You are on page 1of 10

View Online / Journal Homepage / Table of Contents for this issue

PAPER www.rsc.org/dalton | Dalton Transactions

Synthesis, structure and properties of {M4 O4 } cubanes containing nickel(II)


and cobalt(II)†‡
Katharina Isele,a Fabienne Gigon,a Alan F. Williams,*a Gérald Bernardinelli,b Patrick Franzc and
Silvio Decurtinsc
Received 4th October 2006, Accepted 6th November 2006
First published as an Advance Article on the web 21st November 2006
DOI: 10.1039/b614424a

A survey of the crystal structures containing simple {M4 O4 } cubane units is reported. It shows that the
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

average M–M distance in these complexes is relatively constant for a given metal ion M. The structures
are all distorted from the idealised cube to a T d structure, and most show a further distortion which,
however, usually maintains some elements of symmetry. A system for classifying the different types of
ligand in these complexes is proposed. Two new cubanes of cobalt(II) and nickel(II) with the ligand
(R,R)-bis-1,2-(1-methylbenzimidazol-2-yl)ethane-1,2-diol, (R,R)-1 or its enantiomer have been isolated
and the crystal structure of the cobalt(II) complex confirms the cubane structure. Electronic, CD and
Downloaded by Duke University on 28 July 2012

1
H NMR spectra and magnetic susceptibility data are reported. The magnetic data for these and other
compounds in the literature are discussed in terms of the structural parameters.

Introduction
Coordination chemistry is increasingly interested by the study
of polynuclear complexes and the novel properties associated
with the interactions of two or more metal centres.1 Interest has
been focussed on complexes where magnetic exchange is observed
between paramagnetic metal ions, both from a theoretical point of
Scheme 1
view, and also with the aim of developing single molecule magnets.2
We have been interested in the design of ligands which favour the
formation of polynuclear species and which may potentially con- The cubanes, which we shall define in this paper as complexes
trol the structure. The ligand (R,R)-bis-1,2(1-methylbenzimidazol- with a M4 X4 core in which the metal ions occupy four non-
2-yl)ethane-1,2-diol, (R,R)-1 (Scheme 1) possesses potentially neighbouring vertices of a cube, and the triply bridging X units
bridging alcohol functions which may be deprotonated to give the four others (Scheme 2), are a particularly interesting family.
alkoxo-functions which are good bridging ligands, and offer a The Fe4 S4 cubanes are well known in biology as electron transfer
pathway for magnetic exchange.3 We will denote the deprotonated agents.5,6 Very recently, a heteronuclear Mn3 CaO4 cubane was
forms as 1 − H or 1 − 2H. The ligand and its enantiomer (S,S)- shown to be present at the active site of the oxygen evolving
1 are readily prepared in enantiomerically pure form from the centre of photosystem II,7 and a Fe3 MoS4 cubane has been
appropriate isomer of tartaric acid. identified at the active site of nitrogenase.8 M4 O4 units show
We have previously reported on the chemistry of this ligand interesting magnetic exchange properties and may under certain
in simple mononuclear complexes [M(1)2 ]n+ 4 and more recently circumstances act as single molecule magnets.9–13
on the various polynuclear complexes formed with copper(II).3
In this paper we show that ligand 1 forms cubane complexes
with cobalt(II) and nickel(II) and report on the properties of these
complexes.

a
Department of Inorganic chemistry, University of Geneva, 30 quai Scheme 2
Ernest-Ansermet, CH-1211, Geneva, Switzerland. E-mail: Alan.Williams@
chiam.unige.ch
b
Laboratory of X-ray Crystallography, University of Geneva, 4 quai Ernest- Although many such cubane complexes are now known, there
Ansermet, CH-1211, Geneva 4, Switzerland has been little comparative discussion of the structures. We
c
Department of Chemistry and Biochemistry, Universität Bern, Freiestrasse therefore review briefly the structures in the Cambridge Structural
3, CH-3009, Berne, Switzerland Database for cobalt and nickel, and propose a simple structural
† Dedicated to Professor Hans Güdel on the occasion of his retirement.
classification which may be used for correlation with the properties
‡ Electronic supplementary information (ESI) available: IR spectra of
[Co4 ((R,R)-1 − H)4 ](ClO4 )4 (C2 H6 OH)9 and [Ni4 ((S,S)-1 − H)4 ](ClO4 )4 - of the complexes. To simplify discussion of our results, we begin
(H2 O)5 . See DOI: 10.1039/b614424a with this survey.

332 | Dalton Trans., 2007, 332–341 This journal is © The Royal Society of Chemistry 2007
View Online

Results cobalt(III). This constancy of the average M–M distance may be


interpreted in terms of a well-defined M4 O4 unit which may be
Structural analysis of cubanes distorted slightly but which is not greatly affected by the external
ligands.
The term cubane has been used to describe many complexes which
show great structural variety. In this section we discuss the features Distortion modes. The M–M distances afford a simple method
which allow one to distinguish and classify the structures. of analysing the distortion of the M4 tetrahedron (Scheme 3). If the
tetrahedron is distorted by extension along a S4 axis, four M–M
Metric properties. An ideal cubane has twelve equal bond
distances will be increased, and two others will remain the same.
lengths and twenty-four bond angles of 90◦ . Such ideality is
Distortion along a C 3 axis will give two sets of three equal M–
not encountered in practice, notably because three bonds at
M distances, whereas a distortion which maintains three C 2 axes
90◦ are rarely observed at the bridging oxygen atom. An sp3
will give three pairs of M–M distances. The D2 symmetry may
hybridised oxygen atom would have three angles of 109.5◦ , and
be further distorted to a C 2 system. This very simple approach
this may be achieved by compressing the four oxygens along the
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

accounts for the majority of the compounds in the database: out


body diagonals of the cube, increasing the angles at oxygen, but
of 69 compounds, 40 show a S4 distortion, 14 a D2 or C 2 distortion
concurrently decreasing the angles at the metal below 90◦ . We may
and 5 a C 3 distortion, the remaining 10 adopting lower symmetry.
measure the magnitude of this effect by the average M–O–M bond
The magnitude of the distortion may be estimated roughly from
angle or by calculating the sum of the angles at an oxygen vertex,
the standard deviation of the six M–M distances: we have taken
the value for an ideal cubane being 270◦ and that for a sp3 oxygen
values below 0.03 Å to correspond to a low distortion, and those
328.5◦ . As a consequence of this distortion, the M4 O4 units may
greater than 0.1 Å as highly distorted. Table 2 shows mean M–M
Downloaded by Duke University on 28 July 2012

be regarded as interpenetrating M4 and O4 tetrahedra, the whole


distances for the 20 cobalt(II) cubanes in the CSD, and, for a given
possessing T d symmetry. We will refer to this distortion of the ideal
compound, the standard deviation of the M–M distance, and the
cube as the tetrahedral distortion.
deviation of each M–M distance from the mean, which allows
We have searched the Cambridge Structural Database (version
the ready identification of the distortion mode. We have taken
5.27 with updates of January 2006) for structures containing
extensions as shown in Scheme 3 to be positive; a negative S4
the M4 O4 unit where M = Co or Ni, excluding structures with
distortion corresponds to a compression and will give four bonds
errors. We also excluded more complicated structures such as
shorter than average and two longer.
face-sharing cubanes where it appeared that the geometry of
the M4 O4 unit was likely to be perturbed by the oligomerisation.
This left a total of 70 structures, 40 with Ni(II), 20 with Co(II), one
mixed valence structure with two Co(II) and two Co(III), and 9 with
Co(III). A summary of the most important parameters is given in
Table 1. In all cases the tetrahedral distortion discussed above is
observed: the average M–O–M angles are greater than 90◦ , and
the O–M–O angles less than 90◦ . Although the individual angles
vary by several degrees, the sum of the angles at a given oxygen
vertex shows a much smaller variation.
The tetrahedron of metal ions may be characterised by the
six M–M distances which will be equal in an ideal tetrahedral
structure. If we examine these distances we observe that the
individual values vary by up to 0.3 Å within a given structure
(i.e., by about 10%), but from one structure to another the mean
Scheme 3 The M4 tetrahedron of the cubane (shown with dashed lines,
value is remarkably constant for a given metal ion, with a standard top) may be distorted in three ways while retaining some element of
deviation of the order of 0.03 Å. The average values for cobalt(II) tetrahedral symmetry. For the D2 distortion only one C 2 axis is shown.
and nickel(II) are similar but that for cobalt(III) is much smaller.
This is of course a consequence of the shorter M–O bond distances For example, the structure UMAHAR14 shows four bonds
expected for cobalt(III). The average M–M distances for the mixed longer than the average and two which are shorter, corresponding
valence cobalt cubane lie in between the values for cobalt(II) and to an elongation along the S4 axis or positive distortion, while

Table 1 Average structural parameters for cobalt and nickel cubanes

Co(II) Co(III) Ni(II)

Number of compounds 20 9 40
Mean M–M distance/Å 3.14(3) 2.80(3) 3.11(3)
Maximum and minimum M–M distances/Å 3.368, 2.854 3.060, 2.641 3.269, 2.824
Mean M–O distance/Å 2.094(15) 1.885(12) 2.064(11)
Mean M–O–M angle/◦ 97.2(7) 95.9(6) 97.6(6)
Mean pyramidal angle/◦ 291.8 287.8 292.9
Mean O–O distance/Å 2.747(17) 2.510(11) 2.698(18)
Mean O–O/M–M ratio 0.875 0.898 0.869

This journal is © The Royal Society of Chemistry 2007 Dalton Trans., 2007, 332–341 | 333
View Online

Table 2 Data for cobalt(II) cubanes in the CSD. Distortion: magnitude of the distortion based on the s.d. of the mean Co–Co distance (low < 0.03 Å;
high > 0.1 Å). Symmetry: approximate symmetry of the distortion from T d . Dn : the difference between the nth Co–Co distance of the cubane and the
mean value. Tet. Parm: ratio of the average O–O distance to the average M–M distance. Ligand coding as described in the text

Co–Co Distortion, Tet.


CSD code mean/Å s.d./Å symmetry D1 /Å D2 /Å D3 /Å D4 /Å D5 /Å D6 /Å Parm. Ligand coding Ref.

TAWJEG 3.143 0.035 Medium −0.015 0.029 −0.052 −0.012 0.002 0.048 0.871 Co4 /0s4 /m10 b 24
UMAHAR 3.103 0.144 High, S4 + −0.153 0.097 0.084 0.091 0.097 −0.215 0.886 Co4 /0s4 /m4 b2 d2 14
XEZMUK 3.133 0.088 Medium, C 3 − 0.081 0.081 −0.080 0.081 −0.080 −0.080 0.879 Co4 /0s4 /m6 b3 16
XOVNAX 3.118 0.005 Low, C 3 + 0.005 0.004 0.004 −0.005 −0.005 −0.005 0.883 Co4 /0s4 /m6 b3 25
ACABEL 3.134 0.042 Medium, S4 − −0.013 −0.030 0.033 0.069 −0.031 −0.030 0.885 Co4 /0s4 /m4 b4 26
BEWDUC 3.128 0.110 High, C 2 0.018 −0.095 0.201 −0.044 −0.095 0.018 0.883 Co4 /0s4 /m4 b4 27
CEVSEB 3.109 0.046 Medium, S4 − −0.030 0.059 −0.030 −0.030 0.059 −0.030 0.884 Co4 /0s4 /m4 b4 28
MAKFAG 3.132 0.046 Medium, S4 − −0.031 0.052 −0.023 −0.037 0.065 −0.028 0.879 Co4 /0s4 /m4 b4 21
OBOYOT 3.101 0.025 Low, D2 −0.011 0.027 −0.020 −0.020 0.036 −0.011 0.876 Co4 /0s4 /m4 b4 29
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

XOVNEB 3.118 0.021 Low, S4 − 0.027 −0.014 −0.014 −0.014 −0.014 0.027 0.881 Co4 /0s4 /m4 b4 25
HAJCIF 3.099 0.147 High, S4 + −0.139 0.088 0.099 0.095 0.089 −0.234 0.888 Co4 /0s4 /m4 d2 b2 30
HAJFAA 3.099 0.126 High, S4 + 0.090 0.078 −0.169 −0.155 0.083 0.073 0.892 Co4 /0s4 /m4 d2 b2 30
IWOKOU 3.147 0.105 High, D2 −0.144 0.039 0.094 0.095 0.036 −0.119 0.864 Co4 /0s4 /d2 b4 17
RAPXIQ 3.062 0.140 High, S4 + −0.150 0.089 0.083 0.095 0.092 −0.208 0.886 Co4 /0s4 /m4 d2 b2 31
LILTUV 3.147 0.119 High, D2 −0.136 0.131 0.007 0.007 0.129 −0.136 0.860 Co4 /0s4 /d332 23
FUXNOB 3.165 0.053 Medium −0.043 −0.019 −0.026 −0.017 0.002 0.103 0.874 Co4 /1s4 /m6 32
UMUFOX 3.188 0.133 High, S4 − 0.180 −0.098 −0.085 −0.094 −0.066 0.163 0.880 Co4 /1s4 /d4 15
Downloaded by Duke University on 28 July 2012

JAXTOR 3.241 0.012 Low 0.002 0.011 0.002 −0.015 −0.013 0.016 0.848 Co4 /2s4 /m4 22
MAWZOZ 3.242 0.029 Low, C 2 −0.034 0.037 −0.022 −0.013 0.030 0.004 0.842 Co4 /2s4 /m4 20
ILUHUS 3.173 0.036 Medium, C 2 0.014 −0.013 0.043 0.034 −0.032 −0.046 0.861 Co4 /3s4 / 13
[Co4 (1 − H)4 ]4+ 3.229 0.121 High, C 2 0.042 −0.059 0.128 0.128 −0.059 −0.179 0.864 Co4 /2s212 3s2 /

UMUFOX15 has two extended and four shortened bonds, corre- more significant change is when the bridging oxygen atoms are
sponding to a flattening of the tetrahedron or negative distortion. part of alkoxo- or phenoxo-units which are linked to one or more
XEZMUK16 with three bonds shortened and three lengthened has terminal ligands. We will refer to this arrangement as a supported
a negative C 3 distortion, and IWOKOU17 a D2 structure with three bridging ligand, using a term which has previously been used in
different pairs of roughly equal distances. Tables for cobalt(III) and the chemistry of polynuclear copper compounds.3 In this case the
nickel(II) compounds are given in the supplementary material. The different binding sites may be prearranged to favour the formation
S4 distortion has been mentioned previously in studies18 of iron- of the cubane structure, although to our knowledge there are
sulfur protein models where it is often observed18,19 (note that these currently no thermodynamic data which allow one to judge how
authors refer to it as a D2d distortion). great a stabilisation of the polynuclear species is achieved.
The bridging oxygen may be linked to one, two, or three
Oxygen–oxygen distances. Now we may examine the O4 tetra-
metal ions, and we will call these mono-, di- or ter-supported
hedron. The tetrahedral distortion of the cube brings the oxygen
(Scheme 4). An example of a monosupported ligand is 2-
atoms closer to one another, and one may ask to what extent O–
methylquinolin-8-olate (UMUFOX15 ), of a disupported ligand
O repulsion is important. For Ni(II) and Co(II) the average O–O
bis(2-pyridyl)hydroxymethanolate (MAWZOZ20 ), and of a tersup-
distances are similar and close to 2.7 Å, but are slightly smaller for
ported ligand citrate (ILUHUS13 ).
cobalt(III) at 2.51 Å. These values are smaller than twice the van
der Waals radius of oxygen (1.40 Å) and comparable to the values
commonly used for the ionic radius of oxygen. We may therefore
deduce that some O–O interaction is present. The ratio of the
mean O–O distance to the mean M–M distance may be taken as a
measure of the tetrahedral distortion: a value of 1 is expected for
a perfect cube. The values vary only by a few percent for a given
group of compounds (see Table 2 for the cobalt(II) complexes), but
Scheme 4 Classification of bridging ligands.
are significantly greater for the cobalt(III) complexes.
If a mixed valence MII 2 MIII 2 O4 system is considered, we may
One quickly becomes aware of the difficulty of comparing struc-
easily predict that the intermetal distances will follow the or-
tures with different ligands. Often the connectivity only becomes
der MIII –MIII < MIII –MII < MII –MII leading to a C 2 system which
clear after rotating a computer-generated image so as to examine
is indeed observed. For MII MIII 3 O4 and MII 3 MIII O4 one would
the complex from different angles. We suggest here a relatively
predict negative and positive C 3 distorted structures respectively.
simple nomenclature for coding the different types of ligands in
Ligand description. For an octahedral metal ion, the simplest these structures. Any given structure may be regarded as composed
cubane will have the composition [L3 M(OR)]4 where OR is the of four metals and four bridging atoms (in our case oxygen) which
l3 -bridging ligand, and L is a monodentate terminal ligand which form the vertices of the cube. The coordination sphere of the
completes the octahedral coordination of the metal. This structure metal is then completed by a certain number of terminal ligands.
may be modified in several ways. A trivial change is to replace the We denote the terminal ligands as m (monodentate), b (bidentate)
monodentate terminal ligands by bidentate or tridentate ones. A or t (tridentate). The bridging atoms may be unsupported (0s) or

334 | Dalton Trans., 2007, 332–341 This journal is © The Royal Society of Chemistry 2007
View Online

supported (1s, 2s, or 3s, the number designating the number Co4 O4 cubane structure with four deprotonated alcohol functions
of arms linking the bridging atom to a terminal ligand). We of ligand (R,R)-1 supplying the bridging oxygens. Each cobalt
then code the complex as a whole by metal4 /bridging ligand4 / is further coordinated by two benzimidazole functions and an
terminal ligandsn . Thus the simplest cubane [L3 M(OR)]4 is M4 / alcohol function from each ligand. The two crystallographically
0s4 /m12 , involving four unsupported (0s) OR groups and twelve distinct ligands show different coordination modes. Both possess
terminal monodentate ligands m. Substitution of two terminal a triply bridging alkoxide function, but ligand a binds one metal
monodentate ligands on each metal by a bidentate ligand gives through a benzimidazole and an alcohol, and a second through a
M4 /0s4 /m4 b4 . This is a common motif as in MAKFAG21 benzimidazole; ligand b binds all three metal ions bridged by the
[Co(OMe)(dpm)(MeOH)]4 which has bridging methoxides (0s) alkoxide using a benzimidazole, an alcohol and a benzimidazole
and the terminal ligands of each metal are an dipivaloylmethane respectively. Using the nomenclature developed above, ligand a
(dpm, type b) and a methanol (m). When supported ligands are may be described as 2s21, and ligand b as 3s, giving a complete
present, they occupy some of the terminal sites. For example structural description as Co4 /2s212 3s2 /.
JAXTOR22 contains a di-supported bis(2-pyridyl)hydroxy-
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

methanolate 2s ligand where the two pyridine nitrogens occupy


two of the terminal sites, the third being occupied by a
monodentate acetate to give Co4 /2s4 /m4 .
There are two subtleties which need to be taken into account.
There is sometimes a terminal ligand which binds to more than
one metal. For example, carboxylates are often found to bridge
Downloaded by Duke University on 28 July 2012

two metals across a face diagonal of the cube. We denote these


as d (diagonal) as in UMUFOX,15 Co4 /1s4 /d4 where the 2-
methylquinolin-8-olate ligand is monosupported, the quinoline
nitrogen occupying one of the terminal sites while the two
remaining terminal sites are occupied by diagonally bridging
pivalates. A further complication arises when the bridging or
diagonal ligand binds to a metal by more than one functionality.
This may be indicated by adding the number of binding groups as
in LILTUV23 (Scheme 5) where three functions bind to each metal.
In the absence of a qualifier, the ligand is assumed to bind by one
atom to each metal only (i.e., d is equivalent to d11, 3s to 3s111). Fig. 1 Partial structure of the [Co4 ((R,R)-1 − H)4 ]4+ cation showing the
two independent ligands and the Co4 O4 core. The twofold axis generates
the primed atoms.

Both ligands adopt a conformation in which the benzimidazoles


are mutually trans. Consequently the diol functions are mutually
gauche: the diol dihedral angle (measured as O1–C9–C10–O2) is
47.1(7)◦ for ligand a and 49.0(7)◦ for ligand b, smaller than the
60◦ expected for a gauche conformation. The coordination modes
shown by ligand 1 in this complex are quite different from the seven
previously recorded for this ligand. In particular they are different
from that observed for the tetranuclear copper cubane with the
non-methylated ligand, where the unprotonated alcohol function
does not coordinate.3 We may also note that the conformation is
in agreement with the steric effects observed for the methyl group
bound to the benzimidazole nitrogen.33 The terminal cobalt–
oxygen and cobalt–nitrogen bond distances are normal. The
average Co–Co distance (3.229 Å) is slightly greater than the mean
in Table 1, and the degree of distortion (as indicated by the variance
of the Co–Co distances) is high. The general features of the Co4 O4
Scheme 5 core, however, do not show significant differences from those in
the literature. There are two intramolecular stacking interactions
Crystal structure of [Co4 ((R,R)-1 − H)4 ](ClO4 )4 ·9EtOH between benzimidazoles related by a twofold axis, with interplane
angles of 0.5(2) and 13.6(2)◦ for ligands a and b respectively.
Rose coloured crystals suitable for X-ray diffraction were obtained The cations pack in layers in the ab plane (Fig. 2) and the
by slow diffusion of ethanol into a solution of ligand (R,R)-1, layers are separated by perchlorate ions and ethanol molecules.
cobalt(II) perchlorate, and base. The crystal structure shows the Hydrogen bonds are observed between the alcohol functions of
presence of the tetranuclear complex [Co4 ((R,R)-1 − H)4 ]4+ de- the ligand and ethanol solvate molecules which are themselves
duced from ESMS (Fig. 1) which lies on the twofold axis (parallel hydrogen bonded to perchlorate ions. Five ethanol molecules (one
to a) of the orthorhombic space group C2221 . The complex has the of which is disordered about a twofold axis) form a pentagonal

This journal is © The Royal Society of Chemistry 2007 Dalton Trans., 2007, 332–341 | 335
View Online
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A
Downloaded by Duke University on 28 July 2012

Fig. 2 View of one of the cation layers in the crystallographic ab plane.

motif linked by hydrogen bonds in a R55 arrangement.34,35 The


solvate molecules are readily lost and the crystals are thus rather
unstable.
Fig. 3 Visible (a) and CD (b) spectra of [Co4 ((R,R)-1 − H)4 ]4+ in
It proved impossible to obtain crystals of the nickel complex
acetonitrile solution.
suitable for accurate X-ray diffraction studies. The best crystals
obtained allowed us only to obtain an estimate of the unit cell geometry for the metal, and are slightly blue shifted with respect
as orthorhombic P (a = 13.534(4), b = 24.05(2), c = 30.03(1) Å) to the aqua-ions, and red shifted with respect to the complexes
with a volume of 9772.4 Å3 , slightly less than that for the cobalt [M(1)2 ]2+ ,4 as would be expected for a MN2 O4 chromophore. The
compound discussed above. This is at least consistent with the CD spectra are normal for cobalt(II) and nickel(II) complexes.
existence of a tetranuclear complex, and on the basis of the ESMS
results and the similarity of the IR spectra of the two solids (Fig. Magnetic susceptibility
S1, ESI‡), we assume that the nickel complex is also a cubane.
Solvent loss was also observed with these crystals. The magnetic susceptibility of cubane complexes has been studied
extensively in recent years. Their high nuclearity and the magnetic
Electronic and circular dichroism spectra exchange pathways offered by the triply bridging oxygen atoms
make them particularly interesting, and several complexes have
The spectra obtained in solution and those obtained from diffuse shown potential as single molecule magnets.9–13 The susceptibilities
reflectance show maxima at the same wavelengths and support of these complexes fall into three classes which are sketched below
the hypothesis that there is no structural change upon dissolution. (Fig. 4). Type A is generally the most common and shows a rise
The d–d bands (Table 3, Fig. 3) are consistent with an octahedral in vT at low temperatures, until a sharp drop is observed below

Table 3 d–d Transitions (kmax /nm, e/dm3 mol−1 cm−1 ) and circular dichroism (kmax /nm, De/dm3 mol−1 cm−1 ) of [Ni4 ((S,S)-1 − H)4 ]4+ and [Co4 ((R,R)-1 −
H)4 ]4+ recorded in acetonitrile solution, and reflectivity data obtained in a MgO matrix

d–d Transitions
Absorption (kmax (e)) Reflectivity (kmin ) Assignment CD (kmax (De))

[Ni4 ((S,S)-1 − H)4 ]4+


3
1135 (10) 1155 T2g –3 A2g (m1 )
1
763 (sh) (7) 763 (sh) Eg –3 A2g 746 (−0.166)
3
662 (13) 660 T1g –3 A2g (m2 ) 663 (−0.204)
3
402 (31) 399 T1g (P)–3 A2g (m3 ) 414 (−0.236)

[Co4 ((R,R)-1 − H)4 ]4+


4
1044 (8) 1054 T2g –4 T1g (m1 )
4
704 (9) 682 A2g –4 T1g (m2 ) 702 (−0.153)
560 (76) 542 553 (−0.138)
4
527 (sh) (62) 480 (sh) T1g (P)–4 T1g (m3 ) 515 (0.372)
460 (50) 480 (sh) (0.118)
450 (−0.123)

336 | Dalton Trans., 2007, 332–341 This journal is © The Royal Society of Chemistry 2007
View Online
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

Fig. 6 Variable-temperature (300–2 K) magnetic susceptibility data of


compound [Co4 ((R,R)-1 − H)4 (CIO4 )4 (C2 H5 OH)(H2 O)5 . The plot shows
vT against T.

If we assume [Ni4 ((S,S)-1 − H)4 ]4+ to possess an ideal T d cubane


Downloaded by Duke University on 28 July 2012

structure one may use an isotropic Heisenberg Hamiltonian (H =


−2J(S1 S2 + S1 S3 + S1 S4 + S2 S3 + S2 S4 + S3 S4 ). A least-squares
fit of this model to the data gave g = 2.35 and J = −3.57 cm−1
Fig. 4 Schematic representation of the three types of magnetic behaviour in agreement with an antiferromagnetic interaction between the
observed in these systems. nickel(II) ions. However, as we discuss below, this is not a good
model for this system.
10 K. Type B shows a steady drop in vT starting in the range 50– The susceptibility of [Co4 ((R,R)-1 − H)4 ](ClO4 )4 (C2 H5 OH)-
100 K and tending towards zero at very low temperatures. Both (H2 O)5 shows typical type C behaviour. The value of vT at
types A and B have been observed for cobalt(II) and nickel(II) but 300 K is 11.2 emu K mol−1 greater than the spin-only value
type C has only been seen with cobalt(II) compounds. vT falls of 7.5 (S = 3/2, g = 2), but typical for cobalt(II) cubanes, and
slightly with temperature showing rather irregular behaviour, and arises certainly from unquenched orbital angular momentum in
is still quite high even at 2 K. the pseudo-octahedral symmetry. The value of vT decreases with
We have measured the magnetic susceptibility of [Ni4 ((S,S)- temperature, but much less than a type B system, and shows a
1 − H)4 ](ClO4 )4 (H2 O)5 and [Co4 ((R,R)-1 − H)4 ](ClO4 )4 (C2 H5 OH)- small peak at 11 K. This peak was observed in the two other type
(H2 O)5 in the temperature range of 2 to 300 K. Fig. 5 and 6 show C systems in the literature,13,36 and is thus unlikely to be an artefact.
the vT values plotted against temperature. After this peak, vT continues to fall, but is still 9.7 emu K mol−1
at 1.8 K, well above the spin-only value.
It would be desirable to establish a correlation between the
crystal structures and the magnetic properties of these compounds,
and in particular the observation of ferro- or antiferromagnetism.
The magnetic exchange pathway is through the two alkoxo-ligands
which bridge each pair of metals in the cubane. It is generally
accepted that the value of the M–O–M bond angle is critical in
characterising the exchange and for nickel there is a consensus that
below 98–99◦ , the interaction will have ferromagnetic character,
and above this angle it will be antiferromagnetic.38,48 If we return
to our structural discussion above, it will be noted that the
average M–O–M angle in these structures is around 97◦ . This is
just on the ferromagnetic side of the boundary value, and indeed
ferromagnetic (type A) behaviour is observed more often that
Fig. 5 Variable-temperature (300–2 K) magnetic susceptibility data others. It is equally clear, however, that a distortion will change
of compound [Ni4 ((S,S)-1 − H)4 ](ClO4 )4 (H2 O)5 . The plot shows vT these angles by several degrees, and this would be expected to affect
against T. the exchange pathways. If we consider the S4 distortion which is
the most common, then, if we assume that the bridging oxygen
The vT value of [Ni4 ((S,S)-1 − H)4 ](ClO4 )4 (H2 O)5 at 300 K remains in the centre of the triangular face, an S4 + distortion will
is 5.32 emu K mol−1 and decreases to 0.25 emu K mol−1 at 2 K, increase eight of the M–O–M angles, and decrease four; an S4 -
typical for a type B system. The room-temperature value is slightly distortion will decrease eight and increase four. On the basis of
greater than the spin-only value of 4.0 emu K mol−1 for an S = this reasoning we might expect that the S4 -distortion will favour
1 system with g = 2. The shape of the curve indicates an overall type A behaviour, whereas S4 + will favour type B. This is indeed
antiferromagnetic exchange interaction between the nickel(II) ions. generally the case for the nickel systems for which structure and

This journal is © The Royal Society of Chemistry 2007 Dalton Trans., 2007, 332–341 | 337
View Online

Table 4 Comparison of magnetic and structural properties

M–O–M angles/◦

Compound Ligation mode Distortion Magnitude/Å Mean Max. Min. Magnetism Ref.

Nickel complexes
ELELIQ 0s4 /t4 Irreg. 0.067 97.2 101.3 96.1 A 37
EZIGAV 2s4 /m4 S4 + 0.005 97.4 97.5 97.2 A 11
ZAXJUD 0s4 /b4 m4 S4 − 0.036 97.6 100.2 96.0 A 38
AKISOCa 1s4 /m8 S4 − 0.033 98.3 100.5 97.4 A 9
AKISUI 1s4 /m8 S4 − 0.032 98.2 99.6 97.5 A 9
BEPBEDa 1s4 /m8 S4 − 0.025 98.6 100.1 97.5 A 9,39
TOQDEIa 0s4 /b4 m4 Irreg. 0.030 97.5 99.6 96.0 A 40
EHACOFa 1s24 /m4 S4 − 0.080 97.0 100.6 95.1 A 41,42
BEPBONa 1s4 /m8 S4 − 0.058 97.4 99.7 96.3 A 39
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

MAWZUF 2s4 /m2 d D2 0.122 98.3 102.4 92.0 A 43


ZAHROP 0s4 /b4 m4 S4 − 0.023 97.4 99.0 96.5 A 44,45
MSALEN 0s4 /b4 m4 S4 − 0.015 97.7 97.3 98.7 A 46
DARPESa 1s24 /m4 S4 − 0.086 97.0 101.2 94.7 A 10
IBOGAI 0s4 /b2 d2 m4 S4 + 0.189 96.2 100.0 87.7 A 47
HUBZAFa 0s2 1s2s22/d2 m3 S4 + 0.130 97.4 101.2 89.5 B 48
TACNIH10 0s4 /t4 C3+ 0.012 99.0 98.7 99.6 B 49,50
YAQYAR 2s4 /m4 D2 0.072 99.0 102.7 95.6 B 51
Downloaded by Duke University on 28 July 2012

PAGXEA 0s4 /d4 m4 S4 − 0.075 98.3 103.3 95.9 B 52

Cobalt complexes

MAWZOZ 2s4 /m4 Irreg. 0.029 99.3 100.9 98.1 A 43


MAKFAG 0s4 /b4 m4 S4 − 0.046 97.1 99.3 95.8 A 21
JAXTOR 2s4 /m4 D2 0.012 99.0 99.8 98.0 A 22
UMUFOX 1s4 /d4 S4 − 0.133 96.6 104.1 92.7 B 53
ILUHUS 3s4 Irreg. 0.036 98.1 100.0 96.3 C 13
This work 2s122 3s2 Irreg. 0.121 98.4 108.2 94.1 C

Magnitude: degree of distortion as indicated by the e.s.d. of the metal–metal distances; M–O–M angles defined as the mean of the two angles linking a
pair of metals. a Indicates a system for which the average of two crystallographically independent molecules or two independent structure determinations
is given.

susceptibility data are available. We list in Table 4 the structural and temperatures below 20 K, a zero field splitting. Early work used a
magnetic properties of the compounds found in the Cambridge single J value,9,40,45,55 but more recent work has used two or three
Structural Database for which magnetic data have been published. J values,10,37,38,43,47 thereby taking into account the distortion from
For nickel compounds, many S4 -systems have been studied tetrahedral symmetry and leading to better fits. The fitted J values
and show, with one exception, type A behaviour. The exception, show a correlation with the M–O–M angles although the scatter
PAGXEA,52 is unusual in that it has four diagonally bridging 2- is considerable. The cobalt complexes are more complicated by
thionato-1,3-thiazoline ligands which may offer a supplementary virtue of the presence of unquenched orbital momentum, and con-
antiferromagnetic exchange pathway; this is also observed with sequently first order spin–orbit coupling, and most authors have
pivaloate ligands in the cobalt complex UMUFOX.53 EZIGAV,11 limited themselves to a qualitative description of the behaviour.
while formally a S4 + system, is only very slightly distorted and However, in a study of a defect double cubane containing cobalt(II)
all angles are less than 98◦ . The J values are therefore likely to a simulation using a spin Hamiltonian adapted to the symmetry
be very similar and ferromagnetic in nature. IBOGAI,47 while and including an anisotropic term gave a qualitatively satisfactory
the distribution of nickel–nickel distances corresponds to a S4 + fit to data which resembled type C behaviour.56 At present it
system, also has diagonally bridging pivaloate ligands resulting in seems unlikely that a useful magnetostructural correlation may be
two very short Ni–Ni distances. This may account for the type A, established: the presence of many exchange constants in any but
ferromagnetic behaviour. The S4 + system reported by Gladfelter the highest symmetry systems, coupled with the sensitivity of the
et al. shows the expected type B behaviour.54 It is harder to sign and magnitude of the J values to slight distortion makes their
predict the magnetic behaviour in less symmetric systems, since the prediction difficult, and the magnetic behaviour that is actually
dependence of the J values on the M–O–M angle is not obvious. observed depends in a complicated way on the relative values
In the case of [Co4 ((R,R)-1 − H)4 ](ClO4 )4 (C2 H5 OH)(H2 O)5 one of J.
pair of cobalt atoms is linked by two bridges with angles of 96.8
and 108.2◦ and it seems reasonable to assume that the exchange 1
H NMR spectroscopy
will be dominated by the antiferromagnetic coupling due to the
second: is it reasonable to take a mean value of 102.5◦ ? It is well established that polynuclear systems with exchange
For nickel complexes the data have generally been fitted to coupling are often susceptible to study by NMR since the
a model incorporating a Heisenberg spin Hamiltonian, and for exchange coupling creates new relaxation pathways, and if the

338 | Dalton Trans., 2007, 332–341 This journal is © The Royal Society of Chemistry 2007
View Online

relaxation is rapid enough, the paramagnetic broadening of the This could arise from a change in structure but could equally arise
1
H NMR spectrum is reduced to an acceptable level.57 We were from a dynamic effect since the cobalt(II) centre is much more
previously able to characterise polynuclear complexes of 1 with labile than the nickel.
copper(II) by NMR,3 and we have therefore recorded the NMR
spectra of [Ni4 ((S,S)-1 − H)4 ](ClO4 )4 (H2 O)5 and [Co4 ((R,R)-1 − Conclusions
H)4 ](ClO4 )4 (C2 H5 OH)(H2 O)5 as shown in Fig. 7.
We have shown that ligand 1 which we had previously shown
to form tetranuclear clusters with copper(II) can also form such
complexes with cobalt(II) and nickel, using the second, unpro-
tonated, diol function to complete the octahedral coordination
of the metal ion. This implies quite considerable conformational
flexibility on the part of the ligand. The structure is significantly
distorted from the tetrahedral ideal, but not more so than
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

other complexes in the literature. Analysis of the data in the


literature suggests that the M4 O4 has rather constant structure
for a given M and oxidation state, and that in the majority of
cases the distortion maintains some elements of the tetrahedral
symmetry, most commonly the S4 axis. Although the sign of the
magnetic exchange parameters J appears to depend on the M–
Downloaded by Duke University on 28 July 2012

O–M angle, the magnetic properties observed for the complex as


a whole depend on the interaction of several exchange pathways,
both ferromagnetic and antiferromagnetic, and are not readily
predicted.

Experimental
Fig. 7 1 H NMR spectra of [Ni4 ((S,S)-1 − H)4 ](ClO4 )4 (H2 O)5 (top) and
[Co4 ((R,R)-1 − H)4 ](CIO4 )4 (C2 H5 OH)(H2 O)5 (bottom) in acetonitrile at Safety note
293 K (* solvent peak, ◦ H2 O peak, ? tentative signal assignment).
Although no problems were experienced in handling perchlorate
The spectrum of the nickel complex is in agreement with the compounds, these salts are potentially explosive when combined
crystal structure of the cobalt complex. The aromatic protons with organic ligands. The compounds should be handled with care,
were identified by their coupling from a COSY spectrum where and only used in small quantities.
four sets of three coupled protons were identified. The failure
to observe the fourth proton of the benzimidazole presumably General
arises from its proximity to the paramagnetic centre, paramagnetic Starting materials were purchased commercially, and were used
broadening precluding the observation of a COSY signal, as found as supplied unless specified otherwise. Ligand 1 was prepared in
in the copper(II) systems. The missing aromatic peaks may be both enantiomeric forms as described previously.4 Proton NMR
the broadened signals in the region 17–21 ppm. In the spectrum spectra were recorded at 300 MHz on a Varian Gemini instrument.
shown, only three methyl resonances are visible, but the fourth is Infrared spectra were recorded as KBr discs on a Perkin-Elmer
hidden by the solvent and residual water peak. Recording the Spectrum One spectrometer. UV/vis spectra were recorded using
spectrum in DMSO and varying the temperature allowed the a Perkin-Elmer Lambda 900 UV/vis/NIR machine; reflectivity
fourth methyl signal to be shifted out of this peak. The observation spectra used the same instrument equipped with a 60 mm
of four methyl and four benzimidazole signals is consistent with integration sphere. The white standard was PTFE. Samples were
a structure [Ni4 ((S,S)-1 − H)4 ]4+ having C 2 symmetry as observed prepared in a 1 mm cell after mixing the compound with roughly
in the crystal structure of the cobalt complex. nine times the amount of MgO. CD spectra were recorded on a
The spectrum of the cobalt complex is harder to understand, Jasco J-15 spectrometer.
since there is clearly more than one species present. The majority Magnetic susceptibility data were collected with a Quantum
species shows six signals for the aromatic protons of the benz- Design SQUID magnetometer (XL5S) operating in the range
imidazoles (identified as above by the COSY spectrum) and two 300–2 K with a field of 1000 G. The data were corrected for
methyl resonances, corresponding to two different environments sample holder contribution (saran foil) and for the diamagnetic
for benzimidazoles. There are three weak resonances above 50 ppm contribution estimated using Pascal’s constants.58
and three more below −20 ppm. The ESMS suggests strongly
that [Co4 ((R,R)-1 − H)4 ]4+ is the majority species in solutions Synthesis of complexes
under these conditions, and the UV-visible spectrum is quite
similar to that observed for the crystalline solid. We therefore Tetrakis((R,R)-1,2-1H -benzimidazol-2-yl)ethanediolate)tetra-
consider [Co4 ((R,R)-1 − H)4 ]4+ to be the major species present. The cobalt(II) tetraperchlorate–ethanol solvate, [Co4 ((R,R)-1 − H)4 ]-
observation of two different environments requires the symmetry (ClO4 )4 (C2 H5 OH)(H2 O)4 .
of the species in solution to be higher than that observed in the Method (a).A mixture of Co(ClO4 )2 ·6H2 O (0.0732 g, 0.2 mmol)
solid, and that observed for [Ni4 ((S,S)-1 − H)4 ](ClO4 )4 (H2 O)5 . and (R,R)-1 (0.0644 g, 0.2 mmol) was dissolved in 10 ml hot

This journal is © The Royal Society of Chemistry 2007 Dalton Trans., 2007, 332–341 | 339
View Online

ethanol. After adding one equivalent sodium hydroxide (200 ll transformed within a week to green crystals in form of thin sheets.
NaOH 1 M, 0.2 mmol) a rose precipitate is formed. The precipitate The solution was filtered off and the crystals dried in vacuum
could be redissolved by adding 1 ml water. Diffusion of ethanol (0.0392 g, 77%). Solubility: soluble in DMSO, acetonitrile, acetone
via the gas phase led to formation of rose crystals within one day. and nitromethane, less soluble in methanol, insoluble in water,
The solution was filtered after two weeks and the crystals air dried ethanol and chloroform. Elemental analysis: Found: C, 43.2; H,
(0.050 g, 43%). 4.1; N, 11.0%; C72 H68 N16 O8 [Ni4 ((S,S)-1 − H)4 ](ClO4 )4 (H2 O)5
Method (a1). The synthesis was carried out as described in (a), requires C, 43.1; H, 3.9; N, 11.2%. UV-visible (T = 22 ◦ C, c =
but 1.5 ml water were added to dissolve the rose precipitate and 0.00983 M (CH3 CN)) kmax /nm (e/dm3 mol−1 cm−1 ) : 402 (31), 662
crystals started growing within one week. Monocrystals suitable (13), 763 (sh) (7), 1135 (10); CD (T = 22 ◦ C, c = 0.00983 M
for X-ray structure analysis were obtained. (CH3 CN)) kmax /nm (De/dm3 mol−1 cm−1 ) : 414 (−0.236), 663
Method (b). A mixture of Co(ClO4 )2 ·6H2 O (0.0366 g, (−0.204), 746 (−0.166); reflectivity in MgO kmin /nm: 399, 660, 763
0.1 mmol) and (R,R)-1 (0.0322 g, 0.1 mmol) was dissolved in 8 ml (sh), 1155 (broad band low in intensity); IR (KBr) m/cm−1 : 3416br
m; 3059w, 1617m, 1500m, 1481m, 1458m, 1412w, 1336m, 1318m,
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

hot ethanol. After adding two equivalents of 2,5-dimethylpyridine


(232 ll 96%, 0.2 mmol) a rose precipitate formed. The precipitate 1295m, 1276m, 1086br s; 1010w, 925m, 903m, 877m, 804w, 746m,
transformed within three weeks into crystals. Heat accelerates 679w, 626m, 582w, 554w, 539w, 501w; ESMS (CH3 CN): m/z (%) =
this process. The solution was filtered and the crystals air dried 758.3 (100) [Ni4 (1)4 (−6H)]2+ , 506.0 (8) [Ni4 (1)4 (−5H)]3+ , 1617.0 (8)
(0.0430 g, 74%). It was observed that the crystals readily lost [Ni4 (1)4 (−6H)(ClO4 )]+ , 519.6 (7) [Ni4 (1)4 (−5H)(CH3 CN)]3+ , 809.2
solvent, and the composition of the air-dried sample is thus (6) [Ni4 (1)4 (−5H)(ClO4 )]2+ ; 1 H NMR (assignment not complete,
different from the crystals used for X-ray diffraction studies, ? not assigned peaks) (T = 22 ◦ C, CD3 CN, 300 MHz: d (ppm))
Downloaded by Duke University on 28 July 2012

where care was taken to avoid solvent loss. Solubility: soluble 29.82 (1Harom , s), 29.77 (1Harom , s), 25.17 (1Harom , s), 23.30 (1Harom ,
in DMSO, acetonitrile, acetone and a water–ethanol mixture s), 20.68 (1H br, ?), 17.14 (2H br, ?), 13.66 (1H br, ?), 10.16 (1Harom ,
(1 : 9); less soluble in methanol; insoluble in ethanol, water, s), 9.60 (1Harom , s), 8.65 (1Harom , s), 8.55 (1Harom , s), 7.47 (1Harom , s),
ethanol–water mixture (1 : 1), chloroform, and tert-butyl methyl 7.03 (1Harom , s), 6.11 (1Harom , s), 5.62 (1Harom , s), 4.22 (3H, CH3 , s),
ether. Elemental analysis: Found : C, 43.2; H, 4.2; N, 10.8%; 2.66 (3H, CH3 , s), −0.83 (3H, CH3 , s). (H,H) COSY (300 MHz,
C72 H68 N16 O8 ·4Co·4ClO4 ·5H2 O·C2 H5 OH requires C, 43.3; H, 4.1; Varian Gemini 300), cross peaks (ppm): 29.82/8.55, 8.55/10.16,
N, 10.9%; UV-visible (T = 22 ◦ C, c = 0.015 M (CH3 CN)) kmax /nm 29.77/8.65, 8.65/9.60, 25.17/6.11, 6.11/7.03, 23.30/–5.62, 5.62/–
(e/dm3 mol−1 cm−1 ): 460 (50), 527 (sh) (62), 560 (76), 704 (9), 7.47.
1044 (8); CD (T = 22 ◦ C, c = 0.015 M (CH3 CN)) kmax /nm
Crystal structure determination of [Co4 ((R,R)-1 − H)4 ](ClO4 )4 -
(De/dm3 mol−1 cm−1 ) : 450 (−0.528), 480 (sh) (0.4701), 515 (1.487),
(C2 H6 O)9 . Crystals were obtained by vapour diffusion of ethanol
553 (−0553), 702 (−0.613); reflectivity spectrum: in MgO kmin /nm:
into an aqueous ethanol solution of the complex. Crystal data:
480 (sh), 542, 682, 1054; IR (KBr) m/cm−1 : 3414br m, 3060w,
Co4 (C18 H17 N4 O2 )4 (ClO4 )4 (C2 H6 O)9 , M = 2333.8, orthorhombic,
2947w, 2656w, 2022w, 1616m, 1497m, 1483m, 1457s, 1408m,
space group C2221 , a = 19.1066(12), b = 23.3433(10), c =
1318m, 1294m, 1278m, 1247w, 1087br s, 1010w, 970w, 924m,
23.6269(17) Å, U = 10538(1) Å3 , T = 200 K, Z = 4, l(Mo-Ka) =
898m, 876m, 876m, 804w, 744s, 680w, 625s, 591w, 565w, 548w,
0.81 mm−1 , 66 435 reflections measured (Stoe IPDS), 10 276 unique
537w, 485m, 404w, 353w. ESMS (CH3 CN): m/z (%) = 759.2(100)
(Rint. = 0.069), 5838 observed (|F o | > 4r(F o )). Final R value
[Co4 (1)4 (−6H)]2+ , 520.2 (38) [Co4 (1)4 (−5H)(CH3 CN)]3+ , 506.5
0.037 (wR = 0.037) for 663 variables, absolute structure (Flack)59
(16) [Co4 (1)4 (−5H)]3+ , 816.1 (7) [Co3 (1)2 (−5H)]+ , 1516.8 (3)
parameter x = −0.01(2). The structure was solved using direct
[Co4 (1)4 (−7H)]+ . 1 H NMR (? unassigned peaks or tentative
methods (MULTAN-8760 ), and other calculations used XTAL
assignment) (T = 22 ◦ C, c = 0.0428 M (CD3 CN), 300 MHz):
3.261 and ORTEP-II62 programmes. The hydrogens of OH groups
primary species (probably [Co4 (1 − H)4 ]4+ ), only the eight sharpest
were not observed.
signals are listed with their tentative signal assignment: d (ppm):
CCDC reference number 622954.
39.10 (aHarom ?, s), 23.38 (1Harom ?, s), 18.27 (1Harom ?, s), 10.56
For crystallographic data in CIF or other electronic format see
(1Harom ?, s), 5.90 (3H, CH3 , s), −9.81 (1Harom ?, s), −13.30 (1Harom ?,
DOI: 10.1039/b614424a
s), −16.85 (3H, CH3 , s). Unassigned signals due to the primary
species and one or two secondary species: d(ppm): 99.82 (s, br),
83.72 (s, br), 74.75 (s, br), 64.34 (s, br), 63.33 (s, br), 44.12 (s), Acknowledgements
39.09 (s), 27.28 (s), 24.52 (s), 17.63 (s), 11.81 (s), 4.08 (s), 3.51
(s), 2.84 (s), 0.67 (s), −0.67 (s), −3.22 (s), −8.44 (s), −33.66 (s), We gratefully acknowledge the support of this work by the Swiss
−54.05 (s, br), −79.70 (s, br). (HH) COSY (with gradient selection, National Science Foundation.
500 MHz, Bruker), cross peaks (ppm): 39.10/18.27, 10.56/9.81,
−9.81/13.30.
References
Tetrakis((S,S) - 1,2 - 1H - benzimidazol - 2 - yl)ethanediolate)tetra-
1 R. E. P. Winpenny, Adv. Inorg. Chem., 2001, 52, 1.
nickel(II) tetraperchlorate hydrate, [Ni4 ((S,S)-1 − H)4 ](ClO4 )4 - 2 D. Gatteschi and R. Sessoli, Angew. Chem., Int. Ed., 2003, 42, 268.
(H2 O)5 . A mixture of Ni(ClO4 )2 ·6H2 O (0.0366 g, 0.1 mmol) and 3 K. Isele, P. Franz, C. Ambrus, G. Bernardinelli, S. Decurtins and A. F.
(S,S)-1 (0.0322 g, 0.1 mmol) was dissolved in 10 ml ethanol. Williams, Inorg. Chem., 2005, 44, 3896.
4 K. Isele, V. Broughton, C. J. Matthews, A. F. Williams, G. Bernardinelli,
On adding one equivalent 2,5-dimethylpyridine (117 ll, 96%)
P. Franz and S. Decurtins, J. Chem. Soc., Dalton Trans., 2002, 3899.
the colour of the solution turned from turquoise to green. 5 W. Kaim and B. Schwederski, Bioinorganic Chemistry: Inorganic
A green microcrystalline precipitate formed immediately, which Elements in the Chemistry of Life, J. Wiley, & Sons, New York, 1994.

340 | Dalton Trans., 2007, 332–341 This journal is © The Royal Society of Chemistry 2007
View Online

6 P. Giastas, N. Pinotsis, G. Efthymiou, M. Wilmanns, P. Kyritsis, J.-M. 31 R. E. P. Winpenny, G. Timco, S. Parsons and D. Messenger, personal
Moulis and I. M. Mavridis, J. Biol. Inorg. Chem., 2006, 11, 445. communication, 2005.
7 K. N. Ferreira, T. M. Iverson, K. Maghlaoui, J. Barber and S. Iwata, 32 M. M. Olmstead, P. P. Power and G. A. Sigel, Inorg. Chem., 1988, 27,
Science, 2004, 303, 1831. 580.
8 B. Schmid, H.-J. Chiu, V. Ramakrishnan, J. B. Howard and D. C. Rees, 33 A. Kübel-Pollak, C. J. Matthews, S. Verdan, B. Bocquet, X. Melich,
in Handbook of Metalloproteins, ed. A. Messerschmidt, R. Huber, A. F. Williams, F. Lavergnat, P.-Y. Morgantini and G. Bernardinelli,
T. Poulos and K. Wieghardt, Chichester, New York, Weinheim, New J. Chem., 2006, 30, 851.
Brisbane, Singapore, Toronto, 2001, p. 1025. 34 M. C. Etter, Acc. Chem. Res., 1990, 23, 120.
9 E.-C. Yang, W. Wernsdorfer, S. Hill, R. S. Edwards, M. Nakano, S. 35 J. Bernstein, R. E. Davis, L. Shimoni and N.-L. Chang, Angew. Chem.,
Maccagnano, L. N. Zakharov, A. L. Rheingold, G. Christou and D. N. Int. Ed. Engl., 1995, 34, 1555.
Hendrickson, Polyhedron, 2003, 22, 1727. 36 S. G. Telfer, R. Kuroda, J. Lefebvre and D. B. Leznoff, Inorg. Chem.,
10 A. Sieber, C. Boskovic, R. Bircher, O. Waldmann, S. Ochsenbein, 2006, 45, 4592.
G. Chaboussant, H. U. Güdel, N. Kirchner, J. van Slageren, W. 37 T. K. Paine, E. Rentschler, T. Weyhermüller and P. Chaudhuri,
Wernsdorfer, A. Neels, H. Stoeckli-Evans, S. Janssen, F. Juranyi and Eur. J. Inorg. Chem., 2003, 3167.
H. Mutka, Inorg. Chem., 2005, 44, 4315. 38 M. A. Halcrow, J.-S. Sun, J. C. Huffman and G. Christou, Inorg. Chem.,
11 M. Moragues-Cánovas, M. Helliwell, L. Ricard, E. Rivière, W. 1995, 34, 4167.
Published on 21 November 2006 on http://pubs.rsc.org | doi:10.1039/B614424A

Wernsdorfer, E. Brechin and T. Mallah, Eur. J. Inorg. Chem., 2004, 39 A. Escuer, M. Font-Bardı́a, S. B. Kumar, X. Solans and R. Vicente,
2219. Polyhedron, 1999, 18, 909.
12 E.-C. Yang, D. N. Hendrickson, W. Wernsdorfer, M. Nakano, L. N. 40 M. S. El Fallah, E. Rentschler, A. Caneschi and D. Gatteschi, Inorg.
Zakharov, R. D. Sommer, A. L. Rheingold, M. Ledezma-Gairaud and Chim. Acta, 1996, 247, 231.
G. Christou, J. Appl. Phys., 2002, 91, 7382. 41 C. Boskovic, E. Rusanov, H. Stoeckli-Evans and H. U. Güdel, Inorg.
13 M. Murrie, S. J. Teat, H. Stoeckli-Evans and H. U. Güdel, Angew. Chem. Commun., 2002, 5, 881.
Chem., Int. Ed., 2003, 42, 4653. 42 N. Hoshino, T. Ito, M. Nihei and H. Oshio, Chem. Lett., 2002, 844.
14 G. Aromi, A. Batsanov, P. Christian, M. Helliwell, A. Parkin, A. A. 43 M.-L. Tong, S.-L. Zheng, J.-X. Shi, Y.-X. Tong, H. L. Lee and X.-M.
Downloaded by Duke University on 28 July 2012

Smith, G. A. Timco and R. E. P. Winpenny, Chem. Eur. J., 2003, 9, Chen, J. Chem. Soc., Dalton Trans., 2002, 1727.
5142. 44 M. Darensbourg, R. M. Buonomo and J. H. Reibenspies, Z. Kristal-
15 G. Aromı́, A. S. Batsanov, P. Christian, M. Helliwell, O. Roubeau, G. A. logr., 1965, 210, 469.
Timco and R. E. P. Winpenny, Dalton Trans., 2003, 4466. 45 J. A. Bertrand, A. P. Ginsberg, R. I. Kaplan, C. E. Kirkwood, R. L.
16 R.-K. Chiang, C.-C. Huang and C.-S. Wur, Inorg. Chem., 2001, 40, Martin and R. C. Sherwood, Inorg. Chem., 1971, 10, 240.
3237. 46 J. E. Andrew and A. B. Blake, J. Chem. Soc. A, 1969, 1456.
17 H. Zhao, J. Bacsa and K. R. Dunbar, Acta Crystallogr., Sect. E, 2004, 47 G. Chaboussant, R. Basler, H. U. Güdel, S. T. Ochsenbein, A. Parkin,
60, m637. S. Parsons, G. Rajaraman, A. Sieber, A. A. Smith, G. A. Timco and
18 B. A. Averill, T. Herskovitz, R. H. Holm and J. A. Ibers, J. Am. Chem. E. P. Winpenny, Dalton Trans., 2004, 2758.
Soc., 1973, 95, 3523. 48 J. M. Clemente-Juan, B. Chansou, B. Donnadieu and J.-P. Tuchagues,
19 P. V. Rao and R. H. Holm, Chem. Rev. (Washington, DC), 2004, 104, Inorg. Chem., 2000, 39, 5515.
527. 49 B. Aurivillius, Acta Chem. Scand., 1977, 31, 501.
20 M.-L. Tong, H. K. Lee, S.-L. Zheng and X.-M. Chen, Chem. Lett., 50 P. D. W. Boyd, R. L. Martin and G. Schwarzenbach, Aust. J. Chem.,
1999, 1087. 1988, 41, 1449.
21 J. F. Berry, F. A. Cotton, C. Y. Liu, T. Lu, C. A. Murillo, B. S. Tsukerblat, 51 G. S. Papaefstathiou, A. Escuer, F. A. Mautner, C. Raptopoulou, A.
D. Villagrán and X. Wang, J. Am. Chem. Soc., 2005, 127, 4895. Terzis, S. P. Perlepes and R. Vicente, Eur. J. Inorg. Chem., 2005, 879.
22 A. Tsohos, S. Dionyssopoulou, C. P. Raptopoulou, A. Terzis, E. G. 52 L. Ballester, E. Coronado, A. Gutiérrez, A. Monge, M. F. Perpiñan, E.
Bakalbassis and S. P. Perlepes, Angew. Chem., Int. Ed., 1999, 38, 983. Pinilla and T. Rico, Inorg. Chem., 1992, 31, 2053.
23 C. He and S. J. Lippard, J. Am. Chem. Soc., 2000, 122, 184. 53 G. Aromı́, A. S. Batsanov, P. Christian, M. Helliwell, O. Roubeau, G. A.
24 E. K. Brechin, S. G. Harris, S. Parsons and R. E. P. Winpenny, Chem. Timco and R. E. P. Wimpenny, Dalton Trans., 2003, 4466.
Commun., 1996, 1439. 54 W. L. Gladfelter, M. W. Lynch, W. P. Schaefer, D. N. Hendrickson and
25 K. E. Gubina, V. A. Ovchynnikov, J. Swiatek-Kozlowska, V. M. H. B. Gray, Inorg. Chem., 1981, 20, 2390.
Amirkhanov, T. Yu. Sliva and K. V. Domasevitch, Polyhedron, 2002, 55 J. A. Barnes and W. E. Hatfield, Inorg. Chem., 1971, 10, 2355.
21, 963. 56 P. King, R. Clérac, W. Wernsdorfer, C. E. Anson and A. K. Powell,
26 R. Wang, M. Hong, W. Su and R. Cao, Acta Crystallogr., Sect. E, 2001, Dalton Trans., 2004, 2670.
57, m325. 57 V. Clementi and C. Luchinat, Acc. Chem. Res., 1998, 31, 351.
27 Yu. A. Simonov, G. S. Matuzenko, M. M. Botoshanskii, M. A. 58 O. Kahn, Molecular Magnetism, VCH, New York, 1993.
Yampol’skaya, N. V. Gerbeleu and T. I. Malinovskii, Zh. Neorg. Khim., 59 H. D. Flack, Acta Crystallogr., Sect. A, 1983, 39, 876.
1982, 27, 407. 60 MULTAN 87, Universities of York (UK) and Louvaine-La-Neuve
28 Yu. A. Simonov, V. K. Bel’skii, G. S. Matuzenko and N. V. Gerbeleu, (Belgium), 1987.
Kristallografiya, 1984, 29, 82. 61 S. R. Hall, H. D. Flack and J. M. Stewart, XTAL 3.2 User’s Manual,
29 V. G. Kessler, S. Gohil, M. Kritikos, O. N. Korsak, E. E. Knyazeva, Universities of Western Australia, Geneva and Maryland, 1992.
I. F. Moskovskaya and B. V. Romanovsky, Polyhedron, 2001, 20, 915. 62 C. K. Johnson, ORTEP II Report ORNL-5138, Oak Ridge National
30 S. Parsons, R. Winpenny and P. Wood, personal communication, 2004. Laboratory, Oak Ridge, TN, 1976.

This journal is © The Royal Society of Chemistry 2007 Dalton Trans., 2007, 332–341 | 341

You might also like