 230-232.
25. Ref. 23, pp. 55-57.
26. (a) Bent, H. A. J. Chem. Phys. 1960,33,1259. (b) Bent, H. A. J. Chem. Phys. 1960,
33, 1260. (c) Bent, H. A. Chem. Rev. 1961, 61,275.
27. Holmes, R. R. Chem. Rev. 1990, 90, 17.
28. Westheimer, F. H. Acc. Chem. Res. 1968, 1, 70.
29. Trippett, S. Pure Appl. Chem. 1974, 40, 595.
30. (a) Kepert, D. L. Inorg. Chem. 1973, 12, 1938. (b) Kepert, D. L. Inorganic
Stereochemistry, Springer: Berlin, 1982.
3 1. (a) Luckenbach, R. Dynamic Stereochemistry of Pentaco-ordinated Phosphorus and
Related Elements, Georg Thieme Verlag: Stuttgart, 1973. (b) Favas, M. C.; Kepert,
D. L. Pmgr. Inorg. Chem. 1980,27, 325. (c) Demuynck, J.; Strich, A.; Veillard, A.
Nouv. J. Chim. 1977, 1, 217.
32. Klanberg, F.; Muetterties, E. L. Inorg. Chem. 1968, 7, 155.
33. For an excellent review, see: Comu, R. J. P.; Guerin, C.; Moreau, J. J. E. Dynamic
Stereochemistry at Silicon, In The Chemistry of Organic Silicon Compounds, Patai,
S . ; Rappoport, Z., Eds., Wiley: Chichester, UK, 1989, Vol. 1, Chap. 4, p. 305.
34. Cotton, F. A.; Wilkinson, G. Advanced Inorganic Chemistry, Third Ed., Wiley: New
York, 1972, pp. 40-41.
35. Berry, R. S. J. Chem. Phys. 1960,32,933.
36. Much of this discussion is covered more extensively (and beautifully) in: (a)
Bassindale, A. R.; Taylor, P. G. Reaction Mechanisms of Nucleophilic Attack at
Silicon, In The Chemistry of Organic Silicon Compounds, Patai, S.; Rappoport, Z.,
Eds., Wiley: Chichester, UK, 1989, Vol. 1, Chap. 13, p. 839. (b) Bassindale, A. R.;
Glyne, S. J.; Taylor, P. G. Reaction Mechanisms ofNucleophilic Attack at Silicon, In
The Chemistry of Organic Silicon Compounds, Rappoport, Z.; Apeloig, Y., Eds.,
Wiley: Chichester, UK, 1998, Vol. 2, Chap. 9, p. 495.
37. (a) Muetterties, E. L. J. Am. Chem. SOC. 1969,91,4115. (b) Holmes, R. R.; Deiters,
R. M.; Golen, J. Inorg. Chem. 1969, 8, 2612.
38. Spiridonov, V. P.; Ischenko, A. A,; Ivashkevich, L. S. J. Mol. Struct. 1981, 72,
153.
REFERENCES 139

39, Corriu, R. J. P.; Mazhar, M.; Poirier, M.; Royo, G. J. Organomet. Chem. 1986,306,
c5.
40. Corriu, R. J. P.; Kpoton, A.; Poirier, M.; Royo, G.; de SaxcC, A.; Young, C. J.
J. Organomet. Chem. 1990, 395, 1.
41. (a) Stevenson, W. H., 111; Wilson, S.; Martin, J. C.; Farnham, W. B. J. Am. Chem.
SOC.1985, 107, 6340. (b) Stevenson, W. H., III; Martin, J. C. J. Am. Chem. SOC.
1985,107,6352.
42. For related compounds, see: (a) Famham, W. B.; Harlow, R. L. J. Am. Chem. SOC.
1981,103,4608. (b) Famham, W. B.; Whitney, J. F. J. Am. Chem. SOC. 1984,106,
3992.
43. (a) Boudin, A.; Cerveau, G.; Chuit, C.; Corriu, R. J. P.; Reye, C. Angew. Chem., Int.
Ed. Engl. 1986, 25, 473. (b) Boudin, A.; Cerveau, G.; Chuit, C.; Comu, R. J. P.;
Reye, C. Angew. Chem., Int. Ed. Engl. 1986,25,474, and references cited therein.
44. For Y =Si (Chart 5-2), see: Kira, M.; Sato, K.; Kabuto, C.; Sakurai, H. J. Am. Chem.
SOC. 1989, 111, 3747.
45. Boyer, J.; Comu, R. J. P.; Kpoton, A.; Mazhar, M.; Poirier, M.; Royo, G.
J. Organomet. Chem. 1986,301, 131.
46. Damrauer, R.; Danahey, S. E. Organometallics 1986, 5, 1490.
47. Kalikhman, I.; Krivonos, S.; Ellern, A.; Kost, D. Organometallics 1996, 15, 5073.
48. Ref. 33, pp. 340-341.
49. Brelibre, C.; Carri, F.; Comu, R. J. P.; Poirier, M.; Royo, G. Organometallics 1986,
5, 388.
50. Ref. 33, pp. 342-351.
51. Corriu, R.; Royo, G. Tetrahedron 1971, 27, 4289.
52. (a) Comu, R.; Masse, M.; Royo, G. J. Chem. Soc., Chem. Commun. 1971, 252.
(b) Corriu, R.; Fernandez, M.; Guerin, C.; J. Organomet. Chem. 1978, 152, 21.
53. Sommer, L. H.; Parker, G. A.; Lloyd, N. C.; Frye, C. L.; Michael, K. W. J. Am.
Chem. SOC.1967, 89, 857.
54. (a) Corriu, R. J. P.; Fernandez, J. M.; Guerin, C. J. Organomet. Chem. 1978,152,25.
(b) Corriu, R. J. P.; Fernandez, J. M.; Guerin, C. J. Organomet. Chem. 1978,152,31.
55. Sommer, L. H.; Michael, K. W.; Korte, W. D. J. Am. Chem. SOC. 1967,89, 868.
56. Comu, R.; Royo, G. J. Organomet. Chem. 1968, 14, 291.
57. Sommer, L. H.; Korte, W. D.; Rodewald, P. G. J. Am. Chem. SOC. 1967, 89, 862.
58. Corriu, R.; Royo, G. Bull. SOC.Chim. Fr. 1972, 1497.
59. Comu, R.; Royo, G. J. Organomet. Chem. 1972,40, 229.
60. Sommer, L. H.; Frye, C. L.; Parker, G. A.; Michael, K. W. J. Am. Chem. SOC. 1964,
86, 3271.
61. Sommer, L. H.; Parker, G. A.; Frye, C. L. J. Am. Cliem. SOC.1964, 86, 3280.
62. Sommer, L. H.; Frye, C. L.; Parker, G. A. J. Am. Chem. SOC. 1964, 86, 3276.
63. Sommer, L. H.; Korte, W. D. J. Am. Chem. SOC.1967, 89, 5802.
64. Sommer, L. H.; Frye, C. L.; Musolf, M. C.; Parker, G. A.; Rodewald, P. G.; Michael,
K. W.; Okaya, Y.; Pepinsky, R. J. Am. Chem. SOC.1961, 83, 2210.
65. Sommer, L. H.; McLick, J. J. Am. Chem. SOC.1966, 88,5359.
66. Anh, N. T.; Minot, C. J. Am. Chem. SOC.1980, 102, 103.
140 REACTION MECHANISMS FOR NUCLEOPHILIC SUBSTITUTION AT SILICON

67. Ref. 36, pp. 843-848.


68. (a) McKinnie, B. G.; Bhacca, N. S.; Cartledge, F. K.; Fayssoux, J. J. Am. Chem. SOC.
1974,96, 2637. (b) McKinnie, B. G.; Bhacca, N. S.; Cartledge, F. K.; Fayssoux, J.
J. Org. Chem. 1976, 41, 1534.
69. (a) Dubac, J.; Mazerolles, P.; Serres, B. TetrahedronLett. 1972,3495. (b) Dubac, J.;
Mazerolles, P.; Serres, B. Tetrahedron 1974, 30, 759.
70. (a) Cartledge, F. K.; Wolcott, J. M.; Dubac, J.; Mazerolles, P.; Fagoaga, P.
Tetrahedron Lett. 1975, 3593. (b) Wolcott, J. M.; Cartledge, F. K. J. Organomet.
Chem. 1976, 111, C35.
7 1. (a) Cartledge, F. K.; Wolcott, J. M.; Dubac, J.; Mazerolles, P.; Joly, M. J. Organomet.
Chem. 1978, 154, 187. (b) Dubac, J.; Mazerolles, P.; Joly, M.; Cartledge, F. K.;
Wolcott, J. M. J. Organomet. Chem. 1978, 154, 203.
72. Comu, R.; Masse, J. Bull. SOC. Chim. FK 1978, 10, 3491.
73. Sommer, L. H. Intra-Sci. Chem. Rep. 1973, 7, 1.
74. Myers, A. G.; Kephart, S. E.; Chen, H. J. Am. Chem. SOC. 1992, 114, 7922.
75. (a) Denmark, S. E.; Griedel, B. D.; Coe, D. M. J. Org. Chem. 1993, 58, 988. (b)
Denmark, S. E.; Griedel, B. D.; Coe, D. M.; Schnute, M. E. J. Am. Chem. SOC.1994,
116,7026.
76. Matsumoto, K.; Oshima, K.; Utimoto, K. J. Org. Chem. 1994, 59, 7152.
77. Comu, R. J. P.; Guerin, C. J. Chem. SOC., Chem. Comrnun. 1977, 74.
78. (a) Corriu, R. J. P.; Guerin, C.; Masse, J. J. Chem. SOC., Chem. Comrnun. 1975,75.
(b) Corriu, R. J. P.; GuCrin, C.; MassC, J. J. Chem. Res. ( S ) 1977, 160. (c) Comu, R.
J. P.; Guerin, C.; Masse, J. J. Chem. Res. (S) 1977, 160
79. Guerin, C., Ph.D. Thesis, Montpellier, 1978 (see Ref. 7).
80. Comu, R. J. P.; Fernandez, J. M.; GuCrin, C. Tetrahedron 1981, 37, 2467.
81. Bassindale, A. R.; Glynn, S. G.; Jiang, J.; Turtle, R.; Taylor, P. G.; Brown, S. S. D.
Recent Explorations of the Chemistry of Pentacoordinate Silicon, In Organosilicon
Chemistry: From Molecules to Materials II (Proceedings of the Munich Silicon
Days, 1994), Auner, N.; Weiss, J., Eds., VCH: Weinheim, 1996, p. 411.
82. (a) Comu, R. J. P.; Dabosi, G.; Martineau, M. J. Chem. SOC., Chem. Comrnun. 1977,
649. (b) Corriu, R. J. P.; Dabosi, G.; Martineau, M. J. Organomet. Chem. 1978,154,
33. (c) Comu, R. J. P.; Dabosi, G.; Martineau, M. J. Organomet. Chem. 1978,150,
21.
83. (a) Comu, R. J. P.; Guerin, C.; Henner, B. J. L.; Wong Chi Man, W. W. C.
Organometallics 1988,7,237. (b) CarrC, F.; Cerveau, G.; Chuit, C.; Corriu, R. J. P.;
RCyC, C. Angew. Chem., Int. Ed. Engl. 1989, 28, 489.
84. Corriu, R. 3. P. J. Organomet. Chem. 1990, 400, 81.
85. Schott, G.; Kelling, H.; Schild, R. Chem. Ber. 1966, 99, 291.
86. (a) Allen, A. D.; Charlton, J. C.; Eaborn, C.; Modena, G. J. Chem. SOC. 1957,3668.
(b) Allen, A. D.; Charlton, J. C.; Eaborn, C.; Modena, G. J. Chem. SOC. 1957,3671.
87. Chiu, H. K.; Hohnson, M. D.; Frye, C. L. J. Organomet. Chem. 1984, 271,327.
88. Comu, R. J. P. Hypervalent Species of Silicon: Structure and Reactivity, In Frontiers
of Organosilicon Chemistry (Proceedings of the qhInternational Symposium on
Organosilicon Chemistry, Edinburgh, I990),Bassindale, A. R.; Gaspar, P. P., Eds.,
Royal Society of Chemistry: Cambridge, 1991, p. 185.
REFERENCES 141

89. Corriu, R. J. P.; Henner, B. J. L. J. Organomet. Chem. 1974, 71, 393.


90. (a) Comu, R. J. P.; Leard, M. J. Chem. SOC.,Chem. Commun. 1971,1086.(b)Comu,
R. J. P.;Henner, M. Bull. SOC.Chim. Fr. 1974,1447.(c) McKinnie, B. G.; Cartledge,
F. K. J. Organomet. Chem. 1976, 104, 407. (d) Chojnowski, J.; Cypryk, M.;
Michalski, J.; Wozniak, L. J. Organomet. Chem. 1985, 288, 275.
91. Cartledge, F. K.; McKinnie, B. G.; Wolcott, J. M. J. Organomet. Chem. 1976,118,7.
92. Comu, R. J. P.; Henner-Leard, M. J. Organomet. Chem. 1974, 65, C39.
93. Bassindale, A. R.; Stout, T. Tetrahedron Lett. 1985, 261, 3403.
94. Taft, R. W.; Gramstad. T.; Kamlet, M. J. J. Org. Chem. 1982, 47, 4557, and
references cited therein.
95. Pierce, A. E. Silylation of Organic Compounds ( A Technique for Gas Phase
Analysis), Pierce Chemical Co.: Rockford, Ill., 1968.
96. Corey, E. J.; Venkateswarlu, A. J. Am. Chem. SOC.1972, 94, 6190.
97. Comu, R. J. P.; Perz, R.; Reye, C. Tetrahedron 1983, 39, 999.
98. (a) Tamao, K.; Ishida, N.; Tanaka, T.; Kumada, M. Organometallics 1983,2, 1694.
(b) Tamao, K.; Ishida, N. J. Organomet. Chem. 1984, 269, C37. (d) Tamao, K.;
Kumada, M.; Maeda, K. Tetrahedron Lett. 1984,25, 321. (d) Tamao, K.; Hayashi,
T.; Ito, Y. Hydrogen-Peroxide Oxidation of the Silicon-Carbon Bond: Mechanistic
Studies, In Frontiers of Organosilicon Chemistry (Proceedings of the qhInterna-
tional Symposium on Organosilicon Chemistry, Edinburgh, 1990), Bassindale,
A. R.; Gaspar, P. P., Eds., Royal Society of Chemistry: Cambridge, 1991, p. 197.
99. Fleming, I.; Henning, R.;Parker, D. C.; Plaut, H. E.; Sanderson, P. E. J. J. Chem.
SOC.,Perkin Trans. I 1995, 317.
100. (a) Tamao, K.; Kakui, T.; Akita, M.; Iwahara, T.; Kanatani, R.; Yoshida, J.; Kumada,
M. Tetrahedron 1983,39,983. (b) Uozomi, Y.; Lee, S.-Y.; Hayashi, T. Tetrahedron
Lett. 1992,33,7185. (c) Uozomi, Y.; Hayashi, T. Tetrahedron Lett. 1993,34,2335.

You might also like