You are on page 1of 16

Polyhedron Vol. 12, No. 20, pp. 2431-2446, 1993 0277-5387/93 $6.00+ .

oO
Printed in Great Britain Pergamon Press Ltd

POLYHEDRON REPORT NUMBER 49

STABLE COMPOUNDS OF THE HEAVIER GROUP 14 AND 15


ELEMENTS INVOLVING p-p MULTIPLE BONDING: AN
OVERVIEW OF THE FIRST DECADE

NICHOLAS C. NORMAN

The University of Newcastle upon Tyne, Department of Chemistry,


Newcastle upon Tyne NE1 7RU, U.K.

CONTENTS

1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . 2431

2. ASPECTS OF SINGLE AND DOUBLE BOND ENERGIES . . . . . . . . . 2432

3. STERIC PROTECTION . . . . . . . . . . . . . . . . , . 2434

4.GROUP14 ........................ 2435


4.1. Compounds involving double bonds to a first row element ....... 2435
(i) Silenes (S&C) and their heavier congeners ........... 2435
(ii) Silanimines (Si=N) and their heavier congeners ......... 2436
4.2. Compounds involving double bonds without a first row element ...... 2437
(i) Disilenes (Si=Si) and their heavier congeners .......... 2437
(ii) Silaphosphenes (Si=P) and their heavier congeners ........ 2438
(iii) Thiosilanones or silanethiones (S&S) ............. 2439

5. GROUP 15 . . . . . . . . . . . . . . . . . . . . . . . . 2439
5.1. Compounds involving double and triple bonds to a first row element . . _ . . 2439
(i) Phospha-alkenes (P=C), phospha-alkynes (F%C) and related compounds . 2439
(ii) Phospha-imines (P=N) . . . . . . . . . . . . . . . . . 2441
5.2. Compounds involving double bonds without a first row element . . . . . 2441
(i) Diphosphenes (P=P) and their heavier congeners . . . . . . . . 2441

6. CONCLUSION AND GENERAL FEATURES . . . . . . . . . . . . . 2443

ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . 2444

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . 2444

1. INTRODUCTION

In Dasent’s 1965 monograph entitled “Nonexistent Compounds”, ’ molecules of the heavier main
group elements containing pn-pr multiple bonds warranted a chapter to themselves, the point
being made that whilst multiple bonding was common for the first row elements carbon, nitrogen
and oxygen, it was conspicuous by its absence in the chemistry of the heavier congeners (with the
exception, to some extent, of sulphur). This was the situation not for lack of effort on the part of
synthetic chemists, but rather because many of the compounds which had been described in the
early literature as having multiple bonds were later shown to be cyclic oligomeric species containing
only single bonds. For example, the compound, described in 1877 by Kiihler and Michaelis as
2431
2432 N. C. NORMAN

“phosphobenzene”, was formulated as PhP=PPh by analogy with the well known azobenzene,
PhN=NPh, but it was later shown to adopt cyclic pentameric and hexameric forms in the solid
state.’ Similarly, compounds which were once thought to contain As=As, Si=Si and Si=C double
bonds were also later shown to be oligomeric. Not surprisingly, perhaps, these results led to a belief
in a “double bond rule” whereby elements of the second and subsequent rows were deemed not to
form stable prc-pn multiple bonds. 3 This was the state of affairs which existed until a little over ten
years ago and it is possible that it was the belief in a double bond rule which was largely responsible
for the stagnation in this field. In this regard, the parallel with the stories of perbromate4 and noble
gas chemistry’ are illuminating since the important discoveries in each of these cases were preceded
by a period in which many reasons for their nonexistence were advanced ; it is also noteworthy how
many of the compounds in Dasent’s book have, in fact, now been synthesized.
The situation changed dramatically, however, in the early 1980s with the synthesis and char-
acterization of the first stable* compounds with a phosphorus-phosphorus (diphosphenes, RP=PR)
and a silicon-silicon (disilenes, R$i=SiR*) double bond, (1)6 and (2),’ respectively, together with
examples of stable phospha-alkenes (RP=CR,)8 and silenes (R,Si=CR,).9

k-3
- Si Si

8-3 2

The area has expanded considerably since this time and is very much part of the renaissance of
main-group chemistry which has occurred during this same period. In this review, we shall look
back over the first decade of the subject addressing first some thermodynamic aspects regarding the
strengths of single and double bonds for the p-block elements, followed by an overview of some of
the important compounds which have been synthesized, specifically, those in groups 14 and 15
which contain isolated (i.e. non-conjugated) multiple bonds which can be described as of thepn-p7t
type. Conjugated compounds and the extensive chemistry of multiply bonded sulphur species will
not be discussed. Particular emphasis is given to structure and bonding ; details on the methods of
synthesis and on reactivity studies can be found in the various review articles which are referenced.

2. ASPECTS OF SINGLE AND DOUBLE BOND ENERGIES

In seeking an explanation for the varying stability of pn-pn multiple bonds‘between the p-block
elements, we must look first at the strengths of the various bonds involved and an illuminating way
in which this can be done is to examine the 0 and n-bond increments for particular pairs of atoms.
Table 1 lists such increments for several pairs of elements, and is derived largely from the recent
work of Kutzelnigg, lo although the origin of this approach goes back to Pauling. ” The values in
Table 1 are defined such that the first value is that for a typical E-E single or a-bond ; for example
the value of 335 kJ mol-’ for carbon is that of the C-C bond in ethane. The second value is the
difference in energy between a double, or (T+ n-bond and that of the single bond given by the first
value and this can be taken as a measure of the strength of the n-bond. For carbon, the difference
is that between the bond energy of the C=C double bond in ethene (630 kJ mol- ‘) and the single
C-C bond in ethane. Clearly the a-bond in ethane is not the same as the a-bond in ethene but the

*The term “stable” is used here to imply that the compounds can be isolated, characterized and stored
under ambient conditions of temperature and pressure ; in other words, that they are stable with respect to
dimerization or oligomerization and/or other facile reaction types.
Heavier group 14 and 15 elements involving px-pn multiple bonding 2433
Table 1. Approximate bond increments (kJ mol- ‘) for a/n bonds adapted from
ref. 10

N-N c-o
(160/395) (335/380)
Si-Si P-P Si-0
(195/120) (200/145) (27%;5) (420/170) (31$0)
Ge-Ge As-As Se-Se c-s
(1651110) (1751120) (2101125) (2801265) (2GO)

differences are likely to be small ; moreover, it is broad comparisons between different types of
bonds that are important rather than a means of quantitatively predicting particular bond energies.
If we first consider homonuclear bonds, an important point to emerge from Table 1 is that the
It-bond increment for the first row or 2p elements is considerably greater than that for the elements
of the second and subsequent rows. In other words, n-bond strengths are much larger for the first
row elements and by a factor of between two and three. The standard explanation for this feature is
that the second and subsequent row elements, being larger, have correspondingly larger and therefore
more diffuse orbitals. Overlap between these bigger orbitals is poorer which results in weaker bonds,
particularly in the case of rc-bonds in which the atomic orbitals overlap side-on rather than the head-
on overlap of o-orbitals. Whilst this is probably broadly correct, and is an example of a defining
feature of p-block chemistry, wherein bond strengths generally decrease down a group, it is only
part of the explanation for the absence of extensive R-bonding in the chemistry of the heavier
elements. The other aspect which we must consider is the strengths of the single or a-bonds between
the first row elements. Thus for nitrogen and oxygen, the N-N and O-O single bonds (like the
F-F bond) are exceptionally weak as is clear from Table 1. This feature is usually associated with
the small size of the first row atoms, leading to significant repulsions between nonbonded pairs of
electrons, and to the fact that relatively high energy rc*-orbitals are occupied. These factors are of
much lesser importance for the larger, heavier elements, the latter in particular since any factors
which result in weak rr bonding will also mitigate the effects of occupation of rr*-orbitals. As a result,
P-P and S-S single bonds are considerably stronger than N-N and O-O bonds, respectively.
In considering both of these factors, it is therefore apparent that for groups 15 and 16, as the values
in Table 1 clearly indicate, the combination of a c + rc bond is more stable than two o-bonds for the
first row elements whereas the reverse is true for the elements of the second and subsequent rows.
In group 14 the comparison is a little different in that the strength of the C-C single bond is
greater than that of the Si-Si bond since there are no nonbonded electron pairs associated with
the carbon atoms. The absence of an extensive chemistry of multiply bonded silicon compounds is
due, as before, to the weakness of n-bonding for this element (two o-bonds are stronger than one (T
and one x-bond) and it is noteworthy that Si-Si n-bonding is the weakest of the 3p elements (see
Table 1) consistent with the larger size of the silicon atom (vs P or S). The occurrence of R-bonding
in carbon chemistry, however, is not associated with any weakness of the C-C a-bond but much
of the extensive and rather special organic chemistry of carbon is the result of the similarity in the
strengths of the C-C cr and n-bonds. With this final point in mind, it is also clear from the values
for group 14 in Table 1 that, in absence of factors which weaken the first row o-bond, the ratio of
the a/rc bond strengths drops markedly between the first and second row (0.88 vs 0.61) but then
remains nearly constant for the second and third row (0.61 vs 0.66).
An examination of 0 and 7cincrements for heteronuclear bonds is also instructive. For example,
Table 1 reveals that the rc-bond increment is large where both elements are from the first row
although the a-bond strengths are also large where one of the elements is carbon. Heteronuclear
bonds where a second row element is involved, for example Si-0 and P-O, are similar to the
homonuclear second row element cases in having a small 71increment but we should note the rather
special case of sulphur wherein the 0 and A increments for both C-S and S-O bonds are about
equal. This is mainly the result of a large K increment (as opposed to weak a-bonding, cf. C-C)
and in this sense, sulphur is acting (as far as n-bonding is concerned) more like a first row element.
This is undoubtedly due to the relatively small size of the sulphur atom and the consequent better
2434 N. C. NORMAN
rc overlap, and it is therefore not surprising that px-,rr multiple bonding is important in the
chemistry of this element.

3. STERIC PROTECTION

In the previous section it was shown, on the basis of bond energy considerations, why single
bonding is expected to be much more common than pn-pa multiple bonding for the heavier p-
block elements. However, an analysis which deals only with strengths of particular bonds does not
encompass other important features. The synthesis of stable compounds containing isolated multiple
bonds, which are the subject of this report, was, in fact, achieved by employing sterically bulky
substituents, typical examples of which are CsH2R3-2,4,6 (R = Me, Et, Pr’, Bu’, CF& (Me,Si),C,
(Me$i),CH and Ph,Si as shown below.

\
R Ph

Me3Si”““”

Me,Si
t Ph,,,,.. Si-

Ph
/

R = Me, Et, Pr’, Bu’, CF3 R = Me& H

The efficacy of such groups in this regard has often been stated to be kinetic in origin with the
implication that they protect an unstable functional group from oligomerization by means of their
size. However, whilst there is much truth in this assertion, it is not really kinetic stabilization. If we
take as an example some compounds which have already been mentioned, then for the RP or
phosphinidene unit, when R is phenyl, cyclic pentameric and hexameric species are formed whereas
when R is the much more bulky aryl substituent CgH2But3-2,4,6 (hereafter Ar’ and often referred
to as supermesityl), it is a diphosphene which is obtained. If we now consider various possible ring
sizes and also that a diphosphene can be thought of, albeit somewhat perversely, as a two-membered
ring, it is clear from the diagrams below that more bulky substitutents will favour smaller ring sizes
since the R-P-P angle(s), and hence the R * * . R separation, increase as the ring size decreases.
This argument assumes that the P, rings are planar and that the bonded atoms of the R groups are
coplanar with the P, ring which will not generally be the case, but the basic point of the argument
remains in that smaller rings will more readily accommodate larger R groups.

I I
\p/p\p/ \p/p\p/
I I
\ I
/pLp/p\ P-P

I ,”
\ I
P-P
1 I/ /‘\ -PZP-
P-P
/ \ /p-p \

On this basis, it is apparent that the stabilization of a diphosphene by sterically demanding


substituents is not really kinetic stabilization but it is a form of thermodynamic stabilization since
it is certainly not the case that 1 is a kinetic product which is somehow prevented (by a large
activation energy) from forming a larger, and thermodynamically more stable, P, ring. Thus, R-
bonded compounds (two membered rings) can be rendered quite stable if the a-bonded alternatives
(larger rings) are made less stable by using large and bulky side groups ; in effect, these large groups
exert thermodynamic control over oligomerization and kinetic control over other reactions which
may occur at the double bond.
There is an interesting parallel to note here between the use of sterically demanding groups used
to stabilize P=P and Si=Si double bonds in diphosphenes and disilenes, respectively, and the
Heavier group 14 and 15 elements involving px-pa multiple bonding 2435
similar use of such groups to stabilize element-element single bonds in group 13 chemistry. Thus
compounds of the type R2E-ER2 containing two-centre, two-electron E-E single bonds are
particularly uncommon in the chemistry of the group 13 elements, especially for the heavier elements.
This instability is likely to be associated with the presence of a vacant p-orbital at each element
centre rather than with any inherent weakness of the E-E bonds; the B-B single bond, for
example, has a substantial bond energy of about 300 kJ mol- ‘. For boron, the use of strong n-
donors is efficacious in this regardI whereas for aluminium, gallium and indium, the sterically
demanding group (Me,Si)*CH has been utilised by Uhl and co-workers’3 to effect the isolation of
such compounds ; Cowley et al. have also described a di-indium compound stabilized by four
CgH2(CF3)3-2,4,6 groups.14

(Me,Si),CH ,CH(SiMe,),
\
E-E
/ \
(Me,Si),CH CH(SiMe,),

E = Al, Ga, In.

In this case, the sterically bulky (Me3Si)$H groups probably are effective in stabilizing the
E-E bond for kinetic reasons in that they inhibit reactions which would degrade the E-E bond,
particularly disproportionation reactions.
Having looked at some general properties which affect the stability of multiply bonded com-
pounds in the p-block, we will now look at some specific examples. This is certainly not a com-
prehensive or exhaustive list but the compounds described serve to illustrate some of the general
features associated with this subject.

4. GROUP 14

4.1. Compounds involving double bonds to a first row element

(i) Silenes (S&C) and their heavier congeners. The first stable compound in this class to be
structurally characterized was (Me,Si)$i=C(OSiMe,)(l-ad) (3) (ad = adamantyl) described by
Brook and co-workers” in 1982. Key features of this compound include a Si=C bond which is
approximately 10% shorter than a typical Si-C single bond, and the near coplanarity of the S&C
unit and the atoms to which it is bonded. Other important features, derived from NMR spectroscopy,
are the inequivalence of the Me,Si gro.ups, which indicates hindered rotation about the Si=C bond,
a deshielded 29Si NMR chemical shift and a large ‘J(SiC) coupling constant. All of these features
are important in establishing the presence of double bonding akin to that commonly encountered
for the first row elements.

Me,.%, ,OSiMe, ,SiMe,


Me\
,Si==C Si=C,
/
Me,Si Me SiMeBd2

3 4

An additional and important structural feature of 3 is the presence of four bulky groups
surrounding the Si=C unit which is undoubtedly responsible, in large part, for the stability of 3
with regard to dimerization, a usually facile reaction for less hindered silenes. However, the
subsequent isolation and structural characterization of complex 4 by Wiberg and co-workers’ti
shows that massive steric protection is not always a necessary condition, at least as far as the silicon
2436 N. C. NORMAN

centre is concerned, for stability towards dimerization ; there are other examples of complexes (see
later) where steric factors alone are unlikely to prevent dimerization. It is also noteworthy that the
silicon centre in 4 is quite Lewis acidic and readily accommodates a coordinated thf (tetrahydrofuran)
molecule to give complex 5.‘6bLewis acidity is commonly associated with three-coordinate, multiply
bonded silicon centres in silenes, silanimines (see later) and in transition metal silylene complexes,
L,M=SiR,.‘7
Compounds containing Ge=C double bonds (germenes) are fewer in number and have been
reviewed by Barrau et al. I8 Two examples, 619 and 7 *O**’have been structurally characterized and
both contain Ge=C bonds which are significantly shorter than typical Ge-C single bonds in accord
with that found for silenes. However, whilst in 6 the atoms to which the Ge=C unit is bonded are
nearly coplanar, in 7 the average twist angle, 4, defined in the diagram below, is 36”. Relatively
large twist angles are sometimes encountered in complexes of this and related types (see later) which
reflects the relative weakness of the n-bond. It should also be noted that in both 6 and 7 there is the
possibility of 7c-conjugation with the substituents which may enhance their stability.

(MeBihN, ,!z ,SiMe,

(Me&N

6 7

Stannenes (compounds with Sri=== double bonds) related to 6’* and 723 have also been reported
(the latter characterized by X-ray crystallography) together with the stannaketenimine (C6H2(CFJ3-
2,4,6),Sn=C=N(mes) (8),24 the structure of which will be discussed later.
(ii) Silanimines (Si=N) and their heavier congeners. The first stable silanimine (Bu’),Si=N
(SiBut3) (9) was described by Wiberg and co-worker? in 1986. The Si=N bond distance (1.568 A)
is significantly shorter than typical Si-N single bonds and is also shorter than the N-SiBu’,
bond length in 9 (1.695 A), but of more interest is the almost linear coordination about the nitrogen
atom (S&N-Si 177.8”). Such a geometry is clearly inconsistent with a bonding scheme as
represented in A with a localized lone pair on the nitrogen atom ; some degree of delocalization
involving a vacant orbital on the SiButj silicon centre, as shown in B, is implied and in which the
nitrogen is better described as sp hybridized rather than sp2. Delocalization of nitrogen lone pairs
onto adjacent silicon centres is not uncommon and is seen, for example, in the planar structure of
silyl substituted amines such as N(SiMe,)3 and related compounds.26 As with the silenes, the
three-coordinate silicon centres in silanimines are appreciably Lewis acidic and readily coordinate
(albeit weakly) a Lewis base. The structure of the thf adduct Me,(thf)Si=N(SiBu’,) (lo)*’ reveals
a S&N bond (av. 1.581 A for two independent molecules) slightly lengthened compared to 9 due
to the fact that the acceptor orbital on silicon is probably the Si=N n*-orbital, and also that the
nitrogen centre is now noticeably bent (Si=N-Si av. 161.3”). Additional structures of Lewis
base adducts (with Si=N distances and Si-N-Si angles in parentheses) are Bu’,(Ph,CO)Si=N
(SiBu13) (11) (1.601 A, 169.3”)27 and Bu’,(thf)Si=N(SiBu’,Me) (12) (1.596 A, 174.3”).** Moreover,
it is apparent when comparing the structures of the similarly substituted compounds 11 and 12, that
a longer Si=N bond correlates with a smaller Si-N-Si angle which is in accord with a limiting,
zwitterionic structure C.

,Si=Q; SiBu:
BU’ 3

R
‘Si=N ‘Si=N--_R
\
R’ R d R
A B c
Heavier group 14 and 15 elements involvingpn--pn multiple bonding 2437
Many stable germanimines ( G ~ N ) have been described and the topic has been reviewed ~8
although the only crystal structure determinations that have been reported are those for {(R)
(Me3Si)N}2G~N(R) (R = rues, C6H3Pra2-2,6).29 Both compounds have short G ~ N bonds and
planar structures consistent with the presence o f p n - - p n G ~ N bonds.

4.2. Compounds involving double bonds without a first row element


(i) Disilenes 3° (Si=Si) and their heavier congeners. The first stable disilene 2 was reported in
1981 by West, Fink and Michl7 and the crystal structure determination of 2 was reported in 19843t
together with that of the E isomer of (mes)(But)Si--Si(mes)(Bu t) (mes = C6H2Me3-2,4,6) (13), 31
the structures of (C6H3Et2-2,6)2Si=Si(C6H3Et2-2,6)2 (14)32 and (E)-(C6H2Pri3-2,4,6)(R)Si~---Si
(C6H2Pt~3-2,4,6)(R) (15) (R = But, SiMe3) have also been described 33 together with those of the
air stable compound (C6H2Pri3-2,4,6)2Si~---Si(C6H2Pri3-2,4,6)234 and E-(mes)(1-ad)Si--Si(mes)(1-
ad). 35 In all complexes, the Si~--~--Sibond length is around 2.15 ___0.01 ~ which is about 10% shorter
than typical Si--Si single bond lengths consistent with double bond character for the Si--Si unit
(C-----C bonds are typically about 10-12% shorter than C--C bonds with comparably sized sub-
stituents). Moreover, in complexes 13 and 15, the Si---Si unit and the atoms to which it is bonded
are coplanar where the twist angle ~b and the pyramidalization angle 0 are zero (0 is defined as
the angle made by the E--E' vector to the RRE plane as shown in the diagram below). Maximum
deviations from planarity are found in 2 for which 0 is 18° and in 14 for which #5 is 10%

Si : Si Si ....... Si

Bu t
Et Et

13 14

a ......e'~ ..... E']°

Other factors consistent with a double bond formulation for disilenes are the deshielded 29Si
NMR chemical shifts and large ~J(SiSi) coupling constants. Moreover, for 13, E and Z isomers can
be isolated (although these equilibrate in solution, especially under photochemical conditions),
NMR data for which establish a barrier to rotation about the Si=Si bond of 131 kJ mol- J. This
can be compared with the rotational barrier in stilbene which is 179 kJ mol- ~, i.e. about 30% larger.
Tetra-alkylated disilenes are {(Me3Si)2CH}2Si~--~-Si{CH(SiMe3)2}236 and But2Si--SiBut237 the
former being air stable and rather unreactive due to the extremely large substituents.
Digermenes which have been structurally characterized are (C6H3Et2-2,6)2G~----Ge(C6H3Et2-
2,6)2 (16) 38 and (C6H3Et2-2,6)(mes)G~----Ge(C6H3Et2-2,6)(mes) (17) 39 by Masamune et aL, together
with {(Me3Si)2CH}2G~----Ge{CH(SiMe3)2}2 (18) 40 and the distannene, {(Me3Si)2CH}2Sn=-Sn{CH
(SiMe3)2}: (19) 40 which have been described by Lappert and co-workers.

Ge ..... Ge

16

CH(SiMe92

(Me3Si)zCH,~)E E "'~'' C n ( s i M e ~ z

(Me3Si)2CH
18, E = G e ; 19, E = Sn.
2438 N. C. NORMAN

III 16 38,3vthe twist angle, 4, is IO” and the pyramidalization angle, 0, is 12” neither of which is
significantly larger than those found in disilenes. In 17, however, the germanium atoms are markedly
pyramidal (0 = 36”, 4 = 7”)3v and compounds 18 and 19 are significantly distorted from a planar
geometry such that whilst 4 is zero in both cases, they adopt strongly trans bent structures, where
8 is 32” and 4 1’) respectively. 4oClearly a standard cr+ rc bonding description applicable to olefins is
reasonable for the disilenes and the digermene 16 but is not appropriate for the digermenes 17 and
18 and the distannene 19. The bonding in these latter two species has been addressed by Lappert et
aL4’ Pauling4’ and in a series of papers by Trinquier and co-workers.42 The basis of the argument
is that a planar structure as shown in D is only feasible if the triplet state of the ER2 fragment is
energetically accessible. For carbon, this is certainly the case (the triplet state of CH;? is slightly
lower in energy than the singlet state) but as the group is descended (C, Si, Ge, Sn), not only is the
singlet state more stable, but the singlet-triplet energy separation, AEST, increases. In the situation
where AEsT > 1/2E-E(o + rr), where E-E(s + 7t) is the double bond energy, a tram bent structure
should be preferred. For germanium and certainly for tin, AEsT is sufficiently large and the planar
structure of type D is no longer the energy minimum. The bonding in the resultant tram bent
structure may therefore be represented as an interaction of two singlet states as shown in E.

D E

In the stannaketenimine 8,24 the tin atom is also markedly pyramidal (sum of angles = 290.9”)
and this compound may be thought of as linked, singlet stannylene (R,Sn) and isocyanide (C=NR)
species as shown in the diagram below. Consistent with this view is the bent geometry at the carbon
atom (153.9”) and the near linear coordination (175.0”) at nitrogen.

8, R, = C,H2(CF&-2,4,6

An alternative, although complementary, description is provided by Albright, Burdett and


Whangbo43 wherein the change in structure from type D to type E as group 14 is descended, is
considered in terms of a second order Jahn-Teller distortion due to mixing of the blurt and bZgrc*
levels with the 3b,a* and 3a,a* levels, respectively. These levels become closer in energy as the group
is descended leading to greater mixing and hence to a larger tram bent distortion. In this sense, the
increasing pyramidalization is seen as a special case of pyramidalization in eight-electron EX3 species
reasons for which are also discussed. Similar reasons for an increase in bending for six-electron EX2
species with a concomitant increase in AEST are also addressed. The % bond shortenings in 18 and
19 compared with typical single bond values are 4 and 2%, respectively which also indicates that
the standard 0 + rc double bonding model is not appropriate in these cases.
(ii) Silaphosphenes (Si-P) and their heavier congeners. The following compounds which are
examples of silaphosphenes, germaphosphenes (Ge=P) and stannaphosphenes (Sn=P) have been
described: RR’S;=PAr’ (R = R’ = CgHzRllj-2,4,6 for R” = Me, Et, Pr’; R = C6HzPr’,-2,4,6,
R’ = Ph, mes, But),““=* (mes),Ge=PAr’ (20)45 and {(Me3Si)ZCH),Sn=PAr’46 together with the
cumulene Ar’P=Si=PAr’. 47 Compound 20 has been structurally characterized45 and has a Ge==P
bond length which is about 8% shorter than a typical Ge-P single bond. Moreover, the coordination
around the Ge=P unit is essentially planar with important interbond angles as shown below, and
it is noteworthy that there is no appreciable pyramidalization at the germanium centre, similar to
the situation found for 16 ; the uneven angles to the mesityl groups are clearly a consequence of the
Heavier group 14 and 15 elements involving pn-pn multiple bonding 2439
supermesityl cis to one mesityl group. There has been one report of a sila-arsene, (C6H2Pt3-
2,4,6)2Si=AsSiPr’348 which has been characterized spectroscopically.

h+ci
0
GE! = P
0

(iii) Thiosilanones or silanethiones (Si=S). The only compounds to have been reported in
this class4’ are of the form shown below in which the silicon centre is additionally coordinated,
intramolecularly, by an amine function ; the electrophilic nature of three-coordinate silicon was
encountered with silenes and silanimines. Possible canonical forms are also represented.

B I=, B
Fbtr c-
0 Si=S Si- S-

0
( N,.c - -+

R = Ph, Naphthyl

Compounds of this type have also been isolated with Si-NBu’ and Si=PPh units50 and brief
mention should also be made of a compound containing a base stabilized Ge==S bond, 21 see
below,” and also of ESR evidence for the radical anion species [Bu’,Sis]-’ ‘* although no data
is yet available for stable neutral silanones, R,Si=O.

‘I
Si,
/N
k
+N,, /Ie=S
xNy
21

5. GROUP 15

5.1. Compounds involving double and triple bonds to a first row element

(i) Phospha-alkenes (P=C), phospha-alkynes (kc) and related compounds. Phospha-alkenes’


are one of the most extensively studied class of compound containing multiply bonded phosphorus.
The first report of stable non-conjugated phospha-alkenes was by Becker in 197653 who described
the complexes RP=CR’(OSiMe3) (R = R’ = Bu’; R = c-hexyl, R’ = Bu’; R = Ph, R’ = Bu’) and
examples of complexes which have subsequently been structurally characterized are (mes)P=CPh,
(22)54 and Ar’P=C(SiMe3)z (23)5’ which are shown below. Of particular interest are the structures
of the E and Z isomers of Ar’P=CH(Ph) (24) ;56a-cboth of these isomers are also configurationally
rigid in solution under ambient conditions indicating hindered rotation about the P=C bond.

22 23
2440 N. C. NORMAN

E. 24a Z. 24b

In general, a degree of steric protection is required in order for compounds containing isolated
P=C bonds to be stable (with respect to dimerization) but this condition is relaxed somewhat when
the P=C unit is part of a conjugated system or when the phosphorus is bonded to an electronegative
atom or group. Examples of the former class of compound are the phosphabenzenes CSH2R3-
2,4,6-P (R = H and Ph)57 whilst the latter is exemplified by ClP=C(SiMe& (25).‘* In 25, and a
series of related compounds, the chlorine atom, by virtue of its electronegativity, probably leads to
a phosphorus centre with a greater 6+ charge which in turn results in a contraction of the phosphorus
p-orbital which overlaps more effectively with the carbon p-orbital.
An important class of phospha-alkene is the metalla-phospha-alkene of which two basic types
have been characterized depending on the position of the organotransition metal fragment, viz
(L,M)P=CR* and RP=CR(ML,). Structurally characterized examples are (q-C5Me5)
(CO)zFe-P=C(SiMe,)z (26)59 and Bu’P=C(OSiMe~){Re(CO)(NO)(~-C,Me,)) (27)60 in which
the metal atoms are bonded to phosphorus and carbon, respectively.
Further derivatives of this class of compound are the phosphavinylidene complex (q-C5HS)
(CO)zMo=P=C(SiMe,), (28)61 reported by Cowley et al. which is an example where the phos-
phorus is multiply bonded to the metal atom as well as to carbon, and the arsa-alkene
complex (Y&H~)(CO),F~-A~=CB~‘(OS~M~,) (29)62 described by Weber and co-workers, one
of only very few examples of an arsenic-carbon double bond.

‘SiMe, OSiMe,
26 21

29

Important structural features associated with all the phospha-alkenes are a P=C bond length
which is about 8-10% shorter than a P-C single bond and the coplanarity of the atoms bonded
to the P=C unit. Angles at phosphorus are usually about loo”, except for some of the P-bonded
metalla-phospha-alkenes in which these angles are often a little larger. NMR parameters are also
consistent with the presence of a P=C double bond.
The first phospha-alkyne stable to polymerization, Bu’GP (30) was described in 1981 by
Becker(j3 and the chemistry of this and related species has been extensively reviewed.64 The structure
of 30 and other compounds RGP (R = H, Me, Ph) have been determined by microwave spec-
troscopy65 whilst the structure of the crystalline compound Ar’GP (31) has been determined by
X-ray crystallography ; 66in all cases, the P=C bond distances are short and consistent with a PGC
triple bond. The arsenic analogue of 31, Ar’C-As, has been reported by Markl67 although no
structural details are available.
Heavier group 14 and 15 elements involving pn-pn multiple bonding 2441

But-CC---P +@
0 CSP

30 31

Finally in this section, it is important to mention a number of cumulenes involving p=C double
bonds. This area, which has been reviewed by Appel,* has developed to include a large range of
compounds, examples of which are 1,3-phospha-allenes RP==C=PR, I-phospha-allenes
RP=C=CR* and phosphaketenes RP=CLO.
(ii) Phospha-imines (P=N). The chemistry of phospha-imines has a slightly longer history than
many of the other classes of compound described in this review, dating back to 1973, and the subject
has recently been reviewed by Niecke. 68Of particu 1a r in terest are the structures of these compounds
(all of which contain at least one bulky group) which can be classified according to three general
classes. The first type, F, involves a trans and planar arrangement about the P=N bond and is
defined by a RP=NR torsion angle of about 180”. The P=N bond distances are all close to 1.55
8, with the angle at phosphorus (a) typically between 100 and 110” and the angle at nitrogen (b)
between 115 and 125”. The second type, G, has a cis configuration for which the RP=NR torsion
angle is close to 0”. Compounds of this type have similar P=N bond lengths to those of type F and
also similar values for the angle ~1,but values for /? are larger falling in the range 130-145”. In the
final type, H, the nitrogen atom adopts a nearly linear coordination in which B ranges from 155 to
180”. An examination of the structural data indicates that there is essentially a continuum between
types G and H in which the angle at nitrogen increases from 130 to 180”. Moreover, this increase
in B correlates with a decrease in the P=N bond length from about 1.55 to 1.48 A. This latter
observation is consistent with the view that type H compounds are best represented as incipient
ionic species of the form [RN=P]+[R]- in which the PN linkage has substantial triple bond
character. Not surprisingly, therefore, type H compounds are found when the [RI- group is electron
withdrawing such as halide, But&O or But2HC0 and complete ionization occurs in the complex
[Ar’N=P]+[AlCl,]- (32). Compound 32 and related species are analogous to organic diazonium
cations and isoelectronic with phospha-alkynes.

R$f,-f,,R
L\
B R

F G

R B
n
‘.
‘\
P;;;N--_R -N- Art]+ [AICI,]

H 32

5.2. Compounds involving double bonds without a first row element

(i) Diphosphenes (P=P) and their heavier congeners. The first stable diphosphene, 1, was reported
in 1981 by Yoshifuji6 and the subject has been reviewed a number of times by Cowley,69 Yoshifuji7’
and Weber. 7’ In all cases, sterically bulky groups on at least one of the phosphorus atoms (mostly
both) are required in order to stabilize these compounds with salient structural features being a
P=P bond length of about 2.03 A, which is 8-10% shorter than typical P-P single bonds, and
angles at the phosphorus atoms which range from 100 to 110” (c.f. phospha-imines). In 1, for
example, the C-P-P angle is 102.8” which is considerably smaller than an angle of 120” which
might be anticipated from a bonding model involving sp2 hybridization at the phosphorus centre.
Such a model is much more appropriate for the diazo analogue of 1, Ar’N=NAr’ (33),72 for which
the bond angles at nitrogen are 113.5” (av.), and for the nitrogen centres in some phospha-imines.
2442 N. C. NORMAN

In fact, a trend wherein the bond angles decrease as the group is further descended is apparent
when diarsenes are considered (see below).
Most of the diphosphenes which have been structurally characterized to date, including the
metalla-diphosphene (q-CSMeJ(C0)2Fe-tiPAr’ (34)73 adopt a truns or E configuration anal-
ogous to 1 and for a long time it was considered that this was the only isomer which was likely to
be observed since a cis or Z form would present no obvious steric barrier to dimerization. The first
indication of a cis diphosphene was obtained by Koenig and co-workers74 from laser irradiation
studies of 1, and also by Yoshifuji et d7’ who isolated the Cr(CO), complex Ar’P=P[Cr(CO)5](mes)
(35), but subsequently, the uncoordinated cis diphosphenes Ar’P=PR (R = Bu’NH, 1-adNH) were
isolated and structurally characterized by Niecke and co-workers who later reported the structural
characterization of both the cis and tram isomers of Ar’P=P{N(SiMe,)[N(SiMe,),l) (36a,b).76

Me,Si,, ,SiMe,
.N
\ 52
,SiMe,

P=P
0

36b

If one considers the stabilization of diphosphenes to be due to steric protection alone, then the
isolation of uncoordinated cis isomers, which are also stable in solution albeit in equilibrium with
the tram form,76 presents a difficulty since, as mentioned above, the attached groups present no
obvious barrier to dimerization. The reason for their stability undoubtedly lies not so much with
steric effects but with electronic factors. For example, in 36b, the nitrogen atom to which one of the
phosphorus atoms is bonded adopts a trigonal planar geometry in which this plane is coplanar with
the NPP plane. There is, therefore, the possibility of conjugation between the nitrogen lone pair and
the P=P bond giving rise to a three-centre, four-electron system which is undoubtedly associated
with their stability. Nevertheless, the consequences of having two large groups cis to each other is
evident from a comparison of the angles at phosphorus in compounds 36a and 36b. In 36a the two
angles N-P-P and P-P-C are 106.1 and 97.6”, respectively whereas in 36b they are 126.3 and
121.4”.
The angle of 97.6” in 36a is one of the smallest observed for a diphosphene, but even smaller
angles are found in diarsenes in which the relevant bond angles are usually a few degrees less than
those in the analogous phosphorus compounds. Two diarsenes (Me3Si)3CAs=AsC(SiMe3)3 (37)77a
and Ar’As=AsCH(SiMe3)z (38)77band a phospha-arsene Ar’P=AsCH(SiMe,), (39)77cprepared by
Cowley and co-workers have been structurally characterized, important angles for which are shown
below and which illustrate the trend to decreasing interbond angles as the group is descended. This
trend to smaller angles on descending the group is not unexpected and can be considered as a special
case of the same trend observed in other EX2 species. The rationale for this effect is similar to that
offered for the increasing pyramidalization in compounds of group 14 and is described in detail in
ref. 43. It is important to note, however, that in the group 15 compounds, the rc-bonding orbitals
are not involved in the orbital mixing associated with the change in geometry. Thus, whereas in
group 14 the increase in pyramidalization leads to a weakening of the E-E bond and a decrease
in the per cent bond shortening (compared to a single bond), in group 15 no such weakening occurs
Heavier group 14 and 15 elements involving pn-pn multiple bonding 2443
and the bond lengths found in phospha-arsenes and diarsenes are, like diphosphenes, g-10% shorter
than typical single bonds.
No structural data is available for an antimony containing compound although spectroscopic
data have been reported for Ar’P=SbCH(SiMe3)2.77c

D+0
?a
EX\
(Me’si)3c\?
CH(SiMeJ2
a
“T\ C(siieJ,

E = P, CI= 101.2, /I = 96.4” E = P, a,/? = 108.5”(av.)


E = As, ti = 99.9, /? = 93.6” E = As, d~,b= 106.3”(av.)

6. CONCLUSION AND GENERAL FEATURES

In concluding this perspective, it will be useful to recapitulate briefly the salient features associated
with the multiply bonded compounds described herein. In the early years of the subject there was
much discussion as to whether “true” double bonds were really present. In most cases, the answer
to this question must now be seen to be yes inasmuch as the spectroscopic and structural features
are those expected for multiply bonded species. For example, in all compounds where a suitable
NMR nucleus is available, the chemical shifts and coupling constants are usually consistent with
the presence of a multiple bond ; the solid state NMR studies of Zilm and co-workers78 are a
particularly elegant example in this regard.
Moreover, the structural studies on which this report has focused are also consistent with
multiple bonding in that percentage bond shortenings of between 8 and 10% (for double bonds)
are commonly observed in conjunction with a planar or near planar environment around the E-E
unit. The only major exceptions are found for the digermenes and distannenes (and other tin
containing species) in which the standard r~+ rcmodel is clearly not appropriate in most cases. Some
of the structural features of the silanimines and the phospha-imines are also unusual. Nevertheless,
we must exercise a little caution in the discussions of bond lengths particularly with respect to the
percentage shortenings observed. Thus, in an insightful paper, Power et ~1.~~have shown that
comparisons between bond lengths must take due account of the formal hybridizations of the
elements involved. For example, typical P-P single bonds in compounds such R2P-PR2 are about
2.2 A whereas the P=P bonds in diphosphenes are about 2.03 A, a shortening of about 8%. However,
the phosphorus hybridization in diphosphines and diphosphenes is sp3 and sp’, respectively and
Power has estimated that a bond shortening of about 0.1 A (i.e. about 4%) is due to this change in
hybridization alone. It is therefore likely that only half of the shortening observed for P=P bonds
is due to the presence of the n-bond, the other half being due simply to the different hybridization
of the phosphorus atoms. Analogous examples from nitrogen chemistry, however, suggest that the
shorter distances found for N=N double bonds vs N-N single bonds are mostly due to the presence
of the n-bond (about 75%) and less to the different hybridization of the nitrogen (25%). These
observations are consistent with the relatively stronger x-bonding expected for nitrogen vs phos-
phorus as outlined earlier. I
Finally, it is worth mentioning some of the types of compound which have yet to be synthesized.
Clearly there is a paucity of data for species involving the heaviest elements, particularly the 6p
elements lead and bismuth although it is likely that even more sterically bulky ligands will be needed
to afford stability in these cases. An alternative strategy is to exploit the Lewis acidity of these
elements and to use groups which are both bulky and which contain Lewis basic sites in order to
form intramolecular acid-base complexes. However, the concomitant increase in coordination
number means that a description of the bonding in terms of a pure pn--pn: interaction is no longer
entirely appropriate.
In the chemistry of the group 14 elements, the isolation of triply bonded compounds has so far
eluded researchers in the field. Thus whilst C&C and C=N bonds are well characterized, there are
no stable examples of sila-alkynes, RSi=LR, silanitriles, RSi=N, disilyenes, RSi=SiR or analogous
2444 N. C. NORMAN
compounds involving the heavier congeners. This is in contrast to group 15 where the chemistry
of phospha-alkynes is now well established and this probably reflects the fact that phosphorus forms
somewhat more stable n bonds than silicon due to its smaller size. Success in the field of heavier
element triple bonds is also likely to require new and even more bulky ligands, particularly those
with the ability to stabilize what are likely to be extremely electrophilic silicon centres. It is probable
that conjugatively stabilized compounds and transition metal complexes will be prepared before
uncoordinated compounds are isolated as has happened before in this field.

Note added at proof

Since the submission of the manuscript, X-ray crystal structures have been reported for the stannaimine
complex {(Me$i)zN}zSn=N(C6H3Pr’2-2,6)so and for the phosphasilene complex (CgHzBufx-2,4,6)P=Si
(Bu’){P(C,H,Bu’,-2,4,6)(PPh,)).8’

Acknowledgements-I would like to thank Drs Paul D. Lickiss and A. Guy Orpen and Professor Alan H.
Cowley FRS for their comments and advice in preparing this manuscript.

REFERENCES
1. W. E. Dasent, Nonexistent Compounds. Marcel Dekker, New York (1965).
2. H. Kohler and A. Michaelis, Ber. Dtsch. Chem. Ges. 1877, 10, 807; J. J. Daly and L. Maier, Nature
(London) 1964,203,1167 and 1965 ; 208, 383 ; J. J. Daly, J. Chem. Sot. 1964,6147 and 1965,4789.
3. P. Jutzi, Angew. Chem. Int. Ed. Engl. 1975, 14, 232; L. E. Gusel’nikov and N. S. Nametkin, Chem. Rev.
1979,79, 529. See also K. S. Pitzer, J. Am. Chem. Sot. 1948,70, 2140; R. S. Mulliken, J. Am. Chem. Sot.
1950,72,4493 and 1955,77,884.
4. E. H. Appelman, Act. Chem. Res. 1973, 6, 113.
5. P. Laszlo and G. J. Schrobilgen, Angew. Chem. Int. Ed. Engl. 1988,27,479.
6. M. Yoshifuji, I. Shima, N. Inamoto, K. Hirotsu and T. Higuchi, J. Am. Chem. Sot. 1981, 103,4587.
7. R. West, M. J. Fink and J. Michl, Science 1981, 214, 1343.
8. R. Appel and F. Knoll, Adv. Znorg. Chem. 1989,33,259 ; R. Appel, in Multiple Bon& and Low Coordination
in Phosphorus Chemistry (Edited by M. Regitz and 0. J. Scherer). Georg Thieme Verlag, Stuttgart (1990).
9. A. G. Brook and K. M. Baines, Adv. Organomet. Chem. 1986,2&l ; G. Raabe and J. Michl, Chem. Rev.
1985,85,419; G. Raabe and J. Michl, in The Chemistry of Organic Silicon Compounds (Edited by S. Patai
and Z. Rappoport). Wiley, Chichester (1989).
10. W. Kutzelnigg, Angew. Chem. Int. Ed. Engl. 1984, 23, 272.
11. L. Pauling, The Nature of the Chemical Bond. Cornell University Press, Ithaca, New York (1960).
12. A. Moezzi, M. M. Olmstead and P. P. Power, J. Chem. Sot., Dalton Trans. 1992, 2429; A. Moezzi, M.
M. Olmstead, R. A. Bartlett and P. P. Power, Organometallics 1992, 11,2383and refs. therein.
13. (a) W. Uhl, Z. Naturforsch., Teil B 1988,43, 1113 ; (b) W. Uhl, M. Layh and T. Hildenbrand, J. Organomet.
Chem. 1989,364,289; (c) W. Uhl, M. Layh and W. Hiller, J. Organomet. Chem. 1989,368, 139.
14. R. D. Schluter, A. H. Cowley, D. A. Atwood, R. A. Jones, M. R. Bond and C. J. Carrano, J. Am. Chem.
Sot. 1993, 115,207O.
15. A. G. Brook, S. C. Nyburg, F. Abdesaken, B. Gutekunst, G. Gutekunst, R. K. M. R. Kallury, Y. C.
Poon, Y.-M. Chang and W. Wong-Ng, J. Am. Chem. Sot. 1982,104,5667. See also K. M. Baines, A. G.
Brook, R. R. Ford, P. D. Lickiss, A. K. Saxena, W. J. Chatterton, J. F. Sawyer and B. A. Behnam,
Organometallics 1989, 8, 693.
16. (a) N. Wiberg, G. Wagner and G. Mtiller, Angew. Chem., Znt. Ed. Engl. 1985, 24, 229; (b) N. Wiberg, G.
Wagner, G. Miiller and J. Riede, J. Organomet. Chem. 1984, 271, 381.
17. P. D. Lickiss, Chem. Sot. Rev. 1992, 25, 271.
18. J. Barrau, J. Escudie and J. Satge, Chem. Rev. 1990, W, 283.
19. M. Lazraq, J. Escudie, C. Couret, J. Satge, M. DrHger and R. A. Dammel, Angew. Chem. Int. Ed. Engl.
1988, 27, 828.
20. H. Meyer, G. Baum, W. Massa and A. Berndt, Angew. Chem. fnt. Ed. Engl. 1987,26,798.
21. A. Berndt, H. Meyer, G. Baum, W. Massa and G. Berger, Pure Appf. Chem. 1987,59, 1011.
22. G. Anselme, H. Ranaivonjatovo, J. Escudie, C. Couret and J. Satje, OrganometaNics 1992, 11, 2748.
23. H. Meyer, G. Baum, W. Massa, S. Berger and A. Berndt, Angew. Chem. Znt. Ed. Engl. 1987,26, 546.
24. H. Griitzmecher, S. Freitag, R. Herbst-Irmer and G. M. Sheldrick, Angew. Chem. Znt. Ed. Engl. 1992,
31, 437.
25. N. Wiberg, K. Schurz, G. Reber and G. Miiller, J. Chem. Sot., Chem. Commun. 1986, 591.
26. N. Mitzel, A. Schier and H. Schmidbaur, Chem. Ber. 1992, 125, 2711 and refs. therein.
Heavier group 14 and 15 elements involving pa---pn multiple bonding 2445

27. N. Wiberg, K. Schurz, G. Mtiller and J. Riede, Angew. Chem. Znt. Ed. Engl. 198827,935.
28. S. Walter, U. Klingebiel and D. Schmidt-Base, J. Orgunomet. Chem. 1991,412, 319.
29. A. Meller, G. Ossig, W. Maringgele, D. Stalke, R. Herbst-Irmer, S. Freitag and G. M. Sheldrick, J. Chem.
Sot., Chem. Commun. 1991, 1123.
30. R. West, Angew. Chem. Znt. Ed. Engl. 1987,26, 1201; R. West, Pure Appl. Chem. 1984,56, 163.
31. M. J. Fink, M. J. Michalczyk, K. J. Haller, R. West and J. Michl, Orgunometallics 1984,3, 793.
32. S. Masamune, S. Murakami, J. T. Snow, H. Tobita and D. J. Williams, Organometullics 1984,3,333.
33. R. S. Archibald, Y. van den Winkel, A. J. Millevolte, J. M. Desper and R. West, Organometalhcs 1992,
11, 3276.
34. H. Watanabe, K. Takeuchi, N. Fukawa, M. Kato, M. Goto and Y. Nagai, Chem. Lett. 1987,134l.
35. B. D. Shepherd, D. R. Powell and R. West, Organometallics 1989,8,2664.
36. S. Masamune, Y. Eriyama and T. Kawase, Angew. Chem. Znt. Ed. Engl. 1987,26, 584.
37. M. Weidenbruch, B. Flintjer, S. Pohl and W. Saak, Angew. Chem. Znt. Ed. Engl. 1989,28,95.
38. J. T. Snow, S. Murakami, S. Masamune and D. J. Williams, Tetrahedron Lett. 1984, 25,419l.
39. S. A. Batcheller, T. Tsumuraya, 0. Tempkia, W. M. Davis and S. Masamune, J. Am. Chem. Sot. 1990,
112,9394.
40. D. E. Goldberg, P. B. Hitchcock, M. F. Lappert, K. M. Thomas, T. Fjeldberg, A. Haaland and B. E. R.
Schilling, J. Chem. Sot., Dalton Trans. 1986, 2387.
41. L. Pauling, Proc. Natl. Acad. Sci. U.S.A. 1983,80, 3871.
42. (a) G. Trinquier and J.-P. Malrieu, J. Am. Chem. Sot. 1987, 109, 5303 ; (b) J.-P. Malrieu and G. Trinquier,
J. Am. Chem. Sot. 1989,111, 5916; (c) G. Trinquier, J. Am. Chem. Sot. 1990,112,2130.
43. T. A. Albright, J. K. Burdett and M. H. Whangbo, Orbital Interactions in Chemistry. Wiley, New York
(1985).
44. (a) C. N. Smit, F. M. Lock and F. Bickelhaupt, Tetrahedron Lett. 1984, 25, 3011; (b) C. N. Smit and F.
Bickelhaupt, Orgunometallics 1987,6, 1156 ; (c)Y. van der Winkel, H. M. M. Bastiaans and F. Bickelhaupt,
J. Organomet. Chem. 1991,405, 183.
45. M. Drlger, J. Escudie, C. Couret, H. Ranaivonjatovo and J. Satge, OrgunometaZIics 1988,7, 1010.
46. C. Couret, J. Escudie, J. Satge, A. Rahainirina and J. D. Andriamizaka, J. Am. Chem. Sot. 1985, 107,
8280.
47. K. Hassler, F. Mitter and B. Reiter, J. Organomet. Chem. 1989, 376, Cl.
48. M. Driess and H. Pritzkow, Angew. Chem. Znt. Ed. Engl. 1992, 31, 316. See also M. Driess, H. Pritzkow
and M. Sander, Angew. Chem. Znt. Ed. Engl. 1993,32,283.
49. P. Arya, J. Boyer, F. Car& R. Corriu, G. Lanneau, 9. Lapasset, M. Perrot and C. Priou, Angew. Chem.
Znt. Ed. Engl. 1989, 28, 1016.
50. R. Corriu, G. Lanneau and C. Priou, Angew. Chem. Znt. Ed. Engl. 1991,30, 1130.
51. M. Veith, S. Becker and V. Huch, Angew. Chem. Znt. Ed. Engl. 1989, 28, 1237.
52. A. G. Davies and A. G. Neville, J. Organomet. Chem. 1992,436,255.
53. G. Becker, Z. Anorg. Allg. Chem. 1976,423,247.
54. T. A. van der Knaap, L. Jenneskens, H. J. Meenwissen, F. Bickelhaupt, D. Walther, E. Dinjus, E. Uhlig
and A. L. Spek, J. Organomet. Chem. 1983,254, C33.
55. A. H. Cowley, R. A. Jones, J. G. Lasch, N. C. Norman, C. A. Stewart, A. L. Stuart, J. L. Atwood, W. E.
Hunter and H.-M. Zhang, J. Am. Chem. Sot. 1984,106,7015.
56. (a) R. Appel, J. Menzel, F. Knoch and P. Volz, Z. Anorg. Allg. Chem. 1986, 534, 100; (b) M. Yoshifuji,
K. Toyota, K. Shibayama and N. Inamoto, Chem. Lett. 1983. 1153 ; (c) M. Yoshifuji, K. Toyota and N.
Inamoto, Tetrahedron Lett. 1985, 26, 1727.
57. (a) A. J. Ashe, J. Am. Chem. Sot. 1971, 93, 3293; (b) G. Markl, Angew. Chem. Znt. Ed. Engl. 1966, 5,
846.
58. R. Appel and A. Westerhaus, Tetrahedron Lett. 1981,22, 2159.
59. D. Gudat, E. Niecke, A. M. Arif, A. H. Cowley and S. Quashie, Organometallics 1986,5, 593.
60. L. Weber, K. Reizig, R. Boese and M. Polk, Organometallics 1986, 5, 1098.
61. A. M. Arif, A. H. Cowley, C. M. Nunn, S. Quashie, N. C. Norman and A. G. Orpen, Organometallics
1989,8, 1878.
62. (a) L. Weber, G. Meine and R. Boese, Angew. Chem. Znt. Ed. EngI. 1986,25,469; (b) L. Weber, G. Meine,
R. Boese and D. Bungardt, Z. Anorg. Allg. Chem. 1987,549,73.
63. G. Becker, G. Gresser and W. Uhl, Z. Naturforsch., Teil B, 1981, 36, 16.
64. (a) M. Regitz, in Muhiple Bonds and Low Coordination in Phosphorus Chemistry (Edited By M. Regitz
and 0. J. Scherer). Georg Thieme Verlag, Stuttgart (1990); (b) P. Binger, ibid; (c) M. Regitz, Chem. Rev.
1990,90, 191; (d) J. F. Nixon, Chem. Rev. 1988,88, 1327.
65. H. W. Kroto, Chem. Sot. Revs. 1982, 11,435.
66. A. M. Arif, A. R. Barron, A. H. Cowley and S. W. Hall, J. Chem. Sot., Chem. Commun. 1988, 171 and
refs therein.
2446 N. C. NORMAN

67. G. Mark1 and H. Sejpka, Angew. Chem. Znt. Ed. Engl. 1986,25,264.
68. E. Niecke, in Multiple Bonds and Low Coordination in Phosphorus Chemistry (Edited by M. Regitz and
0. J. Scherer). Georg Thieme Verlag, Stuttgart (1990). See also E. Niecke, R. Detsch, M. Nieger, F.
Reichert and W. W. Schoeller, Bull. Sot. Chim. Fr. 1993, 130, 25 and refs therein.
69. (a) A. H. Cowley, J. E. Kilduff, J. G. Lasch, S. K. Mehrotra, N. C. Norman, M. Pakulski, B. R. Whittlesey,
J. L. Atwood and W. E. Hunter, Znorg. Chem. 1984,23,2582 ; (b) A. H. Cowley and N. C. Norman, Prog.
Znorg. Chem. 1986,34,1; (c) A. H. Cowley, Act. Chem. Res. 1984,17,386; (d) A. H. Cowley, Polyhedron
1984,3,389; (e) A. H. Cowley, J. Organomet. Chem. 1990,400,71.
70. M. Yoshifuji, in Multiple Bonds and Low Coordination in Phosphorus Chemistry (Edited by M. Regitz and
0. J. Scherer). Georg Thieme Verlag, Stuttgart (1990).
71. L. Weber, Chem. Rev. 1992,92, 1839.
72. Y. LePage, E. J. Gabe, Y. Wang, L. R. C. Barclay and H. L. Holm, Acta Cryst., Sect. B 1980,36,2846.
73. L. Weber and K. Reizig, Angew. Chem. Znt. Ed. Engl. 198524,865.
74. A.-M. Caminade, M. Verrier, C. Ades, N. Paillous and M. Koenig, J. Chem. Sot., Chem. Commun. 1984,
875.
75. M. Yoshifuji, T. Hashida, N. Inamoto, K. Hirotsu, T. Horiuchi, T. Higuchi, K. Ito and S. Nagase, Angew.
Chem. Znt. Ed. Engl. 1985,24,211.
76. (a) E. Niecke, B. Kramer and M. Nieger, Angew. Chem. Znt. Ed. Engl. 1989, 28, 215; (b) E. Niecke, 0.
Altmeyer and M. Nieger, Angew. Chem. Znt. Ed. Engl. 1991,30, 1136.
77. (a) A. H. Cowley, N. C. Norman and M. Pakulski, J. Chem. Sot., Dalton Trans. 1985, 383 ; (b) A. H.
Cowley, J. G. Lasch, N. C. Norman and M. Pakulski, J. Am. Chem. Sot. 1983, 105, 5506; (c) A. H.
Cowley, J. G. Lasch, N. C. Norman, M. Pakulski and B. R. Whittlesey, J. Chem. Sot., Chem. Commun.
1983,881.
78. (a) K. W. Zilm, G. G. Webb, A. H. Cowley, M. Pakulski and A. Orendt, J. Am. Chem. Sot. 1988,110,
2032; (b) J. C. Duchamp, M. Pakulski, A. H. Cowley and K. W. Zilm, J. Am. Chem. Sot. 1990, 112,
6803.
79. D. C. Pestana and P. P. Power, Znorg. Chem. 1991,X), 528.
80. G. Ossig, A. Meller, S. Freitag and R. Herbst-Irmer, J. Chem. Sot., Chem. Commun. 1993,497.
81. H. R. G. Bender, E. Niecke and Nieger, J. Am. Chem. Sot., 1993, 115,3314.

You might also like