Вы находитесь на странице: 1из 28

This article was published as part of the

2009 Metalorganic frameworks issue


Reviewing the latest developments across the interdisciplinary area of
metalorganic frameworks from an academic and industrial perspective
Guest Editors Jeffrey Long and Omar Yaghi
Please take a look at the issue 5 table of contents to access
the other reviews.
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online / Journal Homepage / Table of Contents for this issue
Magnetic metalorganic frameworksw
Mohamedally Kurmoo
Received 14th November 2008
First published as an Advance Article on the web 24th February 2009
DOI: 10.1039/b804757j
The purpose of this critical review is to give a representative and comprehensive overview
of the arising developments in the eld of magnetic metalorganic frameworks, in particular those
containing cobalt(II). We examine the diversity of magnetic exchange interactions between
nearest-neighbour moment carriers, covering from dimers to oligomers and discuss their
implications in innite chains, layers and networks, having a variety of topologies. We
progress to the dierent forms of short-range magnetic ordering, giving rise to
single-molecule-magnets and single-chain-magnets, to long-range ordering of two- and
three-dimensional networks (323 references).
1. Introduction
One of the many attributes of MetalOrganic Frameworks
(MOF) is magnetism.
1,2
It can be implemented by incorporating
magnetic moment carriers such as paramagnetic metals or
open-shell organic ligands or both.
3,4
This, in itself, is not
enough to render a material magnetic as magnetism is a
cooperative phenomenon and thus requires some kind of
exchange between the moment carriers.
5
Therefore, the connec-
tion between moment carriers at distances within interacting
range is required. This is often achieved in MOFs. Magnetism
can also be introduced as guests while the framework remains
non-magnetic. So far the majority of magnetic frameworks are
those containing paramagnetic metal centres and in particular,
the rst row transition metals (V, Cr, Mn, Fe, Co, Ni and Cu).
These metals, which may exist in dierent oxidation states,
allow variation of the two important parameters, spin quantum
number and magnetic anisotropy. The potential of such a range
of variables opens many challenges in a eld without frontiers.
The readers of this collection of works will notice that many
of the compounds reported today are classied in dierent
categories belonging to the general theme Framework, such
as MOF (MetalOrganic Framework), IRMOF (isoreticular
MetalOrganic Framework), MOP (MetalOrganic
Polyhedra), ZIF (Zeolitic Imidazole Framework), PMOF
(Porous MetalOrganic Framework), CP (Coordination Poly-
mers), PCP (Porous Coordination Polymers), MMOF
(Microporous MetalOrganic Framework) and more.
613
However, there is not one where magnetic is mentioned, for
example MMOF (Magnetic MetalOrganic Framework). This
is, most likely, due to the fact that the eld of magnetism is
much older and there is already an existing classication,
where the classes are Itinerant Magnets, Mineral Magnets,
Organic Magnets, Molecular Magnets and some less
common ones.
1416
Within this classication, Magnetic MOF
are well described by Molecular Magnets. They are both
branches of coordination chemistry where metals are bound
in a solid by coordination bonds to organic ligands. From the
structural point of view, there is a great interest in nding a
logical way to develop a catalogue and thus the development
proposed by OKeefe
17,18
which consequently brings the
notion of Yaghis Secondary-Building-Units
1921
where the
atoms in a crystal are replaced by a cluster and the organic
connections between them make the bonds. Consequently,
many simple and basic crystal structures, for example NaCl,
CsCl, CdCl
2
, rutile, diamond, quartz and more,
22,23
were
realised which happens to be the way magnetic materials were
classied as the structure denes the exchange pathways.
5,14
Interestingly, as the inorganic nodes get bigger and organic
linkers get longer, and of dierent geometry and multitopicity,
a variety of highly stable lightweight materials has been
synthesised.
2426
Some of these materials have spaces between
the nodes containing solvents that can be emptied and relled
without loss of the mosaicity of the crystals. As a result of the
porosity there is a major interest in this area for dierent
applications; among them that of storing fuels and unwanted
gases is of utmost importance.
2732
It is to be noted that the present word Framework is in a
sense related to Cooperative which has been in use for many
years and is still used today.
33,34
The word Cooperative has
several meanings; in our context it principally refers to inter-
action of the valence electrons of neighbouring atoms in a
Dr Mohamedally Kurmoo is now a permanent researcher of the
CNRS (France), based at Universite de Strasbourg, following a
PhD at University College London, and research assistant
positions at the Inorganic Chemistry Laboratory (Oxford)
and Clarendon Laboratory (Oxford). He was Assistant
Director of the Davy-Faraday Research Laboratory at the
Royal Institution of Great Britain (London) for ve years. He
has held several invited research and professorial positions in
France, Italy, United Kingdom, Japan and the USA. He has
published over 300 papers on the synthesis, characterisation and
studies of the properties of metalorganic materials, in
particular, those exhibiting multi-properties (structural,
electrical, optical, magnetic and porosity).
Laboratoire de Chimie de Coordination Organique, CNRS-UMR7140,
Universite de Strasbourg, 4 rue Blaise Pascal, 67000, Strasbourg
Cedex, France. E-mail: kurmoo@chimie.u-strasbg.fr
w Part of the metalorganic frameworks themed issue.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1353
CRITICAL REVIEW www.rsc.org/csr | Chemical Society Reviews
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
molecule or solid. The consequence of these interactions is a
wide range of cooperative phenomena such as electrical
conductivity, magnetism and optical properties. In principle,
going from atoms to SBUs does not change these cooperative
properties and therefore, most of the electrical and optical
ground states are basically the same. However, this is not
always true for the case of magnetism and in this review, we
will develop the progress made in the last ten years which
highlights the fact. As the progress of the eld is exponential, it
is very dicult to cover all the cases in the literature.
Therefore, we have made one selection; that is to present only
magnetic materials of cobalt and where necessary mention
some relevant materials having other magnetic centres or
combination.
This review, therefore, proceeds as follows: we rst discuss the
reasons behind our choice for cobalt highlighting its magnetic
properties as a monomer, then develop our understanding of the
exchange between nearest neighbours, from experimental results,
of its dierent modes of connection and their consequence in
simple dimers, through trimers, tetramers to oligomers. Finally,
we present how long-range magnetic ordering develops in linear
chains, layers and 3D-networks. We also present recent develop-
ments such as single-molecule-magnetism, single-chain-magnetism
as well as the eects of solvents on the magnetic ground states in
some porous magnets.
3537
We conclude with a general view of
where we are coming from and where we are going.
There are several reviews, in addition to the other chapters
of this volume, covering the designs, synthetic approaches,
structures and physical properties of these materials.
3855
Therefore, we will not go into already published details but
will keep this chapter to the relevant parts in view of
developing the structuremagnetic property relationships to
provide a platform for further development of interesting
magnetic materials, both of academic and industrial interests.
Magnetic interaction between nearest neighbour moment
carriers as presented in this review is derived from several
structural modes which may result in dierent signs and
values. The general view of the strength of the interaction,
based on experimental observations and in some cases on
theory, is that it decreases with the number of intervening
bonded atoms between the moment carriers. For example, the
critical ordering temperatures for the three magnetic metals of
the rst row transition metals range between 560 and 1380 K,
while for oxides (one-atom bridge) of the same metals it does
not exceed 900 K and for compounds with two-atom bridges
(e.g., cyanide) it is below 350 K. On increasing the number of
atoms in the bridge to three (e.g., dicyanamide and azide) the
transition temperatures do not get higher than 50 K.
56,57
When the connection is through four-atom bridges, no long-
range magnetic ordering is observed above 2 K. Furthermore,
the transition temperature decreases in the order 3D42D41D.
We have to stress that the above correlation applies to systems
which has all pairwise exchange pathways equivalent. For the
present review we will adopt the denition of magnetic
dimensionality by considering only bridges of less than four
atoms, although bridges of four and more atoms become
important at low temperatures through dipolar interactions
and especially when the ordered SBU (may it be a cluster, a
chain or a layer) has a large resultant correlated moment.
Common to the chemistry of the frameworks, especially for
cobalt, is the presence of water, hydroxide and carboxylate
that form the connections between the magnetic centres. These
are the main examples that will be considered here, though
some examples may contain in addition bipyrimidine,
pyrazole, triazole and tetrazole that provide CoNNCo
bridges.
2. Cobalt
Cobalt, translated from its German name Kobald chosen by
its discoverer Georg Brandt, meaning goblin or evil spirit, is
absolute gold for scientists interested in structural chemistry
and magnetism. One of the interesting structural aspects of
working with cobalt in contrast to nickel, iron and manganese
is the range of geometries, octahedral, tetrahedral, square-
pyramidal, trigonal-bipyramid and square-planar, that are
stable.
58
The order of occurrence in the literature for the
known MOFs is as listed. We note that the coordination
geometry is purely a consequence of serendipity and is not a
designable parameter. As for magnetism it provides the highest
magneto-crystalline anisotropy which results in record
magnetic hardness.
The ligands used in the development of MOF are usually
those having oxygen (oxide (rarely found for cobalt), water,
hydroxide, alkoxide, alcohol and carboxylate) or nitrogen
(amine, pyridine, azide and azole) coordinating atoms and in
few cases, sulfur (thiolate). Water or alcohol can contribute
one (terminal) or two (one-atom bridge) bonds to the metals,
while hydroxide or alkoxide can contribute from one
(terminal) up to three (one-atom bridge) bonds. In contrast,
carboxylate can contribute from one up to four bonds to the
metals, adopting combinations of syn and anti congurations
that have some consequences on the magnetic exchange
between nearest neighbours. Ancillary ligands with nitrogen
can contribute one bond per pyridine nitrogen atom and one
or two per azide nitrogen atom to the metals. The connection
can be a two-atom bridge as in the case of pyrazole, triazole,
tetrazole and bipyrimidine. For these donor atoms the
ligand-eld remains weak favouring the spin state, S = 3/2
(t
2g
5
e
g
2
(O
h
) or e
g
4
t
2g
3
(T
d
)).
As found in nature, cobalt-containing MOFs can have a
wide range of colours which makes it easy, in many cases, to
identify dierent phases.
59
This may be one of the reasons why
cobalt frameworks are also popular with magneto-chemists.
The most common colours are, a very light pink for octahedral
coordinated ones and an intense blue for tetrahedral
coordination. For distorted octahedra, the pink colour can
intensify to orange, dark red, purple and violet depending on
the ligand and the type of distortions. For compounds having
both tetrahedral and octahedral coordinated ions, they are
usually green to blue depending on the distortions and the
ratio of cobalt atoms in the two coordination geometries.
However, some unusual colours such as colourless, gray and
green have been observed for octahedral cobalt ions.
6062
Structural polymorphism is occasionally encountered in
MOFs.
6366
In most cases, dierent condensations of salt
are formed when one works at dierent temperatures
in the hydrothermal synthesis.
6770
For example, in a
1354 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
neutral or slightly acid medium the hydrogen-bonded
network [Co(H
2
O)
6
](dicarboxylate) is formed at temperatures
below 90 1C, and it transforms into the one-dimensional
Co(H
2
O)
4
(dicarboxylate) for temperatures between 100 and
120 1C and to a two-dimensional Co(H
2
O)
2
(dicarboxylate) at
a slightly higher temperature.
7173
By increasing the alkalinity
of the medium, condensed hydroxide containing networks are
often obtained. There is a general tendency to form clusters at
lower temperature, where the clusters increase in size to chains
and ladders through layers and nally 3D-frameworks at
higher temperatures. This progression of structural condensa-
tion provides a whole gamut of clusters that are very
important in the understanding of the magnetism of MOFs;
and consequently, our aim for this review is to develop our
understanding from that of a simple metal centre to a dimer,
progressively through the oligomers to innite 1D, 2D and 3D
structures.
Cobalt, as a metal, has the highest Curie temperature among
the three ferromagnets of the rst-row transition metals and is
also the most anisotropic magnetically. In its oxidised form the
divalent (d
7
) and tetravalent (d
5
) cations are paramagnetic but
the trivalent (d
6
) is generally diamagnetic. In most MOFs of
cobalt the oxidation state is two and therefore, we will only
consider this ionic state in what follows. In octahedral coordi-
nation geometry, the ground state is a
4
T
1g
such that if one uses
the spin-only formula for S = 3/2 and a g-value of 2, the
eective moment is expected to be 3.87 cm
3
K mol
1
. However,
the observed values are usually higher (4.75.2 m
B
) due to the
unquenched orbital-moment as a consequence of spinorbit
coupling (l =B170 cm
1
= 118 K).
58
The spinorbit coupling
lifts the twelve-fold degeneracy of the ground state and gives
rise to a doublet, a quartet and a sextet of Kramers levels. The
lowest level being the doublet. Upon cooling, the depopulation
of the upper levels causes the eective magnetic moment to be
reduced even in the absence of magnetic exchange and Mabbs
and Machin analysed the temperature dependence of the
magnetic susceptibility and arrived at an articial Weiss
constant of 20 K for an isolated Co(II).
74
The temperature
dependence of the magnetic susceptibility has frequently been
simulated using the equation of Mabbs and Machin or
modications thereof to include axial distortion of the six-
coordinated Co(II) ion by taking into account the energy gap
between the singlet
4
A
2
and doublet
4
E levels resulting from the
splitting of the orbital triplet
4
T
1g
ground state under an axial
distortion.
75
For the low temperature state, Abraham and Price
have found an eective S = 1/2 and a g
av
= 13/3 describe best
the EPR spectra of several Co(II) salts.
7678
Consequently, the
isothermal magnetization deviates from that of an expected
Brillouin function for an isotropic S = 3/2 ion with a saturation
value (gS) of 2.32.4 m
B
. For tetrahedral Co(II), the ground state
term is
4
A
2
and the eect of spinorbit is not so large. The
observed range of eective magnetic moment is 4.54.8 m
B
,
g
av
= 2.4 and S = 3/2.
7987
For non-interacting tetrahedral
Co(II), the Weiss constant is close to zero. For a trigonal-
bipyramid and square-pyramid, the behaviour is intermediate
between those of octahedral and tetrahedral coordinated
Co(II).
88
Square-planar Co(II) systems, which are very rare in
MOF, adopt the low-spin state (S = 1/2). The average g-value
is slightly elevated from the free-electron value due to a small
orbital contribution resulting in an eective moment range of
2.22.7 m
B
.
89
3. Magnetism of metalorganic frameworks
There exist several reviews dealing with the dierent aspects of
magnetism.
90105
In addition to the general reviews, some
cover special areas such as (a) simple intuitive synthetic
approaches to designing magnets, (b) the eect of distance,
(c) symmetry relations, (d) and particular ligands (e.g. cyanide,
oxalate, azide etc.). Furthermore, there are some more physics
oriented reviews covering special issues such as frustration,
low-dimensionality and more. These reviews provide a very
good basis of magnetism in molecular systems from which we
can draw much in understanding those of MOFs. We shall
start with a single centre and develop through to more com-
plex clusters and to innite 1D, 2D and 3D networks.
3.1 Monomers
Monomeric compounds of cobalt with carboxylate ligands
are rarely found when the synthesis is carried out hydro-
thermally.
63,106,107
An example of one which is closest to a
perfect octahedral geometry is [Co(H
2
O)
6
](H
2
pm) where pm is
1,2,4,5-benzenetetracarboxylate (Fig. 1 (I)) which has the
hexaaquo-cobalt hydrogen-bonded to the pyromellitate to
form a 3D network.
63
The susceptibility is as described above
with Curie constant (C = 2.95 cm
3
K mol
1
Co) corres-
ponding to an eective moment of 4.86 m
B
and Weiss constant,
Y = 19.9(6) K. The isothermal magnetization saturates at
2.3 m
B
. This set of values is observed for a series of
one-dimensional chain compounds of the general formula,
Co(H
2
O)
4
(dicarboxylate) where the linear dicarboxylate
(fumarate, succinate, 1,4-cyclohexanedicarboxylate,
hexanoate, and others) forms the bridge between the
Co(H
2
O)
4
planar units.
108
These chains can be helical as in
the case of 2,2
0
-biphenyldicarboxylate but with surprisingly a
larger Weiss constant of 51 K.
109
However, when the
cobalt environment is heavily distorted from an octahedron
these values can show major deviations. As an example,
Co
2
(cis-1,4-chdc)
2
(dpa)
2
where chdc is cyclohexanedicarboxylate
and dpa is 2,2
0
-dipyridylamine (Fig. 1 (II)),
110
which has
bond lengths in the range 2.0882.301 A

for CoO and


2.0502.084 A

for CoN. The major distortions are generated


by the presence of chelating carboxylate groups which resulted
in a range of angles of 59.6144.91 for OCoO and 93.6159.91
for OCoN. Analysis of the magnetic susceptibility gives
C = 2.98 cm
3
K mol
1
Co and Y = 1.4(2) K. For an
almost tetrahedral Co(II), as found in Co
2
(2,2
0
-bpdc)
2
(dpa)
2
where bpdc is biphenyldicarboxylate (Fig. 1 (III)),
110
where
the cobalt ions are coordinated by two oxygen atoms and two
nitrogen atoms, C = 2.47 cm
3
K mol
1
Co and Y =1.2(2) K.
The conclusion from these simple cases is that for tetrahedral
system we usually nd the expected values while for octahedral
systems one has to consider the distortions from a regular
octahedron before commenting on the Weiss constants. To
our knowledge, examples of isolated Co-carboxylate in
penta- or square-planar coordination are rather rare in the
eld of MOF.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1355
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
3.2 Dimers
By modelling the temperature dependence of the susceptibility of
dimers and trimers it is usually possible to extract the value of the
exchange interaction between the nearest neighbours, which can
in turn be very useful in our understanding of the magnetism of
more extended systems. However, such analyses for cobalt have
not been very extensive due to the change of regime, consequent
of the eect of spinorbit coupling, in the range of temperature
where experiments are performed. While most analyses are
supercial, some elaborate calculations have been performed
by Ostrovsky et al. and Palii et al. for specic examples.
111114
Such calculations will be very informative and are required for
the dierent modes of connection that are known. These
connections can be just a simple corner-sharing with one bridging
oxygen atom as for Co
2
(2-pzc)(H
2
O)(VO
3
)
3
(Fig. 2 (I)),
115
2-pzc = 2-pyrazinecarboxylate, and Co
5
(bpc)
2
(H
2
O)
12
12H
2
O
(Fig. 2 (III)),
116
bpc = benzenepentacarboxylate, or an edge-
sharing through oxygen bridges as for Co
2
(hypa)
2
(4,4
0
-bpy)
(Fig. 2 (IV)),
117
hypa = hydroxyphenylacetate. A corner sharing
and two carboxylate bridges are observed for an elegant porous
system, Co
2
(nicotinate) (Fig. 2 (II)),
118
where the lattice water
molecules can be removed and other solvent inserted without
damaging the crystals. For the last four examples, the estimation
of the exchange constant have been obtained by simple
CurieWeiss t of the high-temperature data or using equations
for dimers derived from an isotropic Hamiltonian. Many authors
have used the argument of Goodenough and Kanamori to
discuss the types of exchange interaction (J) where a CoOCo
0
angle of less than 971 is assumed positive (ferromagnetic) and
anything above is negative (antiferromagnetic).
119121
This simple
rule appears to operate with success for the case of cobalt(II). As
the examples in Fig. 2 show that modes I and II are antiferro-
magnetic while all the edge-sharing modes III and IV are
ferromagnetic. The latter has been conrmed by neutron
studies on compounds, Co
2
(OH)PO
4
, NaCo
2
(H
3
O
2
)(MoO
4
)
2
(Fig. 2 (V)) and Co
2
(pm), displaying this type of connection
(see later).
122124
When the bridge is a carboxylate, that is
CoOCOCo, most, if not all, of the reported exchange
interactions are antiferromagnetic. However, the absolute values
can vary depending on the modes of connection (synsyn,
synanti or antianti), the angles and also on the distance. One
Fig. 2 Structures of binuclear units showing the dierent modes of
connection via one-atom (O) and three-atom (OCO) bridges.
Fig. 1 Structures containing (I) regular octahedra:
[Co(H
2
O)
6
](H
2
pm), (II) distorted octahedra: Co
2
(cis-1,4-chdc)
2
(dpa)
2
and (III) tetrahedra: Co
2
(2,2
0
-bpdc)
2
(dpa)
2
.
1356 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
exception to this, Kitagawa et al. reported a ferromagnetic
interaction (Y = +8.3 K; C = 4.39 cm
3
K mol
1
Co) for a
dimeric unit bridged by two synanti carboxylato bridges in the
bilayered polymer, Co(3,4-pydc)(H
2
O)
2
H
2
O (Fig. 2 (VI)), pydc
= pyridinedicarboxylate.
125
In the case of one, two or three
OCO bridges, the exchange interactions are usually weak but for
the four OCO bridges, as found in the paddlewheel dimers, it can
be very large. A CurieWeiss t of the susceptibility data of
Co
2
(2,5-Br
2
bdc)
2
(py)
2
(Fig. 2 (VII)) gives Y = 218 K while
using an isotropic dimer model for S = 3/2, we estimate J to
be 68 K. Ouahab et al. used a BleaneyBowers S = 1/2 model
instead and found a singlettriplet gap of 420 cm
1
(292 K) with
a tted g = 4.78.
126,127
To explain the observation, they propose
a molecular-orbital scheme, dierent to that of Cotton et al.,
suggesting the formation of a weak s-bond (d
z
2).
128,129
Using
hydrothermal synthesis with the chiral D-(+)-camphoric acid,
Zeng et al. obtained a chiral two-dimensional (4,4) network of
binuclear paddlewheels, M
x
M
0
2x
(D-cam)
2
(1,4-dimb), which
are then connected by 1,4-di(1-imidazolylmethyl)benzene
(1,4-dimb).
130
They were able to make solid solutions of CoNi
covering the whole range. Surprisingly all the compounds show
long-range antiferromagnetic ordering between 7 and 23 K as
well as weak ferromagnetism, spin-op, and glassy behaviour
that result from the randomness of the mixed metal pairs,
magnetic anisotropy of the metallic cations, and possibly anti-
symmetric exchange. Using the same camphoric acid and
Co(OAc)
2
but this time carrying out the reaction using
ionothermal conditions in 1-ethyl-3-methylimidazolium bromide
(EMImBr), the chiral (EMIm)[Co
2
(D-cam)
2
(OAc)] was obtained
where the (4,4) 2D-network of paddlewheels are bridged by
acetate (OCO) bridges to give a 3D-framework with the cations
sitting in the cavities.
131
No magnetic properties were reported.
Oxalate is another ligand that can be considered as a bridge
through two OCO connections, but it diers electronically by
the presence of the CQC bond which increase the interaction
within a pair due to the p-orbitals (Fig. 2 (VIII)). Dierent
groups have reported estimations of J and all are in agreement
for a weak antiferromagnetic interaction.
132134
3.3 Trimers
Similar types of connection as for dimers are found within
trimers. Magnetic measurements are available for some and
similar approach to analysing the data has been reported. The
elaborate analysis by Calvo-Pe rez et al. considering only a trimer
where the connection between cobalt atoms is via one oxygen
and two OCO bridges and ignoring the connection between
trimers for the chain compound Co
3
(m-OOCCF
3
)
4
(m-H
2
O)
2
)-
(OOCCF
3
)
2
(H
2
O)
2
(C
4
H
8
O
2
)2C
4
H
8
O
2
suggests that J is weakly
antiferromagnetic.
135
For the similar trimer Co
3
(PhCOO)
6
(py)
2
(Fig. 3 (I)) no magnetic data are given.
136
For Co
3
(phcinna)
6
(quin)
2
where phcinna = phenylcinnamate and quin = quinoline,
which has three OCO bridges between cobalt, Oka et al.
estimates a ferromagnetic J by analysing the temperature
and eld dependence of the magnetization.
137
On the other
hand, Yu et al. found antiferromagnetic coupling for a similar
trimer in Co(tatb)
2
(H
2
O)
2
(dma3H
2
O), tatb = 4
0
,4
00
-s-triazine-
2,4,6-triyltribenzoate, dma = dimethylacetamide.
138
Park et al.
made Co
3
(sda)
3
(dmf)
2
(Fig. 3 (II)) where the stilbenedicarboxylate
(sda) bridges trimeric units into hexagonal layers.
139
Each trimeric
unit consists of one octahedral cobalt atom at the centre
symmetrically connected to two tetrahedra through three
OCO bridges each. The Curie constant of 3.12 cm
3
K mol
1
Co was reported to be independent of the applied eld
(100 and 1000 Oe) but the Weiss constant is reported to
be slightly dependent on eld (31.32 K for H = 100 Oe
and 24.69 K for H = 1000 Oe) but negative and higher than
that of a monomer, suggesting that the nearest-neighbour
interaction is antiferromagnetic. Paramagnetism is
reported for a trimer where the connections are through
one oxygen and one carboxylate.
140,141
Unfortunately, no
magnetic data are available for the two edge-sharing trimers,
Co
3
(H
2
O)
6
(succinate)
3
(Fig. 3 (III)) and Na[Co
3
(1,3,5-btc)
2
(m
3
-OH)-
(m
2
-H
2
O)
4
(H
2
O)
7
]1.5H
2
O (Fig. 3 (IV)), btc = benzenetri-
carboxylate,
142,143
where one may expect a ferromagnetic J
between nearest-neighbours.
Using two dierent carboxylate containing ligands, 3-(1H-
benzimidazol-2-yl)propanoate (pa) and isonicotinate (ina), Yao
et al. prepared a 2D network, Co
3
(OH)
2
(pa)
2
(ina)
2
(Fig. 3 (V)),
consisting of Co(II) T
d
O
h
T
d
trimers.
144
Although the trimers
are connected through multiple-atom bridges, the compound is
reported to display long-range antiferromagnetic ordering at
20 K. From eld- and frequency-dependence ac-susceptibility
measurements, the authors conclude that the ground state is a
canted-antiferromagnet deriving from DzyaloshinskiMoriya
interaction in addition to a single-molecule magnetic
behaviour.
145,146
3.4 Tetramers
As the number of moment carriers within a cluster increases,
the magnetism becomes more interesting due to the possibility
of obtaining systems behaving as single-molecule-magnets
given that cobalt(II) has an important single-ion anisotropy
(D) due to the presence of three unpaired electrons in the
d-shell. A positive D is a requirement for the observation of
Fig. 3 Structures of trinuclear units showing the dierent modes of
connection via one-atom (O) and three-atom (OCO) bridges.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1357
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
single-molecule-magnetic properties and its absolute value is
related to the barrier for magnetization reversal.
35
Powell et al. made the rhombus tetramer,
Co
4
(m
3
-OH)
2
(H
2
O)
6
(ntp)
2
2H
2
O (Fig. 4 (I)) where
ntp = nitrilotripropionate, which consists of an edge-sharing
cobalt dimer connected by the apices of two octahedra at the
hydroxide sites.
147
The magnetic properties reveal competing
antiferromagnetic and ferromagnetic interactions between
the four Co(II) ions and below 1 K a long-range Ne el
state is established. Liu et al. reported a network,
Co
4
(1,3,5-btc)
2
(L
3
)
2
(OH)
2
(H
2
O)4.5H
2
O, L
3
= 1,1
0
-(1,4-butanediyl)-
bis(imidazole), consisting of two types of rhombus tetramers
(Fig. 4 (II)) composed of edge-sharing polyhedra.
148
In one
type an edge-sharing octahedral cobalt pair connects to the
edges of two trigonal-bipyramidal cobalt centres. The other
consists of four cobalt octahedra sharing edges. This
compound is found to display dominant antiferromagnetic
interaction (Y = 89 K) and exhibits a very high Curie
constant per cobalt of 6.2 cm
3
K mol
1
, which is twice the
expected value. In a slightly dierent structured tetramer,
Co
4
(OH)
2
(3,4-pbc)
3
(Fig. 4 (III)), 3,4-pbc = 3-pyrid-4-ylbenzoate,
consisting of an edge-sharing octahedral cobalt pair connected
to the apices of two trigonal-bipyramidal cobalt.
149
A
CurieWeiss t of the temperature dependence of the
susceptibility gives a Weiss constant of again 89 K and a
normal Curie constant of 3.5 cm
3
K mol
1
Co.
The 3D network of pyromellitate bridged tetramers
Co
4
(OH)
2
(H
2
O)
2
(C
4
H
11
N
2
)
2
(pm)1.5H
2
O,
150
consisting of an
edge-shared octahedral dimer connected via the apices to two
square-pyramids also displays dominant antiferromagnetic
coupling with features indicating the presence of ferromagnetic
coupling. Co
4
(2,5-pydc)
4
(bpe), pydc = pyridinedicarboxylate,
bpe = bipyridylethane, an eight-connected self-penetrating net
based on a cross-linked a-Po subnet, has tetrameric nodes based
on OCO bridged dimers.
151
It behaves as antiferromagnetically
coupled paramagnets (Y = 61 K and C = 1.48 cm
3
K mol
1
of Co which is only half the expected value). Ma and Zhou
synthesised a highly symmetric network, H
2
[Co
4
O(tatb)
8/3
]
(Fig. 4 (IV)) where tatb = 4,4
0
,4
00
-s-triazine-2,4,6-triyltribenzoate,
exhibiting high gas absorption capacity and consisting of a
beautiful Co
4
(m
4
-O) tetramer.
152
Each cobalt atom adopts a
square-pyramid geometry and sits at the corners of a perfect
square with the appearance of a St Johns cross. If the
exchange between pairs is antiferromagnetic, such an arrange-
ment will be frustrated magnetically and consequently, the
absence of magnetism of this compound will be very
interesting indeed.
Following the synthesis and characterisation of the
[M
4
(cit)
4
] cubane moiety (Fig. 4 (V)) in the series
[C(NH
2
)
3
]
8
[M
II
4
(cit)
4
]8H
2
O (cit = citrate, M = Mg, Mn,
Fe, Co, Ni, Zn),
153,154
the two independent groups performed
extended studies of the ac and dc magnetic properties covering
the milliKelvin (mK) temperature region, eld dependence,
eld-sweep dependence and frequency dependence.
155,156
They
arrived at the same conclusion of a SMM behaviour for the
cobalt complex with the characteristic blocking temperature of
3 K, DE/k
B
= 21 K, and t
0
= 8 10
7
s. Due to the very fast
quantum tunnelling of the magnetization, the hysteresis loops
collapse at zero-eld. Murray et al. also showed that the
related octanuclear cluster [Co
8
(cit)
4
(H
2
O)
12
]24H
2
O,
156
containing a [Co
4
(cit)
4
] cubane moiety, behaves likewise
(DE/k
B
= 20.5 K, t
0
= 1.0 10
7
s). Furthermore, for the two
parent compounds, [(NMe
4
)
3
Na{Co
4
(cit)
4
[Co(H
2
O)
5
]
2
}]11H
2
O,
[(NMe
4
)
4
{Co
4
(cit)
4
[Co(H
2
O)
5
]
2
}]16H
2
O, and their partially de-
hydrated congeners, [(NMe
4
)
3
Na{Co
4
(cit)
4
[Co(H
2
O)
5
]
2
}]7H
2
O,
[(NMe
4
)
4
{Co
4
(cit)
4
[Co(H
2
O)
5
]
2
}]6H
2
O, having the Co
4
O
4
cubane structure connected by two Co(H
2
O)
5
units, Murrie
et al. found SMM behaviour in their dc- and ac-susceptibility
study.
153
The fully solvated samples exhibit a blocking
temperature of 6 K with onset of a frequency dependence
but without a peak in the temperature dependence of the
imaginary ac-susceptibility (w
00
) above 1.8 K. In contrast,
for the partially dehydrated samples, well-resolved peaks for
w
00
are observed from which the gap and relaxation time (t
0
)
were estimated from the Arrhenius ts: DE/k
B
= 26 K and
t
0
= 8.2 10
9
s for [(NMe
4
)
3
Na{Co
4
(cit)
4
[Co(H
2
O)
5
]
2
}]
7H
2
O and DE/k
B
= 32 K and t
0
= 2.1 10
9
s for
[(NMe
4
)
4
{Co
4
(cit)
4
[Co(H
2
O)
5
]
2
}]6H
2
O. The change in co-
ordination geometry upon dehydration is suggested to be
responsible for the dierence in behaviour.
Fig. 4 Structures of tetranuclear units showing the dierent modes of
connection via one-atom (O) and three-atom (OCO) bridges and
face-sharing within Co
4
O
4
cubanes.
1358 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
Xiang et al. found another compound containing the Co
4
O
4
citrate cubane unit in {KCo
3
(cit)(Hcit)(H
2
O)
2
8H
2
O}
8
(Fig. 4 (VII)) which is obtained as violet prismatic crystals
by the hydrothermal reaction of a suspension of cobalt
acetate, citric acid and KOH in ethanolaqueous solution at
120 1C.
157
The structure is highly symmetric and is described
as anatase with an assembly analogous to polyoxometalate.
The key feature is the similar cubane units described above
where each cubane is connected to six neighbours via six
tetrahedral cobalt units; the latter connects four cubanes.
The overall structure was demonstrated to be porous by gas
sorption measurements. The Weiss constant is +1.4 K
suggesting ferromagnetic interaction dominates within the
cluster. Considerable eld dependence is observed in the
dc-magnetization below 20 K which is interpreted as being due
to canting with a remanent magnetization of only 92 cm
3
G mol
1
as expected for a very small canting angle of 0.131. Frequency-
dependent ac-susceptibility is observed below 5 K. The
dehydrated compound behaves in a similar way with almost
the same characteristics.
The Co
4
O
4
cubane was found in [Co
12
(bm)
12
(NO
3
)-
(OAc)
6
(EtOH)
6
](NO
3
)
5
(Fig. 4 (VIII)), where bm = benzimidazole
methanol caps the cubane partially and the coordinating NO
3
sits in the center of three cubanes while the acetate connects
the cubanes on the periphery to give a Co
12
cluster.
158
It
displays SMM behaviour below 3 K where the Argand-
diagram suggests a simple one-barrier model. The isothermal
magnetization at 2 K reaches full saturation in 70 kOe. The
compounds containing the Co
4
O
4
cube appear to be the
smallest true SMM to be realised though Yao et al. hinted
to a trimer having such properties.
144
Chiang et al. found a fusion of two Co
4
O
4
cubanes at one
vertex to give Co
7
(m
3
-OH)
8
(Fig. 4 (VI)) clusters which are
bridged by six oxalate and six piperazines to form a
highly symmetric (space group R

3) 3D-framework
Co
7
(m
3
-OH)
8
(ox)
3
(pip)
3
.
159
Analysis of the high-temperature
susceptibility gave C = 3.14 cm
3
K mol
1
Co and Y =2.76 K.
The latter indicates a nearest-neighbour ferromagnetic
interaction between cobalt atoms within the cluster. Due to
antiferromagnetic coupling between these ferromagnetic
clusters the ground state is an overall antiferromagnet
below T
N
of 26 K. Isothermal magnetization at dierent
temperatures are linear with eld and that at 2 K suggests
a spin-op about 60 kOe. The isostructural nickel analogue
is likewise composed of ferromagnetic clusters (dened by two
exchange pathways J
1
= 40 K and J
2
= 0 K),
160
which
are ordered antiparallel to one another to give a Ne el state
below 17 K.
3.5 Pentamers
Pentamers are rare; a search of the literature found only one
example, Co
5
(OH)
2
(pm)
2
(H
2
O)
4
xH
2
O.
161
It consists of
edge-sharing trimers with apical connection to two octahedral
cobalt atoms on the peripheral ends of the trimer. These
pentamers are linked into frameworks by the benzenetetra-
carboxylate. The link involves bridges consisting of the
carboxylate and the benzene backbone. No magnetic
measurements are reported for this compound.
3.6 Hexamers
Co
6
(OH)
2
(bpc)
2
(H
2
O)
6
(Fig. 5 (I)), bpc = 1,2,3,4,5-benzene-
pentacarboxylate, consists of a linear edge-sharing tetramer
connected to two octahedral cobalt atoms on the peripheral
ends of the tetramer.
116
The links between hexamers involve
only carboxylate connections. Extensive magnetic measure-
ments are reported for this compound demonstrating a long-
range antiferromagnetic ordering below 5 K. The Curie
constant is 3.35 cm
3
K mol
1
Co and the Weiss constant
is 30 K. The saturation magnetization at 2 K of 1.08 m
B
is far
from the expected result of 4.6 m
B
. These results indicate a
ferrimagnetic order within the linear hexamer and weak anti-
ferromagnetic coupling between them to assure a long-range
ordering to a Ne el state. This weak coupling can be overcame
by elds higher than the metamagnetic critical eld of 4 kOe to
realise a ferrimagnet where four ferromagnetically coupled
cobalt atoms are antiferromagnetically coupled to the
remaining two. From the phase diagram, Sevov et al.
established an ordering temperature of 4 K.
116
Using 2-phenylcinnamic acid the hexamer complex,
Co
6
(OH)
2
(phcinna)
10
(Fig. 5 (II)), phcinna = a-phenylcinnamate,
has been isolated and it consists of two edge-sharing octahedra
and four tetrahedral units linked through m-carboxylate
ligands and m-hydroxide ions.
162
Antiferromagnetic coupling
between Co ions is suggested by the temperature dependence
of the susceptibility while the isothermal magnetization at 2 K
indicates an S = 3 ground-state. Assuming the octahedral pair
is ferromagnetically coupled, the moments of the four
tetrahedra must be antiparallel resulting to a saturation of
6 m
B
. Co
6
(OH)
2
(PhCOO)
10
(PhCOOH)
4
3PhCH
3
(Fig. 5 (III))
Fig. 5 Structures of dierent hexanuclear units.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1359
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
contains a closely related hexamer but the magnetic properties
are not reported.
136
As a tripod ligand (phosphonomethyl)iminodiacetate was
found to coordinate via the oxygen atoms of the carboxylate
and phosphite end as well as by the amine nitrogen atoms to
cobalt(II) to form {K
2
[CoO
3
PCH
2
N(CH
2
CO
2
)
2
]}
6
xH
2
O
(Fig. 5 (IV)) by a solvothermal reaction in methanol.
163,164
While the carboxylate groups and the nitrogen block the
cobalt atoms from propagating into a 3D-polymeric structure,
the phosphite group connects penta-coordinated (CoO
4
N)
into a highly symmetric ring. These are then packed into a
3D network by hydrogen bonding with the lattice water
molecules and by ionic bonds with the potassium atoms.
Analysis of the magnetic susceptibility above 100 K,
taken in 100 Oe, gave a Weiss constant of 14 K and
considerable eld dependence below 20 K, consistent with
canted-antiferromagnetism, as conrmed by the linear
dependence of the isothermal magnetization at 5 K and the
presence of a small hysteresis loops. At rst sight, the presence
of a long-range ordering for such a ring is quite surprising but
if the moments within the rings are ferromagnetically coupled
a weak dipolar interaction between the rings can result in an
ordered state. It is further suggested by the authors that the
canting is derived predominantly by the single-ion anisotropy
due to the absence of an inversion centre relating the cobalt
atoms by symmetry.
Reaction of another tripod ligand, 2-amino-2-methyl-1,3-
propanediol (ampdH
2
), with Co(OAc)
2
in methanol aords
the mixed-valent hexamer Co
6
(H
2
O)(MeOH)(OAc)
6
(ampd)
4
(Fig. 5 (V)) compound (four divalent and two trivalent)
formed of a segment of a brucite layer.
165
At high tempera-
tures the magnetic susceptibility data are dominated by the
single-ion properties of high-spin Co(II) centres with distorted-
octahedral coordination geometries. Below 30 K ferro-
magnetic coupling was found using theoretical analysis in
terms of an anisotropic exchange model, and inelastic neutron
scattering data.
3.7 Heptamers
It is interesting to highlight the four examples of heptamers
though they do not strictly belong to our selection; that is
containing a carboxylate group. The rst is an interesting
cubane of cobalt [Co(H
2
O)
6
Co
8
(L1)
12
]X
6
nH
2
O (X = NO
3

,
n = 12; X = HCO
3

, n = 24; HL1 = 4,6-bis(2-pyridyl)-1,3,5-


triazin-2-ol) and inside the cube is a Co(H
2
O)
6
hydrogen-
bonded to the alcohol group of the ligand.
166
The unit is a
fragment of an NbO structure. Though there is a three-atom
bridge between cobalt pair in the cube, the magnetic properties
are close to that of isolated paramagnetic cobalt(II) centres.
The other three examples are all cyclic wheels of six cobalt
atoms on the perimeter and one in the centre (Fig. 6 (I)). This
cyclic wheel is a fragment of a brucite layer where each
octahedron shares three edges with its neighbours. They dier
in the valencies of the cobalt atoms and in their disposition
within the cluster. The rst contains seven divalent cobalt
atoms, while the second has four divalent ones (one at the
centre and three symmetrically positioned on the perimeter)
and three trivalent ones (on the perimeter). The third has three
divalent cobalt atoms (dening the diameter) and two pairs of
trivalent ones on each side of it.
The rst example, [Co
7
(hdeo)
6
(N
3
)
6
](ClO
4
)
2
2H
2
O contains azide
and oxo bridged Co(II) (Hhdeo = 2-hydroxy[1,2-di(pyridin-2-yl)]-
ethane-1-one).
167,168
Magnetic data were theoretically tted
and intramolecular ferromagnetic couplings between nearest
neighbours were found. It shows a blocking temperature of
ca. 3 K with clear frequency dependence of both components
of the ac-susceptibility in elds of zero and 600 Oe, suggesting
SMM behaviour.
The same structural unit is found in [Co
II
4
Co
III
3
(HL)
6
-
(NO
3
)
3
(H
2
O)
3
]
2+
by using the tripod ligand H
3
L = H
2
NC-
(CH
2
OH)
3
,
169
but this time the complex is a mixed-valent salt.
In contrast to azido bridges in the above compound, this
species has oxygen atom bridges. In this electronic conguration
only the Co(II) is the moment carrier and they are arranged in
a star of three corners and one centre. The complex is reported
to be an exchange-biased single-molecule magnet. The
blocking temperature is similar to that of the fully reduced
system.
Extending the tripod ligand to 2-amino-2-methyl-1,3-
propanediol (H
2
ampd) Ferguson et al. obtained the hepta-
nuclear ring structure when the reaction with Co(OAc)
2
was
performed solvothermally in methanol.
169
The cluster is
connected into chain by a cobalt(II) to give the overall formula
Co
8
(H
2
O)
2
(OAc)
7
(ampd)
6
, where the heptanuclear cluster
contains four divalent and three trivalent cobalt ions. The
magnetic data at high temperatures were analysed showing
dominant single-ion properties of high-spin Co(II) centres with
distorted-octahedral coordination geometries, while the
low-temperature data suggest ferromagnetic exchange within
the heptanuclear unit and negligible interactions along the
chain between the hepta- and mononuclear fragments.
Despite the presence of negative magnetic anisotropy of the
Co(II), a requisite for the observation of SMM behaviour,
Luneau et al. observed a lack of such behaviour for the
mixed-valent heptanuclear wheel, Co
II
3
Co
III
4
(L)
6
(MeO)
6
,
LH
2
= 1,1,1-triuoro-7-hydroxy-4-methyl-5-azahept-3-en-2-one.
170
Structural study showed that the three paramagnetic ions are
in a line along the diameter of the wheel. The temperature- and
eld-dependence magnetizations evidence ferromagnetic inter-
actions within the compound. Using ab initio calculations
including spinorbit coupling to simulate the susceptibility
and magnetization, good agreements were obtained for
nearest-neighbour and next-nearest-neighbour exchange para-
meters (1 and 3.8 K, respectively). Facile inter-valence
Fig. 6 Structures of two heptanuclear units.
1360 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
electron-transfer appears to contribute to the strong ferro-
magnetic interaction between distant cobalt ions. The absence
of single-molecule-magnet behaviour is associated with large
tunnelling rates between states in the presence of dipolar
transverse magnetic elds in the crystal.
KCo
7
(OH)
3
(1,3-bdc)
6
(H
2
O)
4
12H
2
O (1,3-bdc = isophthalate),
is another framework formed of a CoOCo connected Co
7
cluster and these MOPs are connected into a high-symmetry
(R

3c) system by the isophthalate.


171
Water molecules occupy
the space within the lattice and single-crystal-to-single-crystal
transformation is demonstrated after removal of the water
providing a void space (10 A

radius and 4.5 A

aperture) of
22% of the total volume. The Co
7
cluster (Fig. 6 (II)) has an
interesting topology where all the cobalt adopt octahedral
geometry in which four of them are edge-sharing forming a
plane and the other three are vertex-sharing with them as well as
between themselves. Each cluster has six neighbours. Analysis
of the magnetic susceptibility gave a Curie constant of
2.93 cm
3
K mol
1
Co and Weiss constant of 49 K. A plot of
the eective moment vs. temperature indicates a ferrimagnetic
behaviour. A possible coupling mechanism for a ferrimagnetic
arrangement is presented. Interestingly, ac-susceptibilities at
dierent frequencies indicate a single-molecule-magnet
behaviour with a blocking at around 6 K. The dehydrated
compound behaves as a paramagnet.
So far, the experimental results have established some
general trends in the pairwise magnetic interaction between
nearest-neighbour cobalt(II), whereby one-oxygen atom
bridges (vertex-sharing) generate antiferromagnetic inter-
action while two-oxygen bridges (edge-sharing) give rise to
ferromagnetic ones. In addition, it is observed that one, two or
three-atom bridges through carboxylate are weak interactions
and can be either ferromagnetic as in the case when the mode
of coordination is synanti or antiferromagnetic if they are
synsyn. In contrast, when there are four OCO bridges as in
a paddlewheel complex, the interaction can be large and
antiferromagnetic if the structure is regular but it can become
ferromagnetic if there are severe distortions.
4. Chains
Developing through to chains and ladders we observe the same
connections between the cobalt centres and therefore the
general trends, remarked above, for the nearest-neighbour
magnetic exchange interactions can be considered valid. As
for clusters where well-separated ones with a resultant moment
can behave as single-molecule-magnet, for chains single-chain-
magnetism has been observed as a result of the high
anisotropy of cobalt and consequently, realising one-dimensional
Ising systems following Glauber dynamics.
172
We will follow
as above by describing the results from the simpler to the more
complex systems, as there are many fascinating examples of
chain compounds in the literature.
The doubly carboxylate bridged chain of ve-coordinated
cobalt in Co(2,5-dibromoterephthalate)py (Fig. 7 (I)) is obtained
hydrothermally from cobalt chloride and dibromoterephthalic
acid neutralised with pyridine.
126
It behaves as a paramagnet
with C = 2.76 cm
3
K mol
1
Co and Y = 14.9(6) K.
A similar conclusion was obtained for the single carboxylate
bridged chain in Co
2
(mpda)(H
2
O)
6
H
2
O (Fig. 7 (II)),
mpda = 1,2,3,4-benzenetetracarboxylate.
173
The saturation
magnetization at 2 K is 2.35 m
B
. Similarly, a simple chain of
octahedra is formed by a combination of a one-atom (O)
and a three-atom (OCO) bridge which are separated by very
bulky ligand in the compound, Co
2
(2,5-diphenylterephthalate)
2
-
(H
2
O)
2
(Fig. 7 (III)).
126
The susceptibility (C = 3.12 cm
3
K mol
1
Co and Y = 28.2 K) suggests weak antiferromagnetic
coupling within the chain. A similar conclusion was reported
for {[Co(dpyo)(1,4-bdc)(H
2
O)
2
][Co(H
2
O)
6
](1,4-bdc)H
2
O}
(Fig. 7 (IV)), dpyo = 4,4
0
-bipyridyl N,N
0
-oxide, by Manna
et al.
174
A closely related chain was observed for the trimesate
containing compound K[Co
3
(1,3,5-btc)(1,3,5-Hbtc)
2
]5H
2
O
(Fig. 7 (V)) where the octahedra were found to be more
regular.
175
The CurieWeiss t of the susceptibility data gave
C = 3.1 cm
3
K mol
1
Co and Y = 0 K; indicating the nearest-
neighbour interaction may be ferromagnetic. However,
replacing the single-atom oxygen bridge by that of a nitrogen
atom from azide, as has been reported by Gao et al. for
((CH
3
)
2
NH
2
)[M(N
3
)
2
(HCOO)],
176
M = Co or Fe, the isolated
chain is obtained. These compounds exhibit metamagnetism
because of the strong intrachain ferromagnetic and weak
interchain antiferromagnetic couplings. The latter may be
due to pp interaction between the azides of neighbouring
Fig. 7 Structures of single-strand chains with dierent modes of
connection.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1361
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
chains. Nataraj et al. reported a low-dimensional magnetic
behaviour for the edge- and vertex-sharing chain of cobalt,
(piperazinium)[Co
4
(HPO
3
)
2
(ox)
3
].
177
Two simple chain compounds of formula, Co(PhCOO)
2
, are
formed with the benzene ring of the benzoate isolating the
chains magnetically from each other.
136,178
They are composed
of alternating octahedra and tetrahedra cobalt centres bridged
by one oxygen atom and one carboxylate (Fig. 7 (VI)). The
chain is therefore zigzag in nature. While the orthorhombic
(Pcab) modication is reported to be a long-range ordered
ferromagnet with a Curie temperature of 3.7 K, the monoclinic
(C2/c) modication remains paramagnetic to very low
temperature and become a SCM below 0.6 K. Both have very
anisotropic susceptibilities. For the former the easy axis is not
parallel to the chain axis but for the latter it is parallel.
However, there is no isothermal magnetization of the ortho-
rhombic phase to justify the ferromagnetism. If the moments
within the chain are antiferromagnetically aligned and all the
chains are oriented in the same direction, a ferrimagnet will
be most likely due to the dierence in g-values. There seems to
be no structural dierence within the chain to explain the
dierence in magnetism.
Co(4,4
0
-bpdc)(H
2
O)
2
H
2
O, 4,4
0
-bpdc = 4,4
0
-biphenyldi-
carboxylate, is the only one of the MOFs to have a linear
edge-sharing chain (Fig. 7 (VII)).
62
The connection at the edge
is via two oxygen atoms of water molecules while one of the
carboxylate group bridges nearest-neighbour cobalt and the
other is hydrogen-bonded to the coordinating water molecules
of adjacent chains. The nearest-neighbour interaction was
found to be ferromagnetic (Y = +26 K) and an antiferro-
magnetic ordering is observed at 5.5 K which was conrmed
by heat capacity measurements.
179
A metamagnetic transition
is observed at 2 K with a critical eld of 500 Oe. This kind of
chain, but without the bridging carboxylate, is found in mineral
derivatives such as natrochalcite, NaCo
2
(H
3
O
2
)(MoO
4
)
2
,
and adamite, Co
2
(OH)PO
4
,
122,123
where neutron diraction
studies have shown that the moments are indeed parallel to the
chain axis in both cases and the long-range ordering is
antiferromagnetic due to the presence of antiferromagnetic
exchange pathways between chains.
There are three known compounds with zigzag chain of
edge-sharing octahedra, namely M(1,4-napdc) (Fig. 7 (VIII))
(M = Co or Mn; 1,4-napdc = napthalene-1,4-dicarboxylate),
170
M
2
(dobdc)(H
2
O)
2
8H
2
O (Fig. 7 (IX)) (M = Co, Ni or Zn;
dobdc = 2,5-dioxyterephthalate)
181,182
and their dehydrated
forms (Fig. 7 (X)) and M
2
(pm) (Fig. 7 (XI)) (M = Co, Mn or
Fe; pm = 1,2,4,5-benzenetetracarboxylate).
63,183,184
The rst
consists of pair-wise zigzag chains which are well separated by
the bulky naphthalene and for the second the zigzag chains
have a periodicity of three octahedra and are also separated
form each other by the benzene ring. In contrast, the third has
a dimer periodicity and the chains are connected by OCO
bridges to give a 2D-network which are separated by the
benzene. These slight dierences, though subtle, exert major
dierences in the magnetic properties. For Co(1,4-napdc) and
Co
2
(pm) the CurieWeiss t of the high-temperature data gave
positive Weiss constants of +27 and +16 K, respectively. On
the other hand that of Co
2
(dobdc)(H
2
O)
2
8H
2
O is negative
(6 K). All these values suggest that the exchanges within the
chains in each of these compounds are ferromagnetic.
However, due to transverse antiferromagnetic coupling
between them, long-range antiferromagnetic order sets in at
5.5, 8 and 16 K, respectively. However, in each case a eld high
enough to overcome the transverse coupling energy reverses all
the moments to a ferromagnetic state and the compounds then
exhibit hysteresis. The critical metamagnetic elds are
quite dierent, 250 Oe for Co(1,4-napdc), 430 kOe for
Co
2
(dobdc)(H
2
O)
2
8H
2
O and 1400 Oe for Co
2
(pm). The large
variation in critical eld does not scale with the interchain
distance as one may expect. Most unexpectedly, the
ac-susceptibility of Co(1,4-napdc) and Co
2
(pm) shows the
presence of both the real and the imaginary components. A
spontaneous magnetization was conrmed for Co
2
(pm) by
zero-eld and eld-cooled measurements in a eld of 1 Oe
which demonstrated the presence of a canted state at 13 K
below the collinear Ne el state at 16 K (Fig. 8). The lack of such
measurements for the other two compounds limits a full
comparison. Heat capacity measurements for Co
2
(pm) esti-
mates that 60% of the entropy is found below the Ne el
temperature and the eective spin of the cobalt is 1/2.
Unpublished neutron diraction data of Co
2
(pm) in zero eld
show that the moments are perpendicular to the chain and the
nearest-neighbour chains connected by OCO bridges point
also in the same direction while those of adjacent layers are
antiparallel.
124
However, in a eld of 40 kOe all the moments
Fig. 8 ZFCFC magnetization (top) and isothermal magnetization at
dierent temperatures in the three regions of the phase diagram
(bottom) of Co
2
(pm).
1362 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
are parallel to one another and points perpendicular to the
chain axis. This is as expected from the observed saturation
moment of a ferromagnet. No LRO has been found for the
isostructural Mn(1,4-napdc) while Ne el transitions are
observed for Mn
2
(pm) (18 K) and Fe
2
(pm) (26 K). There are
no magnetic data for Ni
2
(dobdc)(H
2
O)
2
8H
2
O. It is also to
be noted that Co
2
(dobdc)(H
2
O)
2
8H
2
O can be reversibly
dehydrated/rehydrated without loss of crystallinity and gases
such as hydrogen and carbon dioxide have been loaded in the
channels. Again it will be of interest to study the dehydrated
Co
2
(dobdc) which now has the cobalt in square-pyramidal
coordination geometry and also the absence of any exchanges
through hydrogen-bonds.
Co
2
Na(4-cpa)
2
(m
3
-OH)(H
2
O) (Fig. 7 (XII)) (4-cpa = 4-carboxy-
phenoxyacetate) was obtained hydrothermally and it is
constructed of zigzag chains of edge-sharing octahedra.
These chains are separated by the sodium cations and the
organic ligands.
185
The analysis of the high temperature
dc-susceptibility data gave C = 3.61 cm
3
K mol
1
Co and
Y =56 K. At low temperature blocking of the moments was
observed which gave a clear ZFCFC bifurcation very close to
the blocking temperature of 4 K and also the magnetization
attained almost the saturation value. Field dependent suscept-
ibility is encountered well above the blocking temperature. A
minimum is wT was interpreted as due to ferrimagnetic
behaviour. Most interestingly, the ac-susceptibility at dierent
frequencies clearly exhibits slow relaxation eect. The
relaxation time was obtained from the Arrhenius law
t(T) = t
0
exp(2DE/k
B
T) with t
0
= 2.8 10
10
s and
DE/k
B
= 106 K. A constricted hysteresis loop is observed at
2 K with coercive eld of ca. 6 kOe and a saturation
magnetization at 70 kOe of 2.88 m
B
consistent with a non-
compensated resultant moment within the chain.
Kumagai et al. observed chains of mixed edge- and vertex-
sharing octahedra within the isostructural compound
M
3
(1,3,5-chtc)
2
(m-H
2
O)
2
(H
2
O)
2
5H
2
O (Fig. 7 (XIII)) (M =
Co, Ni), where 1,3,5-chtc is the exible cis,cis-1,3,5-cyclo-
hexanetricarboxylate.
186
The chains consist of trinuclear motifs
of pseudo-octahedra linked through m-OH
2
and m-carboxylate.
The temperature dependence of the magnetic susceptibilities
(Fig. 9) indicates dominant antiferromagnetic interaction within
the chain of the cobalt complex (C = 2.9 cm
3
K mol
1
Co;
Y =27 K) tending towards a ferrimagnetic alignment at low
temperatures due to the uncompensated moment with two of
one kind and one of the other. For the nickel complex the
interaction is ferromagnetic (C = 1.35 cm
3
K mol
1
Co; Y = 6
K) with a possible antiferromagnetic order below 5 K.
Zheng et al. obtained a layered structure consisting of
magnetically isolated chains of paddlewheels connected
by 1,2-cyclohexanedicarboxylate in Co
II
(trans-1,2-chdc)
(Fig. 7 (XIV)) where the dimers are connected by edge-sharing
via the oxygen of the carboxylate.
187
This mode of connection
is classied as at-ribbon for the copper carboxylate
derivatives.
128,129
The inter-dimer CoOCo angles are ca.
1001 and the octahedra are clearly distorted with two short,
two intermediate and two elongated CoO distances. These
distortions are known to aect the magnetic exchange within
the dimers. Most interestingly Co
II
(trans-1,2-chdc) was
found to display a single-chain-magnetic behaviour at low
temperatures. The susceptibility data above 30 K are
unexpected, Curie constant of C = 2.83 cm
3
K mol
1
Co
and Weiss constant of +15.87 K, t well to a ferromagneti-
cally coupled alternating chain (S = 3/2) with intrachain
J
1
= +11.51 K, J
2
= +3.95 K and a mean-eld inter-chain
J
0
= 0.01 K. Considerable anisotropy was shown by
the measurement of the susceptibility on a single crystal to
demonstrate the presence of an Ising one-dimensional ferro-
magnetic chain. Consequently, a SCM behaviour was realized
as demonstrated by the frequency-dependent ac-susceptibility
around the blocking temperature of 6 K. From the Arrhenius
plot, two regimes were identied where the relaxation times
are 5.1 10
11
and 5.5 10
8
s for the high- and low-
temperature regions, respectively, and two dierent corres-
ponding barriers (DE/k
B
= 80.9 and DE/k
B
= 50.2 K). A
hysteresis loop with a constriction at zero-eld was observed at
several temperatures.
More complexities in the chain-structures are brought about
by the dierent modes of connection and these have sometimes
been described as ribbons or strips. The following
section details some of these compounds. A 3D-network
Co
3
(bime)
2
(m
3
-OH)
2
(HO-BDC)
2
(Fig. 10 (XV)), bime = 1,2-
bis(imidazol-1-yl)ethane, HO-BDC = 5-hydroxyisophthalate,
consisting of chains of alternate edge-sharing octahedra and
edge-sharing ve-coordinated cobalt dimers separated by the
ligands.
188
This fully edge-sharing chain behaves as a ferri-
magnet. From ac-susceptibility measurements using a range of
oscillating frequencies substantial dependence was observed
demonstrating a single-chain-magnetic behaviour. The
hysteresis loop is quite wide (22 kOe at 1.77 K) and the
saturation magnetization of 2.35 m
B
is consistent with
one cobalt resultant moment in a ferrimagnetic state.
Staying with chains of edge-sharing octahedra,
Co
3
(OH)
2
(C
4
O
4
)3H
2
O (Fig. 10 (XVI)) contains a strip
segment of a brucite layer.
189,190
These are connected to each
other by the squarate dianion that provide four-atom bridges
(OCCO) between the cobalt. Due to the sharing of three edges
at each hydroxide centre the chain is strictly at. The 3D
structure contains channels occupied by three water molecules
per formula unit and these can be removed and reinserted
repeatedly without aecting the framework as has been shown
Fig. 9 Temperature dependence of wT for M
3
(1,3,5-chtc)
2
-
(m-H
2
O)
2
(H
2
O)
2
5H
2
O, M = Co (blue), Ni (red).
186
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1363
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
by in situ single-crystal crystallography (Fig. 11). Analysis of
the susceptibilities of the hydrated, dehydrated and
rehydrated forms gave C = 3.32 0.02 cm
3
K mol
1
Co and
Y = +0.2, +2.7 and 0.0 K, respectively, suggesting ferro-
magnetically coupled cobalt within the chain. The virgin sample
orders as a linear antiferromagnet at 8 K and a canting is
observed at 6 K associated with the bifurcation seen in the
ZFCFC measurements. Upon dehydration the sample orders
as a ferromagnet at 8 K (Fig. 12). Ac-susceptibilities and heat
capacity conrm the magnetic ordering. The original magnetic
state is regained upon rehydration. Heat capacity conrms the
LRO and the lack of frequency-dependent ac-susceptibility
eliminates the possibility of SCM behaviour. The nickel
analogue is an antiferromagnet below 7.4 K.
Among the most abundant of linear chains within MOFs is
the edge- and vertex-sharing strip of a [110] layer of a rutile
(Fig. 10 (XVIIXIX)) having the general formula,
M
3
(OH)
2
(dicarboxylate)
2
(H
2
O)
4
nH
2
O. It is found for several
metals (Co, Ni, Mn and Zn) and is bridged by several
dicarboxylate, including fumarate, both cis- and trans-1,4-
cyclohexanedicarboxylate and 4,4
0
-biphenyldicarbarboxy-
late.
108,125,191194
For nickel compounds all the sites adopt
octahedral coordination but for cobalt the edge-sharing pair
can be either octahedral or square-pyramidal while for zinc
they are tetrahedral. The simplest case is when the cobalt
coordination are all octahedral thus the chain is very regular,
though there is a slight bend in the chains due to the m
3
-OH.
The dicarboxylate connects these chains into three-dimensional
structures and in some cases water molecules occupy the space
(Fig. 13). The magnetism in most cases, cobalt and nickel, is
dominated by antiferromagnetic interaction while the ferro-
magnetic interaction between the edge-sharing pairs are not
detected in the CurieWeiss analyses. For cobalt and nickel
the chains are ferrimagnetic and this is seen as a prominent
minimum in the wT vs. T plot. However, at temperatures below
the minima the moments are saturated at a certain tempera-
ture, for example for the fumarate it is 13 K (Co) and 6 K (Ni)
and for cis-1,4-chdc (Fig. 10 (XVII)) it is 10.2 K (Co) and
2.1 K (Ni). The magnetic behaviour of the nickel compounds
behaves in an expected way as for classical ferrimagnets but
that of the cobalt compounds does not. For example the
ZFCFC measurements of the cobalt fumarate compound
shows a weak spontaneous magnetization at the Ne el
transition while that of 1,4-chdc behaves in a similar way as
that of Co
2
(pm) discussed above, where the collinear state is
established at 10.2 K and a canting takes place at 9.3 K.
193,63
Temperature and eld dependence magnetization measure-
ments on powders are quite confusing while those on one
aligned single crystal along the three crystallographic axes
suggest a magnetic easy-axis perpendicular to the
Fig. 11 Structure of Co
3
(OH)
2
(C
4
O
4
)3H
2
O showing the squarate
connected brucite strips with the hydrogen-bonded water molecules
within the channels (yellow).
Fig. 12 In situ eld-cooling magnetization of Co
3
(OH)
2
(C
4
O
4
)3H
2
O
in its virgin, dehydrated and rehydrated forms.
Fig. 10 Structures of multiple-strand chains with dierent modes of
connection.
1364 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
crystallographic chain-axis and very unusual metamagnetic
behaviour along the easy axis whereas normal ferromagnetic-
like behaviour is shown perpendicular to it (Fig. 14). Further
work is in progress to elucidate the origin of this character,
which has never been encountered in the eld of magnetism.
Another interesting aspect of this structure is that the water
molecules can be removed upon heating and in the case of the
nickel compound it was found to reversibly transformed from
a ferrimagnet at T
C
of 2.1 K for the virgin sample to a
ferromagnet at T
C
of 4.5 K (Fig. 15).
192
Tong et al. made the
corresponding compound containing 3,4-pyridyldicarboxylate,
Co
3
(OH)
2
(3,4-pydc)
2
(H
2
O)
2
(Fig. 10 (XVIII)), where the
structure is analogous except for the coordination of the
pyridine to the cobalt taking the position of one water
molecule.
125
There is no space to accommodate any solvent.
The magnetic behaviour was described as a ferrimagnet with
T
C
of ca. 15 K. The hysteresis loop for a powdered sample was
constricted at zero eld and may be a result of dierent
orientations of the crystallites.
Co
3
(OH)
2
(btca)
2
3.7H
2
O and its dehydrated phase
Co
3
(OH)
2
(btca)
2
(Fig. 10 (XIX)), btca = benzotriazole-5-
carboxylate, were reported to have the sra network
topology (analogous to that of SrAl
2
).
195
It contains chains
where edge-sharing pairs of two square-pyramids are
vertex-sharing with an octahedral cobalt and the hydroxide
functioning as a m
3
-bridge as in the examples above. The Weiss
constant is quite high (58 K) but negative suggesting the
antiferromagnetic exchange dominates over the ferromagnetic
one. The hydrated phase is reported to order magnetically at 8 K
as a ferrimagnet and single-chain-magnet-like behavior is observed
by ac-susceptibility while the dehydrated phase shows eld-
induced metamagnetism from a compensated antiferromagnet
(T
N
= 4.5 K) to a ferrimagnet. The coexistence of long-range
ferrimagnetism and single-chain-magnetism appears to be
contradictory, as the latter is not supposed to be ordered at all.
However, the frequency dependence of the ac-susceptibility
takes place below the ordering temperature of 8 K and so
it is most likely due to domain eects rather than SCM
characteristics.
Pink needle crystals of Co
4
(phcinna)
6
(OH)
2
(H
2
O)
4
2H
2
O
(Fig. 10 (XX)), phcinna = a-phenylcinnamate, shows the
formation of chains consisting of alternate orthogonal edge-
sharing pairs where the adjacent pairs are connected at the
vertices by a hydroxide and carboxylate oxygen atoms.
196
It is
a diagonal strip of a [110] rutile layer. These chains are not
connected by bonds and the minimum distance between the
centre of one chain and its neighbours is 14 A

. It exhibits a
long-range antiferromagnetic ordering at 5.5 K. Due to the
isolation of the chain with its neighbours a very small eld of
25 Oe is enough to reverse the moments. However, the
moment at saturation is only a third of that expected if the
moments of all the carriers were aligned parallel to the eld.
Co
3
(2,4-pydc)
2
(m
3
-OH)
2
5H
2
O and Co
3
(2,4-pydc)
2
(m
3
-OH)
2
-
(H
2
O)7H
2
O (Fig. 10 (XXI)), 2,4-pydc = pyridine-2,4-
dicarboxylate, have been hydrothermally synthesized and
shown to have 3D frameworks with hydroxide-bridged metal
chains.
197
The chain is constructed by vertex-sharing quad-
rangles formed via edge-sharing triangles. Magnetic studies
show that the former exhibits spin-canted antiferromagnetism
and a eld-induced spin-op transition while the latter behaves
as a normal antiferromagnet. The magnetic properties are
largely unchanged by removal of the water molecules from
Fig. 14 Temperature dependence of the susceptibility and isothermal
magnetization of Co
3
(OH)
2
(cis-1,4-chdc)
2
(H
2
O)
4
2H
2
O along its three
crystallographic axes. See Fig. 13 for referencing the orientations.
Fig. 13 Structure of Co
3
(OH)
2
(cis-1,4-chdc)
2
(H
2
O)
4
2H
2
O showing
the connected rutile strips with the hydrogen-bonded water molecules
within the channels (yellow).
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1365
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
the channels. Li et al. prepared Co
3
(2,4-pydc)
2
(m
3
-CH
3
O)
0.5
-
(m
3
-OH)
1.5
(H
2
O)4.5H
2
O and solved its crystal structure. It
contains a 3D-structure with a modied arrangement of the
octahedra within the chains.
198
There are tetrameric edge-
sharing units connected at the vertex with an edge-sharing
dimer. It is to be noted that part of the hydroxide is replaced
by methoxide. If the reaction is carried out in ethanol the fully
hydroxide substituted complex is obtained. The magnetic
properties indicate ordering as an antiferromagnet below
6 K. A slightly more complex chain structure is found in
Co
5
(m
3
-OH)
2
(pm)
2
(bpp) (Fig. 10 (XXII)), bpp is 1,3-bis(4-
pyridyl)propane, where three octahedral cobalt centres share
edges and these are then bridged by two ve-coordinated
cobalt centres.
199
These chains are bridged by the pyromellitate
and the bending bpp ligand to give a 3D-framework. In an
applied eld less than 3.5 kOe, it behaves as a compensated
antiferromagnet and in elds exceeding the metamagnetic
critical eld of 4 kOe a ferrimagnetic alignment is found. The
saturation moment amounts to that of one cobalt (2.4 m
B
). It
suggests that the ferrimagnetism is due to the three octahedral
cobalt atoms having their moments parallel while they are
antiparallel to those of the two ve-coordinated cobalt centres. The
Ne el temperature is 12 K. Cheetham et al. obtained a cobalt
succinate, Co
7
(H
2
O)
3
(OH)
6
(C
4
H
4
O
4
)
4
7H
2
O (Fig. 10 (XXIII)),
having a hexamer cluster bridged by a monomer forming a
chain. No magnetic properties of the complexes are given.
200
The exchange interaction between two paramagnetic centres
through an oxalate bridged has been of much interest in the past
and several theoretical and experimental studies have been
performed to evaluate its magnitude and sign. Many chains and
layers containing such bridges are known to display long-range
magnetic ordering. This remains a matter for debate amongst
experimentalists and theoreticians given that one-dimensional
magnetic chain should not show long-range magnetic ordering
at nite temperature.
201,202
There are two types of chains that can
be formed with cobalt octahedra and oxalate bridges while there
are none made of tetrahedra. The rst type is a linear one with the
oxalate bridges in trans-position and the other is zigzag with
oxalate in cis-position.
174,203217
A whole gamut of ligands is
known to bond to the remaining two sites of the octahedron.
For the former (Fig. 16 (XXIV)) the ligands are monodentate
(water, dipyridylethane, dipyridylethylene, 4,4
0
-bipyridine,
4,4
0
-bipyridine N-oxide, piperazine, aminopyridine) and for the
latter (Fig. 16 (XXV)) they are usually chelating (2,2
0
-dipyridyl-
amine, 1,3-propanediol, 3-hydroxypyridine, isoquinoline,
adenine, purine and oxalate). All the known compounds show
antiferromagnetic exchange interaction between nearest neigh-
bours which is characterized by a broad hump in the
temperature dependence of the susceptibility. In several cases
the latter has been modelled. The diaquo compound was the
rst to be studied and was found by magnetization, EPR and
heat capacity measurements as well as neutron scattering to be
a collinear antiferromagnet below 6.3 K. It was modelled to a
one-dimensional Ising magnet where the intrachain nearest-
neighbour exchange is 30 K and the ratio of the interchain to
intrachain is 0.003. The other linear chain with aminopyridine is a
canted antiferromagnet at 8 K. In contrast to the linear chains, there
are more examples of canted antiferromagnets for the zigzag
compounds,
211216
for example those with 1,3-propanediol
(10.6 K), 4,4
0
-bipy (13 K), 2,2
0
-dipyridylamine (7.5 K) and oxalate
(Fig. 16 (XXVI)) in K
2
Co(ox)
2
(37 K). The high transition
temperature for the latter may be due to the ionic-bonded network
formed by the potassium and the non-coordinated oxygen
atoms of the oxalate. It is interesting to note the observation of
an SCM behaviour for the alternate mixed-metal complex
[(C
12
H
24
O
6
)K][(C
12
H
24
O
6
)(FC
6
H
4
NH
3
)][Co(H
2
O)
2
Cr(ox)
3
]
2
where the bulky ions and crown ether sits in between layers
consisting of hydrogen-bonded zigzag chains.
217
Co
4
(pico)
4
(4,4
0
-bpy)
3
(H
2
O)
2
2H
2
O (Fig. 16 (XXVII)), pico
= hydroxypicolinate, is closely related to the oxalato-bridged
chain discussed above except that one of the arms is a ve-
membered ring chelate as opposed to four-membered.
218
The
chains are bridged by the 4,4
0
-bpy into two types of layers
which are interpenetrated to give the 3D-structure. The chains
are magnetically isolated from each other. The magnetic
properties are similar to those of Co
2
(pm) where a very sharp
drop in susceptibility is observed at 3.5 K due to collinear
antiferromagnetic ordering.
63
Although the canting has not
been observed, the ac-susceptibility suggests there is a
spontaneous magnetization. The isothermal magnetization
establishes a metamagnetic critical eld of 160 Oe and shows
Fig. 15 Temperature dependence of the susceptibility of Ni
3
(OH)
2
-
(cis-1,4-chdc)
2
(H
2
O)
4
2H
2
O in its virgin, dehydrated and rehydrated
forms.
189
1366 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
a superposition of hysteresis loops analogous to that of
Co
2
(pm) where the saturation magnetization is not reached
even in 70 kOe at 2 K. Using hydroxyphosphonoacetic acid
(pa) neutralized with ethylenediamine (enH
2
)[Co
2
(pa)
2
(H
2
O)
2
]
was obtained and it contains a similar bridged chain which are
then connected by the PO
3
unit into layers. It behaves as a
canted-antiferromagnet below 3.3 K.
219
5. Ladders
Well-isolated ladder structures in the MOF family are very
rare. There are three known systems, Na
2
Co
2
(ox)
3
(H
2
O)
2
(Fig. 17 (I)), Co
3
(2,5-pydc)
2
(m
3
-OH)
2
(OH
2
)
2
(Fig. 17 (II)),
pydc = pyridinedicarboxylate, and Co(C
8
H
8
O
4
) (Fig. 17 (III)),
that can be classied in this category.
220222
Magnetic studies
are available for only the rst two. Na
2
Co
2
(ox)
3
(H
2
O)
2
has
one oxalate forming the rung and a OCO bridge for the legs.
These two-leg ladders are separated by sodium cations.
Co
3
(2,5-pydc)
2
(m
3
-OH)
2
(OH
2
)
2
has MOM legs and OCO
rungs. The legs are closely related to the Co
3
(OH)
2
rutile strip
presented above except for the fact that two of the octahedra
are now replaced by distorted ve-coordinated cobalt atoms.
All the carboxylate oxygen and the pyridine nitrogen atoms of
the 2,5-pyridinedicarboxylate are involved in the coordination
to the cobalt atoms. For the third example, Co(C
8
H
8
O
4
), the
carboxylate groups of 4-cyclohexene-1,2-dicarboxylate forms
the OCO connections of the legs and rungs of the ladder.
The rst reported susceptibility of Na
2
Co
2
(ox)
3
(H
2
O)
2
exhibits a broad maximum (21 K) which was associated with
short-range correlations, and a phase transition to a long-
range ordered Ne el state at 9 K.
220
Above T
N
the data are
tted to several models which suggest the intra-ladder
couplings are of the order 2.83.5 K and the inter-ladder
coupling is approximately 0.14 K. Later magnetic measurements
eliminate the presence of long-range magnetic ordering
(Fig. 18) and specially those on aligned single crystals were
used to t the low-temperature data to a S = 1/2 spin-ladder
model giving a gap of 16 K for H
8
and 13 K for H
>
.
223,224
There was no anomaly in the ZFCFC measurements and the
two superpose for the whole temperature range (2300 K). A
broad hump at B10 K was observed in the heat-capacity data
which was tted using the same model to give a gap of 20 K.
Doping the crystals with 10% Zn the broad maximum in the
susceptibility is still present while a very clear bifurcation at
10 K in the ZFCFC magnetization in 100 Oe was interpreted
as being due to spin-glass behaviour.
225
The analogous nickel
and iron compounds are also known and their magnetic
properties reported.
The magnetism of Co
3
(2,5-pydc)
2
(m
3
-OH)
2
(OH
2
)
2
is more
complex but has some similarity to that of the rutile Co
3
(OH)
2
chains discussed above.
221
A long-range antiferromagnetic
state with a small canting is developed at 30 K in low applied
elds which corresponds to antiferromagnetically coupled
ferrimagnetic chains. In the absence of a neutron diraction
study one cannot be certain if the antiferromagnetism is
between individual legs within a ladder or between ladders.
Several steps were observed in the isothermal magnetization
measured on a polycrystalline sample and has been described
as multiple bistability. At elds higher than the metamagnetic
Fig. 17 Structures of three known ladders.
Fig. 16 Structures of oxalate-bridged chains and one with hydroxy-
picolinate bridge.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1367
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
critical eld of 15 kOe, the moments are reversed. For such a
high critical eld compared to those described above it is quite
likely that the reversal of moment is taking place within a
ladder. From the saturation value of 3 m
B
mol
1
, a ferri-
magnetic state is the nal state. All the behaviour is similar to
the rutile chains described earlier. The interesting point is that
the transition temperature is 23 times that of single-strand
chains.
6. Layers
The most condensed of the layers in terms of MOM con-
nections is that of brucite. There are no example of a purely
brucite layer in the MOFs. It does exist for Co(OH)
2
and for
two others, Co
2
(OH)
3
NO
3
which contains a coordinated NO
3
in
the place of one OH and Co
5
(OH)
6
(SO
4
)
2
(H
2
O)
4
(Fig. 19 (I))
which is pillared by OSO
3
Co(H
2
O)
4
O
3
SO at the missing
OH sites.
226231
The interlayer distances are therefore
increased from 4.7 A

for Co(OH)
2
to 6.95 A

for Co
2
(OH)
3
NO
3
and 10.3 A

for Co
5
(OH)
6
(SO
4
)
2
(H
2
O)
4
. Modication of the
brucite layer is observed with periodic vacancies in the layer
where each vacant position is then replaced by tetrahedral
units above and below the layer. The number and disposition
of the tetrahedra vary from case to case. Only four such
compounds, Co
5
(OH)
8
(trans-1,4-chdc)4H
2
O (Fig. 19 (II)),
232
[Co
7
(OH)
12
(H
2
O)
2
](C
2
H
4
(SO
3
)
2
) (Fig. 19 (III)),
233
and
Co
8
(OH)
12
(SO
4
)
2
(diamine)nH
2
O (Fig. 19 (IV)),
234,235
where
n = 3 for ethylenediamine or 1 for DABCO (C
6
N
2
H
12
) have
been fully characterised by single-crystal X-ray diraction,
while a large number are inferred to be isostructural from
XRPD. Co
5
(OH)
8
(trans-1,4-chdc)4H
2
O has adjacent layers
bridged by the dicarboxylate via the apices of the tetrahedral
units and similarly for Co
8
(OH)
12
(SO
4
)
2
(diamine)nH
2
O by
the diamines. XRPD of both compounds show they can lose
water of crystallization without loss of crystallinity, and for
the former in-situ single-crystal diraction revealed the
complete structure of the fully dehydrated compound, and
its reversibility to the original state after exposure to air.
232
In
contrast, [Co
7
(OH)
12
(H
2
O)
2
](C
2
H
4
(SO
3
)
2
) has coordinated
water while the disulfonate sits in the gallery and is hydrogen
bonded to the layers. It is worth noting that the ratio of
tetrahedral to octahedral units in the structure may be related
to the length of the sulfonate.
The magnetic properties of these layers are quite varied and
the reason for the dierence is logical. First, Co(OH)
2
and
Co
2
(OH)
3
NO
3
are metamagnets with Ne el transitions of 10 1 K
and critical eld at 2 K of less than 1500 Oe.
227229,236,237
The
interpretation is that the cobalt atoms within the layers are
ferromagnetically coupled and the layers are then anti-
ferromagnetically coupled. A weak critical eld reverses the
moments to the ferromagnetic state as judged by the satura-
tion magnetization approaching 5 m
B
. On the other hand all
the studies, magnetization at ambient and elevated pressures,
heat capacity and neutron scattering, indicate that
Co
5
(OH)
6
(SO
4
)
2
(H
2
O)
4
has the characteristics of a pure ferro-
magnet at all elds and the Curie temperature is 14 K.
230,231
Neutron diraction analysis suggests that the moments of the
cobalt atoms are in the plane of the layers (Fig. 19 (I)). Further
analysis conrms the xy nature of the magnetic system.
Fig. 19 Structures of structurally characterised layered cobalt
hydroxides.
Fig. 18 Temperature dependence of the magnetic susceptibility of
Na
2
Co
2
(ox)
3
(H
2
O)
2
(red) and Na
2
Ni
2
(ox)
3
(H
2
O)
2
(blue).
1368 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
Measurements on aligned single-crystals of Co
5
(OH)
6
(SO
4
)
2
(H
2
O)
4
point to an easy-plane magnet. However, the entropy sum of
the l-anomalies in the heat capacity data accounts for
only four of the cobalt moments at the long-range magnetic
ordering temperature at 14 K and the other one, presumably
that corresponding to the cobalt in the gallery, is ordered at
lower temperatures. This interesting observation, conrmed
by the neutron data, strongly suggests that these layers are
very weakly coupled and thus can be regarded as single-layer
magnets. Magnetization measurements under pressure show a
linear increase of T
C
of 1% kbar
1
.
238
The coercive eld is
rather low. In contrast, Co
5
(OH)
8
(trans-1,4-chdc)4H
2
O
(Fig. 20) and [Co
7
(OH)
12
(H
2
O)
2
](C
2
H
4
(SO
3
)
2
) are ferri-
magnets with T
C
of 61 and 33 K, respectively. The saturation
magnetization is the dierence between the moments on the
octahedral and the tetrahedral sites. In these two compounds
the ferrimagnetism implies that the coupling between layers is
ferromagnetic.
Measurements on aligned single-crystals of Co
5
(OH)
8
(trans-
1,4-chdc)4H
2
O also point to an easy-plane magnet as for
Co
5
(OH)
6
(SO
4
)
2
(H
2
O)
4
. In contrast to the latter, Co
5
(OH)
8
-
(trans-1,4-chdc)4H
2
O has a high coercive eld reaching
22 kOe at 2 K (Fig. 20).
232
The Curie temperature was found
to also increase at a rate of 1% kbar
1
.
238
Dehydration and
rehydration of Co
5
(OH)
8
(trans-1,4-chdc)4H
2
O do not
inuence the magnetic properties (T
C
, Weiss constant and
coercive eld) even though the interlayer is reduced upon
dehydration. All the results point to single-layer behaviour. The
amine bridged compounds, Co
8
(OH)
12
(SO
4
)
2
(diamine)nH
2
O,
behave dierently (Fig. 21).
235
Both are antiferromagnets with
T
N
of 14 K (en) and 21 K (DABCO). A small canting is
present for the DABCO compound. Both show metamagnetism
with critical eld scaling with the number of exchange
pathways (one for en and three for DABCO) between the
layers, arming the interaction through bonds rather than
through space. The reason for the dierent ground-state
remains unclear. Easy-plane magnetism is suggested by
anisotropy in the magnetization of an aligned-crystals sample.
Due to the absence of structures with dierent ratio of
tetrahedral and octahedral sites (N(tet)/N(oct)), and based
on the independence of the Curie temperatures on the inter-
layer distances for a series of Co
5
(OH)
8
LnH
2
O where L is a
dicarboxylate, sulfonate or dicyanamide, I had proposed that
the long-range magnetic ordering is induced by dipolar inter-
action between layers of highly-correlated moments.
239241
In
view of a clear dependence of T
C
on the newly reported
structures with dierent N(tet)/N(oct), I have re-examined this
point (Fig. 22). The linear dependence of T
C
vs. N(tet)/N(oct),
if one assume T
C
is a direct measure of the overall exchange
interaction and taking account of the sign convention (positive
for ferromagnetic and negative for antiferromagnetic), is clear
indication that the Curie temperature depends on the number
of exchange pathways between tetrahedral and octahedral
Fig. 20 Temperature dependence of the ac and dc (ZFCFC)
susceptibilities in an applied eld of 1 Oe (top) and the isothermal
magnetization at 2 K (bottom) for Co
5
(OH)
8
(trans-1,4-chdc)4H
2
O.
Fig. 21 Isothermal magnetization at 2 K for eld parallel and
perpendicular to the layers of Co
8
(OH)
12
(SO
4
)
2
(DABCO)H
2
O (top)
and the temperature dependence of the critical eld (bottom) of
Co
8
(OH)
12
(SO
4
)
2
(diamine)nH
2
O on polycrystalline samples including
the critical exponent ts (solid lines).
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1369
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
sites since these are the dominant interactions. This result
suggests that an increase of T
C
would require replacement of
more octahedra by tetrahedra, which is not likely due to the
loss of connectivity within the brucite layer.
K[Co
14
O(OH)
21
(ox)
3
] (Fig. 19 (V)) is a new member in this
series.
242
It is a very interesting system whereby there are
two alternating layers, one of Co-hydroxide and the other
Co-oxalate. The two are matched and they share the oxygen
atoms of oxalate at their interfaces. It behaves as a ferrimagnet
with a T
C
of 33 K. Assuming that the magnetism is derived
principally from the hydroxide layers, the Curie temperature
ts well to the above correlation (see blue point of gure). The
value of the saturation magnetization at 2 K indicates that
the moments of the octahedral cobalt atoms (in both hydroxide
and oxalate layers) are aligned in one direction and those of
the tetrahedral cobalt atoms point in the opposite direction.
Francois et al. have recently reported a non-brucite
cobalt hydroxide layered compound, Co
5
(OH)
6
(C
8
H
12
O
4
)
2
(Fig. 19 (VI)),
243
where all the cobalt atoms adopt octahedral
geometry and there are two dicarboxylate per ve cobalts as
opposed to only one in Co
5
(OH)
8
(trans-1,4-chdc)4H
2
O. The
structure has an unusual arrangement of the octahedra where
both edge- and vertex-sharing are observed with one very
unusual hydroxide which is coordinated to four cobalts in a
nearly square-planar fashion. It is found to be a ferrimagnet
(T
C
= 16.3 K) with a saturation magnetization of 4.3 m
B
per
formula unit and does not sum to an integral number of cobalt
moments. The nickel analogue is also reported and it is an
antiferromagnet at 26.5 K.
Going down in the hierarchy in the MOM connection, the
series containing the rutile [110] layers is the next in line.
244246
This particular layer consists of all octahedrally coordinated
cobalt atoms where both edge- and vertex-sharing are present,
resulting in an expected alternate ferromagnetic chains
coupled antiferromagnetically. The orthogonal competing
interactions in this system have some striking consequences.
This structure is known from the Rietveld renement of
PXRD for Co
2
(OH)
2
(1,4-bdc) (Fig. 19 (VII)). Although
similar PXRDpatterns have been observed for related compounds
containing 4,4
0
-biphenyldicarboxylate, carboxycinnamate,
2,9-naphthalenedicarboxylate, the quality of the XRPD does
not permit full structural analyses. However, their basic
magnetic properties are very similar. The key magnetic proper-
ties are those of metamagnets, but with a particular behaviour
where on cooling a sample of Co
2
(OH)
2
(1,4-bdc) in a small
eld, a collinear antiferromagnetic ordering at 48 K is
observed followed by a very weak spontaneous magnetization
at 43 K (Fig. 23). On increasing the strength of the eld a
gradual reversal of the moments takes place which does not
reach saturation to a ferromagnetic state. The isothermal
magnetization measured at dierent temperatures reveals
paramagnetic behaviour above 48 K, an s-shape consistent
with a metamagnet without hysteresis for 43 4 T 4 48 K, a
constricted loop with some coercivity between 20 4 T 4 43 K
and below 20 K the coercive eld varies very rapidly to the
point that 50 kOe is not enough to reverse the moments.
This is amongst the highest coercive magnets known. The
saturation magnetization is never reached even in 330 kOe. All
the compounds exhibit similar characteristics, two ordering
transitions and fractional moments. Neither the transition
temperature nor the width of the loop at 2 K has any direct
relation with the interlayer distances. In contrast, the
metamagnetic critical eld in the temperature region between
the two transitions appears to relate to the interlayer distance,
lowering as the distance increases. There may be other variants
of cobalt hydroxide terephthalate which have dierent
transition temperatures (38 and 22.5 K) and similar wide
hysteresis loops.
247
Fig. 22 Correlation between the Curie temperature and the ratio of
tetrahedral to octahedral cobalt sites in the layer.
Fig. 23 Field cooling ac and dc susceptibilities (top) for
Co
2
(OH)
2
(1,4-bdc) and isothermal magnetization (bottom) at
temperatures in three dierent regions of the phase diagram.
1370 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
The highly distorted kagome layered compound, Co
3
(m
3
-OH)
2
-
(trans-1,2-chdc)
2
(Fig. 24 (VIII)), where chdc = cyclohexane-
dicarboxylate, contains edge-sharing trimers of cobalt, in three
dierent coordination geometries, connected at their vertices
and bridged by the carboxylate groups.
248
The layers are
separated by the cyclohexane backbone. Due to the asymmetry
of the triangles, where two of the three magnetic pathways
between cobalt are expected to be ferromagnetic and one
antiferromagnetic, the expected frustration of a regular
kagome compound is not present and a long-range antiferro-
magnetic order is instead observed at 11 K. A small canting of
the moments accompanies the antiferromagnetic transition.
Interestingly, the two kagome (Fig. 24 (IX)) layered
compounds pillared by bpy, Co
3
(idc)
2
(4,4
0
-bpy)(H
2
O)
4
2H
2
O
and Co
3
(idc)
2
(4,4
0
-bpy)
3
(6H
2
Odmf), idc = imidazole-
dicarboxylate, reported by Wang et al., having three-atom
bridges (OCO and NCN) were found to be weakly interacting.
249
For a sulfur-bridged Co
2
O
2
edge-sharing dimers within the
layers of Co(1,2-thiosalicylate) (Fig. 24 (X)), Wood et al.
reported a canted antiferromagnetic order at 9 K.
60
Using
limited susceptibility data in the range 100300 K, an estimate
by the CurieWeiss law gave an amazingly large Weiss
constant of 585 K which has been associated with the greater
covalency of the CoS bond and greater delocalisation of
spin-density. Using azide and isonicotinic acid, Liu et al.
obtained the layered compound Co
3
(N
3
)
2
(OH)
2
(isonicotinate)
2
which contains chains of edge-sharing CoO connected by
1,3-azido.
250
It behaves as a paramagnet. Other edge- and
vertex-sharing cobalt antiferromagnetic layered compounds
such as Co
2
(hypa)
2
(4,4
0
-bpy) (Fig. 24 (XI)), hypa = hydroxy-
phenylacetate (T
N
= 15.2 K),
251
are known and for the series
Co
2
(ma)(ina) (Fig. 24 (XII)) and its solvated forms Co
2
(ma)(ina)S,
where S = 2H
2
O, MeOH or HCONH
2
, ma = malate,
ina = isonicotinate) T
N
are o2, 3.5, 3.5 and 8 K,
respectively.
252
Cobalt glycerolate (Fig. 24 (XIII)) is isostructural to its zinc
analogue which is used in most cosmetic products because zinc
repairs the skin and it is also a very good UV blocker.
253
Most
importantly, it has a lamellar structure where upon rubbing it
is delaminated very easily and smoothly like graphite. This is
because the layer surfaces are all hydrogen atoms and
consequently are weakly held together. We were interested
by the presence of its CoOCo connection within the pseudo-
hexagonal layer geometry. Each cobalt is distorted
ve-coordinate and vertex-sharing with its three neighbours.
An earlier magnetic measurement reported a canted anti-
ferromagnetic ordering.
254
We re-studied its magnetic properties,
and extended them by heat capacity, neutron diraction
and muon spin rotation measurements. Its magnetic behaviour
is exactly analogous to that of Co
2
(OH)
2
(1,4-bdc) and
Co
2
(pm) but with transition temperatures of 36 K (collinear)
and 35 K (canting) and with a coercive eld at 2 K of
20 kOe.
255
The metamagnetic critical eld of a mere 160 Oe
may be related to the absence of bonded bridges between the layers.
Furthermore, the pressure dependence of the susceptibility
shows a similar change of the transition temperatures of
ca. 1% kbar
1
. The neutron magnetic structure reveals that
the moments are antiferromagnetically coupled and lie within
Fig. 24 Structures of layered compounds with one-atom and three-atom bridges.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1371
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
the layer with a small out-of-plane resultant moment which is
compensated by the opposite alignment of those in adjacent
layers. From the muon results all but two of the principal
critical exponents were evaluated and the two remaining ones
were obtained by a scaling sum.
255,256
They match pretty well
those of the N = 2 chiral universality class.
257
There are several cobalt hydroxy-succinate layered
compounds with dierent arrangements of both edge- and
vertex-sharing polyhedra and most of them are reported to be
antiferromagnetic below 10 K.
258261
The controlled synthesis,
structures and properties of these compounds have been
reviewed on several occasions.
A number of square-grid layered compounds, Co(H
2
O)
2
(L)
(Fig. 24 (XIV)) where L is a carboxylate from tere-
phthalate, acetylenedicarboxyolate, 4-carboxycinnamate,
1/2 pyromellitate, 2 trans-cinnamate or malonate has
been characterised.
63,65,244,262267
In each case, the cobalt
atoms are connected by OCO bridges in the plane and the
two water molecules are in trans-positions. The layers can be
either pillared when the two carboxylate groups are from the
same acid or separated when a monoacid is used. When phenyl-
glycinate and mandalate are the coordinating ligands, the
amine and the alcohol group of the a-carbon replace the water
molecules. Similar layered structures are obtained with
pimelate or glutarate but these are not hydrated and thus
the cobalt is in tetrahedral coordination.
268,269
The rst series
of octahedral cobalt species were found to be paramagnetic
with classical behaviour which has been tted using the
MabbsMachin model.
74
However, the heat capacity of the
pyromellitate compound shows a clear l-type anomaly
characteristic of a long-range magnetic ordering at 0.9 K.
270
In the absence of further magnetization measurements at these
temperatures, the ground state has not been identied. The
nickel analogue is a metamagnet with a Ne el temperature of
3 K and a critical eld of 1300 Oe at 2 K.
271
For the
tetrahedral containing compounds a maximum was observed
in the temperature dependence of the susceptibility, which is
interpreted by Fe rey et al. as due to canted antiferromagnetic
ordering at 20 K for the pimelate, but Lee et al. used the Lines
model for a square-grid of isotropic S = 3/2 to t the data of
the glutarate very well for the whole temperature range.
272
The magnetic exchange between cobalt is weakly antiferro-
magnetic. Rettig et al.
273
found the related carboxylate-
bridged layered compound Co(formamide)
2
(HCOO)
2
is the
exception within this series. The key feature of the structure is
that the layers are separated by the formamide in an inter-
digitated fashion. It behaves as a canted antiferromagnet
below 9 K, a feature similar to the 3D-perovskites
A[Co(HCOO)
3
] where A is an alkylammonium cation (see
later). The high-temperature data can be tted to the square-
lattice model of Lines,
272
and the estimated canting angle from
the remanent magnetization and an expected saturation value
of 2.4 m
B
is 1.51.
Among the metal oxalate layered compounds the ferromagnetic/
ferrimagnetic ones, (cat
+
)[MM
0
(ox)
3
], composed of two
dierent metallic centres, one divalent (M) and the other
trivalent (M
0
), are the most studied.
274
This is due to the wide
variety of compounds that can be prepared for a variety of
cations with most of the rst-row transition metals, thus
allowing the control of spin states at each site and therefore,
the types of interaction and anisotropy. These studies are well
documented and will not be discussed further. The structure
consists of the honeycomb layers of alternate M and M
0
, with
one stringent requirement that one site is a D-enantiomer and
the other is L.
275277
The layers are therefore non-chiral.
Cobalt can occupy either of the sites. Making purely single
valent compounds required either no cation if the metals are
trivalent or a dication if the metals are divalent. To our
knowledge, the latter exists only for cobalt,
(bpyph
2+
)[Co
II
2
(ox)
3
]5.5H
2
O (Fig. 24 (XV)), where bpyph is
1,4-bis(4-pyridyl-1-pyridinio)phthalazinium.
278
The neighbouring
cobalt atoms are antiferromagnetically coupled (Y =38 K).
There is evidence of low-dimensional behaviour but no LRO
was observed above 2 K. Colacio et al. replaced the oxalate ion
by pyrimidine-2-carboxylato (pymca) which they obtained
in situ by hydrothermal reaction of 2-cyanopyrimidine and
either CoCl
2
6H
2
O or FeCl
2
4H
2
O to crystallise the 2D
hexagonal coordination polymers [M
2
(m-pymca)
3
]OHH
2
O.
279
For M = Co spin-canted antiferromagnetism is observed
below 10 K whereas for M = Fe a normal antiferromagnet
at T
N
of 21 K results. Slight modications to the MM
0
layered
compounds by introducing crown ether and a small cation
(such as K, Na) in the galleries has been reported by Coronado
et al.
280,281
For example, the series [K(18-crown-6)]
3
[M
II
3
(H
2
O)
4
-
{M
III
(ox)
3
}
3
] (M
III
= Cr, Fe; M
II
= Mn, Fe, Ni, Co, Cu;
ox = C
2
O
4
2
) contains some disruption of the oxalate M
II
bonds
and the Cr
III
M
II
derivatives behave as ferromagnets with
critical temperatures up to 8 K while the Fe
III
M
II
present
ferrimagnetic or weak ferromagnetic ordering up to 26 K.
Major modication was found in the neutral compound
{[Co(H
2
O)
2
]
3
[Cr(ox)
3
]
2
(18-crown-6)
2
} where the crown ether
is involved in the layer structure again with disruption of the
oxalate cobalt bonds. The resulting compound, now soluble
and recrystallisable as opposed to the MM
0
honeycomb
compounds, is a ferromagnet at 7.4 K.
Three more known layered oxalates are known. The rst,
Co(H
2
O)(ox) (Fig. 24 (XVI)), consists of octahedral coordi-
nated cobalt centres with both CoOCo and CoOCOCo
connections.
282
Magnetic susceptibility shows a low dimensional
behaviour with Weiss constant of 42 K and a long-range
ordering at 21.5 K to a canted antiferromagnetic ground state.
A similar structure and magnetic properties were reported for
Fe(CH
3
OH)(ox) with T
C
of 23 K.
283
The second is not
known on its own and it is part of K[Co
14
O(OH)
21
(ox)
3
]
(Fig. 24 (XVII)) reported by Tong et al.
242
The oxalate layer
is sandwiched between CoOH layers as discussed above. The
combined layers order as a ferrimagnet and the results suggest
that the oxalate moiety is ferromagnetically coupled. The third
exists for three transition metals, Ba
2
M
2
(ox)
3
Cl
2
(H
2
O)
4
(M = Mn, Fe or Co (Fig. 24 (XVIII)) and the metal atoms
form a structure analogous to black phosphorus where they
are the corners of a chair cyclohexane connected in opposite
orientation to produce a corrugated sheet.
284
The cobalt atoms
are bridged by OCO and are antiferromagnetically coupled.
Antiferromagnetic LRO may be present at 11 K. Two other
layered compounds composed of edge- and vertex-sharing
between octahedral cobalt centres are known.
285,286
The layers
in Co
2
(L-Asp)
2
(4,4
0
-bpy)1.5H
2
O, Asp = aspartate, is bridged
1372 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
by 4,4
0
-bpy to give a homochiral 3D network with pores that
can absorb hydrogen gas and is an antiferromagnet with a T
N
of 15 K. Co
2
(tzc)
2
(bpe) (Fig. 24 (XIX)), tzc = tetrazole-5-
carboxylate, bpe = 1,2-bis(4-pyridyl)ethane, forms a Cotzc
layer separated by bpe.
286
Within the layer both two-atom and
three-atom bridges are present. It orders as an antiferromagnet
with a slight canting at 9 K. Metamagnetism is also observed
with a critical eld of 500 Oe. The 4,4
0
-bpy pillared layered
compounds, Co(HCOO)
2
(4,4
0
-bpy)nH
2
O are canted antiferro-
magnets with T
N
= 7.4 K (n = 0) and 3 K (n = 5).
287
7. 3D-networks
Compounds containing an all MOM framework in MOFs
are not commonly found due to the complexities of the organic
ligands used and consequent structural arrangements. Few are
known, here, we will show several compounds for which some
magnetic properties have been described.
The rst is Co
5
(OH)
2
(OAc)
8
2H
2
O (Fig. 25 (I)) which was
obtained in high yield by solvothermal treatment of Co(acac)
3
in a THF solution containing 4 vol% H
2
O.
288
All the cobalt
atoms are octahedrally coordinated by oxygen. Interestingly it
consists of four-fold (I4
1
/a) helical chains of trimeric units,
Co
3
(OH), bridged by another cobalt. Unfortunately, only a
brief study of the magnetism is reported showing paramagnetism
at high temperatures and possibly a canted antiferromagnetic
behaviour below 30 K.
The second one is the isostructural series of porous
M
3
(HCOO)
6
S (Fig. 25 (II)) where M can be paramagnetic
(Fe, Mn, Co and Ni) or diamagnetic (Zn and Mg), S can be
one of over forty solvents.
289299
The parent compound,
M
3
(HCOO)
6
(H
2
O)CH
3
OH), can be desolvated by thermal
treatment or under reduced pressure without damage to the
framework. The resulting pores can be relled in the solidgas
or solidliquid media by dierent solvents. Gases such as N
2
,
H
2
and CH
4
have also been inserted in the desolvated
compounds at low temperatures. The structure shows an
interesting arrangement of the octahedra in MM
4
tetrahedral
nodes which then form a distorted diamond structure, which
we believe leads to the stability of the framework. The metals
are bridged by both one single-oxygen and one three-atom
carboxylate bridge. Each formate unit contributes to three
coordination bonds, thus there are no vacant sites for coordi-
nation of solvent and gases, and the surface of the channels is
lined with hydrogen and oxygen atoms providing both hydro-
philic and hydrophobic characteristics. The consequence of
this is the wide range of molecules that can be accommodated
in the channels. From a magnetic point of view, we anticipated
that the continuous MOM through the network should
result in magnetic ordering at temperatures above 50 K, but
this was not the case. The observed types of magnetic ground-
state and transition temperatures, which is dependent on
solvent, were for Fe (ferromagnetic, 1522 K), Mn (ferrimag-
netic, 58 K), Co (unknown type, o2 K) and Ni (ferrimag-
netic, 2.7 K). Due to the slight exibility of the OCO bridges,
the volume of the framework can be stretched up to 12% by
the insertion of solvents, and continuously by 11% by
changing the metal cations or by tuning the proportion of
metal cations in solid solutions. This capability of continuous
tuning of the volume of the frameworks, thus the channel
sizes, is very appealing for selective sorption. Interestingly, we
have demonstrated the development from ferromagnetism to
spin-glass behaviour by doping the Fe-compound with a
non-magnetic cation such as Zn. Taking advantage of the
occurrence of the chirality of the channels, 50% M and 50%
P, the insertion of a chiral alcohol (2-chloropropan-1-ol) into
the channels results in well ordered guests when the chirality of
the guests matches that of the channel and disorder is present
when this match is absent.
298
The consequence is that the non-
chiral (P2
1
/c) framework is rendered chiral (P2
1
). Therefore,
the magnetism can be ne-tuned by the choice of solvents and
mixture of cations while the ground state can be controlled by
the metal electronic and spin states and furthermore, chirality
can also be controlled by the chirality of the solvents. Solvent
dependent dielectricity has been reported for the manganese
formates.
300,301
In some way, there is some analogy to the
Prussian Blue family where the type and temperature of the
magnetic transitions can be controlled and solvent can also
modify the intrinsic properties. As yet there is no control of
chirality for this type of materials.
Fig. 25 Structures of some 3D-networks.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1373
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
The third compound where all the cobalt atoms are
connected via one oxygen atom is Co(OH)(HCOO)
(Fig. 25 (III)).
302,303
It consists of a framework based on a
hexaprismane MOP, Co
6
(OH)
6
, like those common in the
chemistry of titanium. Each hydroxide is m
3
-bridging and each
formate provides three coordination bonds to satisfy the
octahedral coordination of the cobalt atom. Within the
hexaprismane the cobalt octahedra are edge-sharing to each
other. Each hexaprismane is connected to six of its neighbours
by edge sharing. The overall structure is highly symmetric.
Consequently, each cobalt atom has three cobalt neighbours
and therefore, the spins are highly frustrated.
304308
The
magnetic properties show no long-range ordering though the
exchange interaction is quite strong and antiferromagnetic
(Y = 105 K).
The ambidentate nature of the cyanide ligand is known to
provide linear two-atom connections between most of the
metals in the periodic table and its triple bond (CRN) results
in strong magnetic exchange interaction between nearest
neighbours. In contrast, the carboxylate in MOFs is at a
disadvantage, providing three-atom bridges and weak inter-
actions due to the 1.5 bond order. There are a few compounds
known where all the bridges are OCO. The rst is a family of
metal formates having the perovskite ABX
3
structures
(Fig. 25 (IV)) where the cation A is an organic ammonium,
the metal B is a divalent metal (paramagnetic Mn, Co or Ni or
non-magnetic Zn and Mg) and X is the formate acting as
bis-monodentate bridges.
309313
The only requirement for
obtaining this particular phase is that the ammonium cation
has to be of a size to t the space with the cube. There are
several variants of the structures adopting dierent space
groups depending on the shape, size and charge of the cations.
All the compounds studied so far exhibited canted-antiferro-
magnetism where the estimated canting angles are rather
small. The transition temperatures are in the order
Mn o Co o Ni as the order of the cationic radii which is
surprisingly in the opposite sense of the spin quantum number
as one may expect within mean-eld-theory. For cobalt
compounds, the dependence of the Ne el temperature on the
organic cation is dimethylammonium (15 K), ammonium
(9 K) and di(N-methyl)ethanediammonium (16.7 K). The
perovskite contains the square-grid layers which have been
discussed above and in Co(H
2
O)
2
(HCOO)
2
(Fig. 25 (V)) the
same layers are pillared by a Co(H
2
O)
4
units in a similar
fashion as found in Co
5
(OH)
6
(SO
4
)
2
(H
2
O)
4
.
314,315
Therefore,
it might be more appropriate to write the formulation as
[Co(HCO
2
)
4
{Co(H
2
O)
4
)}]. The magnetic properties of this
compound have fascinated physicists for a long time as there
is a transition to a long-range magnetic order at 6 K but the
heat-capacity data suggest a second one in the mK region.
316
The interpretation is that the layers are fully ordered at 6 K
while the moment of the Co(H
2
O)
4
unit is still uctuating and
it comes into full order at the mK transition. There is very
strong orientation dependence in the magnetic susceptibility
suggesting a possible 2D Ising-type anisotropy.
Interestingly, Cheetham et al. has identied a dielectric
change associated with a subtle phase transition for
((CH
3
)
2
NH
2
)[Zn(HCOO)
3
].
313
If this transition is present for
the cobalt compound, it does not appear to have any eect on
the magnetism as is also the case for the porous Mn
3
(HCOO)
6
S
compounds.
300,301
Related to this series of three-dimensional OCO bridged
cobalt atoms, Tong et al. obtained Co
3
(L)(H
2
O)
6
(Fig. 25 (VI))
which is formed by a conformational transformation of
cis,cis,cis,cis,cis-1,2,3,4,5,6-cyclohexanehexacarboxylic acid (H
6
L)
to the trans,trans,trans,trans,trans-isomer during the hydro-
thermal reaction with CoCl
2
6H
2
O.
317
It crystallises in a
unique high-symmetry (R

3) three-dimensional metalorganic
framework. Each ligand connects nine Co(II) atoms, and each
octahedral Co(II) atom is connected to four carboxylate
groups and two water molecules in trans-positions. The
structure is a variant of quartz where the OCO group replaces
the oxygen atoms in SiO
2
where the CoO
4
deviates from the
perfect tetrahedron of SiO
4
. This is in contrast to the
perovskite compounds where each Co atom has six neigh-
bours. It behaves as a paramagnet (C = 2.99 cm
3
K mol
1
Co,
Y =28 K) above 3 K where it orders as an antiferromagnet
with a metamagnetic critical eld of ca. 10 kOe.
As for chains where both one-atom and three-atom bridges
are involved in their formation, similar situations are found
for 3D-networks. One that uses both, reported by Powell et al.,
is Co
2
(OH)
2
(ox) (Fig. 25 (VII)). It consists of edge-sharing
chains which are vertex-sharing to form layers, and the latter
are then pillared by the tetracoordinated oxalate to give a very
stable 3D-network. The high-temperature magnetic data
suggest dominant antiferromagnetic coupling with a Weiss
constant of 105 K.
318
Low-eld ZFCFC dc-magnetization
and ac-susceptibilities measurements reveal a most unusual
behaviour below 100 K. First, a long-range collinear anti-
ferromagnetic ordering is observed at 82 K, the highest
ordering temperature in frameworks of cobalt, and then a
canting of the moments is evidenced at 41 K (Fig. 26). By
increasing the applied eld the spontaneous magnetization due
to the canting is obscured by the saturation of the moment.
The isothermal magnetizations at dierent temperatures are
linear with eld up to 150 kOe. The iron analogue has a higher
transition temperature for the collinear ordering but a lower
one for the canting.
319
While the honeycomb metaloxalate layers are formed when
a non-chiral cation is employed in the synthesis,
320
a 3D-chiral
structure results when chiral cations are used.
321
The 3D-net is
induced by that imposed by the handedness of the cation.
When a tris(bidentate ligand)metal cation such as M(bpy)
3
is
used, complete segregation of the enantiomers occurred. A
series of 3D-chiral networks have been therefore prepared and
their magnetic properties studied. Regarding the kind of
interaction between nearest neighbours, similar rules as those
found for the hexagonal layers were established. When the two
metals within the 3D-network are cobalt(II), and the cation is
Co
III
(bpy)
3
3+
an additional anion is needed in the solid as
found in [Co
III
(bpy)
3
][Co
II
2
(ox)
3
]ClO
4
. When the cation is
M
II
(bpy)
3
2+
no counter-anion is required as for [M
II
(bpy)
3
]-
[Co
II
2
(ox)
3
], M
II
= Fe or Ni. For the cobalt containing
3D-network compounds the magnetic properties are those of
canted-antiferromagnets with T
C
of less than 10 K.
322
1,2,3-Triazole-4,5-dicarboxylate (tzdc) has three chelating
ends at nearly 1201 to each other and when they are coordi-
nated to divalent metals adopting octahedral geometry
1374 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
automatically chiral selection takes place and the crystals are
all enantiomer-pure as found for (H
2
NMe
2
)[M(tzdc)]0.5H
2
O,
M = Co or Mn, and (NH
4
)[Mn(tzdc)]2.6H
2
O.
323
Similar to
the observation made for the perovskite metal formate the size
of the cations play an important role on the crystal packing
and thus the space group. The methylamine derivatives adopt
the cubic space group P2
1
3 and a porous (10,3)-a network
while the NH
4
+
-templated compound adopts a trigonal space
group P6
1
a and a uniform etd (8,3) network. Due to the
three-/four-atom bridges between the cobalt atoms a very
low temperature magnetic transition is observed below 2.5 K.
The only 3D-network where every pair of cobalt atoms is
connected by a four-atom bridge (OCCO) is Co(H
2
O)
2
(C
4
O
4
).
105
The cobalt atoms in this cubic phase (a = 16.244 A

) are all
crystallographically equivalent and have two water molecules
in the trans positions and four equatorial oxygen atoms from
four squarate units. The two dierent closest CoCo distances
are 5.71 and 5.78 A

. The temperature dependence of the


magnetic susceptibility can be tted to that of non-interacting
cobalt for all temperatures above 2 K and the isothermal
magnetization shows no anomaly.
8. Conclusion
The success in the synthesis and characterization of framework
compounds over the last ten years has been phenomenal. The
large number of compounds witnesses the progress made so
far and the increase in literature reports is exponential. The
structures are diverse, providing examples of all types for
structureproperty relationship, though in most cases the
design is uncontrollable by the synthetic approach. The
magnetic properties have generally been well explored and
they illustrate the potential of producing magnets of many
dierent long-range ordered ground states, ferromagnet,
ferrimagnet, antiferromagnet, canted-antiferromagnet and
most unusual metamagnets, in addition to short-range ordered
states, single-molecule magnets and single-chain magnets. In
the present review, we have attempted to build up our under-
standing of the ultimate magnetic behaviour by progressing
through the dierent oligomers to chains, layers and
3D-network to provide a solid ground for future elaborations
with particular aims in mind. It is clear that metalorganic
systems will continue to provide new properties and improve
existing knowledge of their structures as well as their proper-
ties. We nally hope the day will come when enough informa-
tion will be available to permit designing magnetic materials
for specic purposes. Using the porosity of the frameworks,
several groups have successfully tuned the magnetic properties
between the dierent ground states listed above by the solvent
content. Introducing bifunctionality using spin-crossover,
electrical conductivity, dielectricity, mixed-valency and optical
eects are very promising, and some examples are already
shown. With magnetic transition temperatures above that of
liquid nitrogen and magnetic hardness greater than 60 kOe,
exceeding all known oxide magnets, there appear to be no
limits to the avenues for development and therefore are
encouraging for our community.
Acknowledgements
I am indebted to H. Kumagai, S. Vilminot, A. Jouaiti, F. L. Pratt,
S. J. Blundell, S. Takagi, C. J. Kepert, Z.-M. Wang, M. A. Green
and K. Inoue for their contributions, as well as to the funding
agencies, CNRS-France, EPSRC-UK, ESF-Europe and
JSPS-Japan. I thank the Cambridge Crystallographic Data Centre
(UK) for providing many of the crystal data. I am grateful to the
Royal Society of Chemistry (UK) for a travel grant.
References
1 MetalOrganic and Organic Molecular Magnets, , Spec. Publ. - R.
Soc. Chem., ed. P. Day and A. E. Underhill, Cambridge, UK, 2000,
vol. 252.
2 C. N. R. Rao, A. K. Cheetham and A. Thirumurugan, J. Phys.:
Condens. Matter, 2008, 20, 083202.
3 S. J. Blundell and F. L. Pratt, J. Phys.: Condens. Matter, 2004, 16,
R771R828.
4 Molecular Magnetism, New Magnetic Materials, ed. K. Itoh and
M. Kinoshita, Gordon Breach-Kodansha, Tokyo, 2000.
5 S. J. Blundell, Magnetism in Condensed Matter, Oxford
University Press, 2001.
6 O. M. Yaghi, H. L. Li, C. Davis, D. Richardson and T. L. Groy,
Acc. Chem. Res., 1998, 31, 474.
7 S. R. Batten and R. Robson, Angew. Chem., 1998, 110, 1558.
8 O. M. Yaghi, Acc. Chem. Res., 2005, 38, 176.
9 M. Eddaoudi, D. B. Moler, H. L. Li, B. L. Chen, T. M. Reineke,
M. OKeee and O. M. Yaghi, Acc. Chem. Res., 2001, 34,
319330.
10 D. J. Tranchemontagne, Z. Ni, M. OKeee and O. M. Yaghi,
Angew. Chem., Int. Ed., 2008, 47, 5136.
11 M. J. Rosseinsky, Microporous Mesoporous Mater., 2004, 73, 15.
12 R. Robson, J. Chem. Soc., Dalton Trans., 2000, 3735.
13 S. Kitagawa, R. Kitaura and S. Noro, Angew. Chem., Int. Ed.,
2004, 43, 2334.
14 M. M. Schieber, Selected Topics in Solid State Physics Experi-
mental Magnetochemistry, ed. E. P. Wohlfarth, North-Holland,
Amsterdam, 1967, vol. VIII.
15 Physics meets Mineralogy, ed. H. Aoki, Y. Syono and
R. J. Hemley, CUP, Cambridge, 2000.
16 Magnetism: A Supramolecular Function, ed. O. Kahn, Kluwer
Academic Publishers, 1996.
17 M. OKeee and B. G. Hyde, Crystal Structures I: Patterns and
Symmetry, Mineralogical Association of America, Washington,
DC, 1996.
18 M. OKeee, M. Eddaoudi, H. Li, T. Reineke and O. M. Yaghi,
J. Solid State Chem., 2000, 152, 320.
19 H. Li, M. Eddaoudi, M. OKeee and O. M. Yaghi, Nature, 1999,
402, 276.
Fig. 26 Temperature dependence of the FC susceptibility of
Co
2
(OH)
2
(ox) in dierent applied magnetic elds (100 Oe, 5 kOe
and 150 kOe), and ZFCFC in 1 Oe.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1375
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
20 M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter,
M. OKeee and O. M. Yaghi, Science, 2002, 295, 469.
21 M. OKeee, M. A. Peskov, S. J. Ramsden and O. M. Yaghi, Acc.
Chem. Res., 2008, 41, 1782.
22 A. F. Wells, Further Studies of Three-Dimensional Nets, American
Crystallographic Association, Monograph 8, Mineralogical
Society of America, Washington, 1979.
23 S. Han and J. V. Smith, Acta Crystallogr., Sect. A, 1999, 55, 33.
24 T. Uemura, S. Horike and S. Kitagawa, Chem.Asian J., 2006, 1,
36.
25 S. Kitagawa and K. Uemura, Chem. Soc. Rev., 2005, 34, 109.
26 G. Fe rey, Chem. Soc. Rev., 2008, 37, 191.
27 D. Bradshaw, J. B. Claridge, E. J. Cussen, T. J. Prior and
M. M. Rosseinsky, Acc. Chem. Res., 2005, 38, 273.
28 D. Nocera, Science, 2007, 315, 789.
29 M. Dinca and J. R. Long, Angew. Chem., Int. Ed., 2008, 47, 2.
30 P. D. C. Dietzel, B. Panella, M. Hirscher, R. Bloma and
H. Fjellva g, Chem. Commun., 2006, 959.
31 X. Zhao, B. Xiao, A. J. Fletcher, K. M. Thomas, D. Bradshaw
and M. J. Rosseinsky, Science, 2004, 306, 1012.
32 B. Chen, N. W. Ockwig, A. R. Millward, D. S. Contreras and
O. M. Yaghi, Angew. Chem., Int. Ed., 2005, 44, 4745.
33 P. Day, J. Chem. Soc., Dalton Trans., 2000, 3483.
34 P. Day, J. Chem. Soc., Dalton Trans., 1997, 701.
35 D. Gatteschi, R. Sessoli and J. Villain, Molecular Nanomagnets,
Oxford University Press, Oxford, 2006.
36 C. Coulon, H. Miyasaka and R. Cle rac, Struct. Bonding, 2006,
122, 163.
37 L. Bogani, A. Vindigni, R. Sessoli and D. Gatteschi, J. Mater.
Chem., 2008, 18, 4750.
38 C. J. Kepert, Chem. Commun., 2006, 695.
39 A. J. Fletcher, K. M. Thomas and M. J. Rosseinsky, J. Solid State
Chem., 2005, 178, 2491.
40 P. J. Hagrman, D. Hagrman and J. Zubieta, Angew. Chem., Int.
Ed., 1999, 38, 2638.
41 B. Moulton and M. J. Zaworotko, Chem. Rev., 2001, 101, 1629.
42 O. R. Evans and W. B. Lin, Acc. Chem. Res., 2002, 35, 511.
43 G. J. Halder and C. J. Kepert, Aust. J. Chem., 2006, 59, 597.
44 J. J. Vittal, Coord. Chem. Rev., 2007, 251, 1781.
45 M. E. Davis, Nature, 2002, 417, 813.
46 M. E. Davis, Top. Catal., 2003, 25, 3.
47 D. W. Lewis, G. Sankar, J. K. Wyles, J. M. Thomas,
C. R. A. Catlow and D. J. Willock, Angew. Chem., Int. Ed. Engl.,
1997, 36, 2675.
48 D. Maspoch, D. Ruiz-Molina and J. Veciana, Chem. Soc. Rev.,
2007, 36, 770.
49 M. OKeefe, M. Eddaoudi, H. Li, T. Reineke and O. M. Yaghi,
J. Solid State Chem., 2000, 152, 3.
50 C. Janiak, Dalton Trans., 2003, 2781.
51 A. K. Cheetham, C. N. R. Rao and R. K. Feller, Chem. Commun.,
2006, 4780.
52 P. M. Forster and A. K. Cheetham, Top. Catal., 2003, 24, 79.
53 S. Singh and H. W. Roesky, Dalton Trans., 2007, 1360.
54 W. Lin, J. Solid State Chem., 2005, 178, 2486.
55 T. D. Hamilton and L. R. MacGillivray, Cryst. Growth Des.,
2004, 4, 419.
56 M. Kurmoo and C. J. Kepert, New J. Chem., 1998, 1515.
57 S. R. Batten and K. S. Murray, Coord. Chem. Rev., 2003, 246,
103.
58 J. S. Grith, The Theory of Transition-Metal Ions, Cambridge
University Press, 1964.
59 A. P. B. Lever, Inorganic Electronic Spectroscopy, Elsevier,
Amsterdam, 1986.
60 D. Cave, J.-M. Gascon, A. D. Bond, S. J. Teat and P. T. Wood,
Chem. Commun., 2002, 1050.
61 A. Distler, D. L. Lohse and S. C. Sevov, J. Chem. Soc., Dalton
Trans., 1999, 1805.
62 L. Pan, N. Ching, X. Huang and J. Li, Inorg. Chem., 2000, 39,
5333.
63 H. Kumagai, C. J. Kepert and M. Kurmoo, Inorg. Chem., 2002,
41, 3410.
64 Y.-H. Liu and M.-T. Ding, Acta Crystallogr., Sect. E, 2007, 63,
m1828.
65 O. Fabelo, J. Pasan, L. Canadillas-Delgado, F. S. Delgado,
F. Lloret, M. Julve and C. Ruiz-Pe rez, Inorg. Chem., 2008, 47, 8053.
66 M. Kurmoo, C. Estournes, Y. Oka, H. Kumagai and K. Inoue,
Inorg. Chem., 2005, 44, 217.
67 C. S. Cundy and P. A. Cox, Chem. Rev., 2003, 103, 663.
68 K. Byrappa and M. Yoshimura, Handbook of Hydrothermal
Technology, Technology for Crystal Growth and Material
Processing, William Andrew Publishing, New York, USA, 2001.
69 R. M. Barrer, Hydrothermal Chemistry of Zeolites, Academic
Press, London, 1982.
70 A. Rabeneau, Angew. Chem., Int. Ed. Engl., 1985, 24, 1026.
71 A. Ramanan and M. S. Whittingham, Cryst. Growth Des., 2006,
6, 2421.
72 P. M. Forster, N. Stock and A. K. Cheetham, Angew. Chem., Int.
Ed., 2005, 44, 7608.
73 P. Mahata, M. Prabu and S. Natarajan, Inorg. Chem., 2008, 47,
8451.
74 F. E. Mabbs and D. J. Machin, Magnetism and Transition Metal
Complexes, Chapman and Hall, London, 1973.
75 O. Fabelo, J. Pasan, F. Lloret, M. Julve and C. Ruiz-Pe rez, Inorg.
Chem., 2008, 47, 3568.
76 A. Abragam and M. L. H. Pryce, Proc. R. Soc. London, Ser. A,
1951, 206, 173.
77 A. Abragam and B. Bleaney, Electron Paramagnetic Resonance
of Transition Metal Ions, OUP, Oxford, 1970, p. 446.
78 J. R. Pilbrow, Transition Ion Electron Paramagnetic Resonance,
Oxford Science Publications, Oxford, 1990, p. 144.
79 R. L. Carlin, Magnetochemistry, Springer-Verlag, Berlin, 1986.
80 Theory of Molecular Paramagnetism, ed. L. N. Mulay and
E. Boudreaux, Wiley, New York, 1976.
81 A. H. Morrish, The Physical Principles of Magnetism,
R. E. Krieger, Huntington, NY, 1980.
82 R. Boca, Coord. Chem. Rev., 2004, 248, 757.
83 R. Becker, M. Prester, H. Berger, P. H. Lin, M. Johnsson,
D. Drobac and I. Zivkovic, J. Solid State Chem., 2007, 180, 1051.
84 M. E. Lines, J. Chem. Phys., 1971, 55, 2977.
85 R. P. van Stapele, J. C. M. Henning, G. E. G. Hardeman and
R. F. Bongers, Phys. Rev., 1966, 150, 310.
86 R. P. van Stapele, H. G. Beljers, P. F. Bongers and H. Zijlstra,
J. Chem. Phys., 1966, 44, 3719.
87 D. Dai, H. Xiang and M.-H. Whangbo, J. Comput. Chem., 2008,
29, 2187.
88 M. E. Foglio, M. C. dos Santos, G. E. Barberis, J. M. Rojo,
J. L. Mesa, L. Lezama and T. Rojo, J. Phys.: Condens. Matter,
2002, 14, 2025.
89 B. R. McGarvey, Can. J. Chem., 1975, 53, 2498.
90 A. S. Wills, Annu. Rep. Prog. Chem., Sect. A: Inorg. Chem., 2004,
100, 509.
91 A. S Wills, Annu. Rep. Prog. Chem., Sect. A: Inorg. Chem., 2005,
101, 472.
92 M. Murrie and D. J. Price, Annu. Rep. Prog. Chem., Sect. A:
Inorg. Chem., 2007, 103, 20.
93 A. Harrison, J. Phys.: Condens. Matter, 2004, 16, S553.
94 S. T. Bramwell, Annu. Rep. Prog. Chem., Sect. A: Inorg. Chem.,
2003, 99, 467.
95 X.-Y. Wang, Z.-M. Wang and S. Gao, Chem. Commun., 2008,
281.
96 S. K. Pati and C. N. R. Rao, Chem. Commun., 2008, 4683.
97 M. Pilkington, M. Gross, P. Franz, M. Biner, S. Decurtins,
H. Stoeckli-Evans and A. Neels, J. Solid State Chem., 2001,
159, 262.
98 N. Guillou, C. Livage and G. Fe rey, Eur. J. Inorg. Chem., 2006,
4963.
99 K. Murray, Eur. J. Inorg. Chem., 2008, 3101.
100 P. Day, Inorg. Chim. Acta, 2008, 361, 3365.
101 L. M. C. Beltran and J. R. Long, Acc. Chem. Res., 2005, 38, 325.
102 J.-P. Zhang and X.-M. Chen, Chem. Commun., 2006, 1689.
103 W. Ouellette, J. R. Gala n-Mascaro s, K. R. Dunbar and
J. Zubieta, Inorg. Chem., 2006, 45, 1909.
104 G. J. Halder, C. J. Kepert, B. Moubaraki, K. S. Murray and
J. D. Cashion, Science, 2002, 298, 1762.
105 H. Kumagai, H. Sobukawa and M. Kurmoo, J. Mater. Sci., 2008,
43, 2123.
106 O. Fabelo, J. Pasan, F. Lloret, M. Julve and C. Ruiz-Pe rez,
CrystEngComm, 2007, 9, 815.
107 Y.-Q. Zheng, J.-L. Lin, W. Hu, H.-Z. Xie, J. Sun and
X.-W. Wang, Inorg. Chem., 2008, 47, 10280.
1376 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
108 Y.-Q. Zheng and H.-Z. Xie, J. Solid State Chem., 2004, 177, 1352.
109 H. Kumagai, K. Inoue and M. Kurmoo, Bull. Chem. Soc. Jpn.,
2002, 75, 1283.
110 H. Kumagai, M. Akita-Tanaka, S. Kawata, K. Inoue,
C. J. Kepert and M. Kurmoo, Cryst. Growth Des., submitted.
111 H. Sakiyama, J. Comput. Chem. Jpn., 2007, 6, 123.
112 S. M. Ostrovsky, K. Falk, J. Pelikan, D. A. Brown,
Z. Tomkowicz and W. Haase, Inorg. Chem., 2006, 45, 688.
113 A. V. Palii, B. S. Tsukerblat, E. Coronado, J. M. Clemente-Juan
and J. J. Borras-Almenar, Inorg. Chem., 2003, 42, 2455.
114 K. Fink, C. Wang and V. Staemmler, Inorg. Chem., 1999, 38,
3847.
115 L.-M. Zheng, X. Wang, Y. Wang and A. J. Jacobson, J. Mater.
Chem., 2001, 11, 11001105.
116 X.-Y. Wang and S. C. Sevov, Inorg. Chem., 2008, 47, 1037.
117 Y. Hu, G. Li, X. Liu, B. Hu, M. Bi, L. Gao, Z. Shi and S. Feng,
CrystEngComm, 2008, 10, 888.
118 Y.-H. Liu, H.-L. Tsai, Y.-L. Lu, Y.-S. Wen, J.-C. Wang and
K.-L. Lu, Inorg. Chem., 2001, 40, 64266431.
119 J. B. Goodenough, Magnetism and the Chemical Bond,
John Wiley and Sons, New York, 1963.
120 J. Kanamori, J. Phys. Chem. Solids, 1959, 10, 87.
121 J. Kanamori, in Magnetism, ed. G. T. Rado and H. Suhl,
Academic Press, New York, 1963, vol. I, ch. 4, p. 127.
122 J. M. Rojo, J. L. Mesa, L. Lezama, J. L. Pizarro, M. I. Arriortua,
J. Rodriguez Fernandez, G. E. Barberis and T. Rojo, Phys. Rev.
B, 2002, 66, 094406.
123 S. Vilminot, G. Andre , F. Boure e-Vigneron, P. J. Baker,
S. J. Blundell and M. Kurmoo, J. Am. Chem. Soc., 2008, 130,
13490.
124 M. A. Green, H. Kumagai and M. Kurmoo, unpublished results.
125 M.-L. Tong, S. Kitagawa, H.-C. Chang and M. Ohba, Chem.
Commun., 2004, 418.
126 M. Kurmoo, A. Jouaiti and N. Kyritsakas, in preparation.
127 N. Benbellat, K. S. Gavrilenko, Y. Le Gal, O. Cador, S. Golhen,
A. Gouasmia, J.-M. Fabre and L. Ouahab, Inorg. Chem., 2006,
45, 10440.
128 F. A. Cotton, R. A. Walton and C. A. Murillo, Multiple Bonds
Between Metal Atoms, Springer-Verlag, New York, 2005, p. 454.
129 H. Kumagai, M. Akita-Tanaka, K. Inoue, K. Takahashi,
H. Kobayashi, S. Vilminot and M. Kurmoo, Inorg. Chem.,
2007, 46, 5949.
130 M.-H. Zeng, B. Wang, X.-Y. Wang, W.-X. Zhang, X.-M. Chen
and S. Gao, Inorg. Chem., 2006, 45, 7069.
131 S. Chen, J. Zhang and X. Bu, Inorg. Chem., 2008, 47, 5567.
132 J. Glerup, P. A. Goodson, D. J. Hodgson and K. Michelsen,
Inorg. Chem., 1995, 34, 6255.
133 G. Li, Z. Shi, Y. Xu and S. Feng, Inorg. Chem., 2003, 42, 117.
134 D. M. Duggan and D. N. Hendrikson, Inorg. Chem., 1975, 14,
1944.
135 V. Calvo-Pe rez, S. Ostrovsky, A. Vega, J. Pelikan, E. Spodine and
W. Haase, Inorg. Chem., 2006, 45, 644.
136 K. S. Gavrilenko, S. V. Punin, O. Cador, S. Golhen, L. Ouahab
and V. V. Pavlishchuk, J. Am. Chem. Soc., 2005, 127, 12246.
137 Y. Oka and K. Inoue, Chem. Lett., 2004, 33, 402.
138 C. Yu, S. Ma, M. J. Pechan and H.-C. Zhou, J. Appl. Phys., 2007,
101, 09E108.
139 G. Park, H. Kim, G. H. Lee, S.-K. Park and K. Kim, Bull. Korean
Chem. Soc., 2006, 27, 443.
140 S.-N. Wang, J. Bai, Y.-Z. Li, Y. Pan, M. Scheer and X.-Z. You,
CrystEngComm, 2007, 9, 1084.
141 K. S. Suslick, P. Bhyrappa, J.-H. Chou, M. E. Kosal,
S. Nakagaki, D. W. Smithenry and S. R. Wilson, Acc. Chem.
Res., 2005, 38, 283.
142 P. M. Forster, A. R. Burbank, C. Livage, G. Fe rey and
A. K. Cheetham, Chem. Commun., 2004, 368.
143 D. Cheng, M. A. Khan and R. P. Houser, Cryst. Growth Des.,
2004, 4, 599.
144 M.-X. Yao, M.-H. Zeng, H.-H. Zou, Y.-L. Zhou and H. Liang,
Dalton Trans., 2008, 2428.
145 I. Dzyaloshinski, J. Phys. Chem. Solids, 1958, 4, 241.
146 T. Moriya, Phys. Rev., 1960, 120, 91.
147 P. King, R. Cle rac, W. Wernsdorfer, C. E. Anson and
A. K. Powell, Dalton Trans., 2004, 2670.
148 Y. Y. Liu, J.-F. Ma, J. Yang and Z.-M. Su, Inorg. Chem., 2007,
46, 3027.
149 J. Luo, Y. Zhao, H. Xu, T. L. Kinnibrugh, D. Yang,
T. V. Timofeeva, L. L. Daemen, J. Zhang, W. Bao,
J. D. Thompson and R. P. Currier, Inorg. Chem., 2007, 46, 9021.
150 R.-K. Chiang, N.-T. Chuang, C.-S. Wur, M.-F. Chong and
C.-R. Lin, J. Solid State Chem., 2002, 166, 158.
151 Z.-G. Li, G.-H. Wang, H.-Q. Jia, N.-H. Hu, J.-W. Xu and
S. R. Batten, CrystEngComm, 2008, 10, 983.
152 S. Ma and H.-C. Zhou, J. Am. Chem. Soc., 2006, 128, 11734.
153 M. Murrie, S. J. Teat, H. Stckli-Evans and H. U. Gu del, Angew.
Chem., Int. Ed., 2003, 42, 4653.
154 T. A. Hudson, K. J. Berry, B. Moubaraki, K. S. Murray and
R. Robson, Inorg. Chem., 2006, 45, 3549.
155 K. W. Galloway, A. M. Whyte, W. Wernsdorfer, J. Sanchez-
Benitez, K. V. Kamenev, A. Parkin, R. D. Peacock and
M. Murrie, Inorg. Chem., 2008, 47, 7438.
156 B. Moubaraki, K. S. Murray, T. A. Hudson and R. Robson, Eur.
J. Inorg. Chem., 2008, 45254529.
157 S. Xiang, X. Wu, J. Zhang, R. Fu, S. Hu and X. Zhang, J. Am.
Chem. Soc., 2005, 127, 16352.
158 M.-H. Zeng, M.-X. Yao, H. Liang, W.-X. Zhang and
X.-M. Chen, Angew. Chem., Int. Ed., 2007, 46, 1832.
159 R.-K. Chiang, C.-C. Huang and C.-S. Wur, Inorg. Chem., 2001,
40, 3237.
160 T. D. Keene, M. B. Hursthouse and D. J. Price, New J. Chem.,
2004, 28, 558.
161 S. O. H. Gutschke, D. J. Price, A. K. Powell and P. T. Wood, Eur.
J. Inorg. Chem., 2001, 2739.
162 H. Kumagai, Y. Oka, S. Kawata, M. Ohba, K. Inoue,
M. Kurmoo and H. Okawa, Polyhedron, 2003, 22, 1917.
163 S. O. H. Gutschke, D. J. Price, A. K. Powell and P. T. Wood,
Angew. Chem., Int. Ed., 1999, 38, 1088.
164 G. E. Lewis and C. S. Kraihanzel, Inorg. Chem., 1983, 22, 2895.
165 K. G. Alley, R. Bircher, O. Waldmann, S. T. Ochsenbein,
H. U. Gudel, B. Moubaraki, K. S. Murray, F. Fernandez-Alonso,
B. F. Abrahams and C. Boskovic, Inorg. Chem., 2006, 45, 8950.
166 M.-L. Cao, H.-G. Hao, W.-X. Zhang and B.-H. Ye, Inorg. Chem.,
2008, 47, 8126.
167 X.-T. Wang, B.-W. Wang, Z.-M. Wang, W. Zhang and S. Gao,
Inorg. Chim. Acta, 2008, 361, 3895.
168 Y.-Z. Zhang, W. Wernsdorfer, F. Pan, Z.-M. Wang and S. Gao,
Chem. Commun., 2006, 3302.
169 A. Ferguson, A. Parkin, J. Sanchez-Benitez, K. Kamenev,
W. Wernsdorfer and M. Murrie, Chem. Commun., 2007, 3473.
170 L. F. Chibotaru, L. Ungur, C. Aronica, H. Elmoll, G. Pilet and
D. Luneau, J. Am. Chem. Soc., 2008, 130, 12445.
171 X.-N. Cheng, W.-X. Zhang, Y.-Y. Lin, Y.-Z. Zheng and
X.-M. Chen, Adv. Mater., 2007, 19, 1494.
172 R. J. Glauber, J. Math. Phys., 1963, 4, 294.
173 Y. Su, S. Zang, C. Duan, H. Zhu and Q. Meng, Inorg. Chem.
Commun., 2007, 10, 339.
174 S. C. Manna, E. Zangrando, J. Ribas and N. R. Chaudhuri,
Dalton Trans., 2007, 1383.
175 C. Livage, N. Guillou, J. Marrot and G. Fe rey, Chem. Mater.,
2001, 13, 4387.
176 T. Liu, Y. Zhang, Z. Wang and S. Gao, Inorg. Chem., 2006, 45,
2782.
177 S. Mandal and S. Natarajan, J. Solid State Chem., 2005, 178,
2376.
178 K. S. Gavrilenko, O. Cador, K. Bernot, P. Rosa, R. Sessoli,
S. Golhen, V. V. Pavlishchuk and L. Ouahab, Chem.Eur. J.,
2008, 14, 2034.
179 T. Yuen, C. L. Lin, J. A. Zan, L. Pan and J. Li, J. Appl. Phys.,
2005, 97, 10B 315.
180 T. K. Maji, W. Kaneko, M. Ohba and S. Kitagawa, Chem.
Commun., 2005, 4613.
181 P. D. C. Dietzel, Y. Morita, R. Blom and H. Fjellva g, Angew.
Chem., Int. Ed., 2005, 44, 6354.
182 S. R. Caskey, A. G. Wong-Foy and A. J. Matzger, J. Am. Chem.
Soc., 2008, 130, 10870.
183 N. Snejko, E. Gutie rrez-Puebla, J. L. Martinez, M. A. Monge and
C. Ruiz-Valero, Chem. Mater., 2002, 14, 1879.
184 H. Kumagai, K. W. Chapman, C. J. Kepert and M. Kurmoo,
Polyhedron, 2003, 22, 1921.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1377
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
185 X.-N. Cheng, W.-X. Zhang, Y.-Z. Zheng and X.-M. Chen, Chem.
Commun., 2006, 3603.
186 H. Kumagai, M. Akita-Tanaka, K. Inoue and M. Kurmoo,
J. Mater. Chem., 2001, 11, 2146.
187 Y.-G. Huang, D.-Q. Yuan, L. Pan, F.-L. Jiang, M.-Y. Wu,
X.-D. Zhang, W. Wei, Q. Gao, J. Y. Lee, J. Li and
M.-C. Hong, Inorg. Chem., 2007, 46, 9609.
188 X.-J. Li, X.-Y. Wang, S. Gao and R. Cao, Inorg. Chem., 2006, 45,
1508.
189 S. O. H. Gutcshke, M. Molinier, A. K. Powell and P. T. Wood,
Angew. Chem., Int. Ed. Engl., 1997, 36, 991.
190 M. Kurmoo, H. Kumagai, K. W. Chapman and C. J. Kepert,
Chem. Commun., 2005, 3012.
191 S. Konar, P. S. Mukherjee, E. Zangrando, F. Lloret and
N. R. Chaudhuri, Angew. Chem., Int. Ed., 2002, 41, 1561.
192 M. Kurmoo, H. Kumagai, M. Akita-Tanaka, K. Inoue and
S. Takagi, Inorg. Chem., 2006, 45, 1627.
193 M. Kurmoo and H. Kumagai, in preparation.
194 N. L. Rosi, M. Eddaoudi, J. Kim, M. OKeee and O. M. Yaghi,
Angew. Chem., Int. Ed., 2002, 41, 284.
195 X.-M. Zhang, Z.-M. Hao, W.-X. Zhang and X.-M. Chen, Angew.
Chem., Int. Ed., 2007, 119, 3526.
196 Y. Oka, H. Kumagai and K. Inoue, Annual Review, Institute of
Molecular Science, Department of Applied Molecular Science,
2003, p. 98.
197 Z.-G. Li, G.-H. Wang, H.-Q. Jia, N.-H. Hu and J.-W. Xu,
CrystEngComm, 2008, 10, 173.
198 H.-P. Jia, W. Li, Z.-F. Ju and J. Zhang, Dalton Trans., 2007,
3699.
199 Y.-Z. Zheng, M.-L. Tong, W.-X. Zhang and X.-M. Chen, Angew.
Chem., Int. Ed., 2006, 45, 6310.
200 P. M. Forster, A. R. Burbank, M. C. OSullivan, N. Guillou,
C. Livage, G. Fe rey, N. Stock and A. K. Cheetham, Solid State
Sci., 2005, 7, 1549.
201 L. J. De Jongh and A. R. Miedema, Adv. Phys., 1974, 23, 1.
202 M. Steiner and J. Villain, Adv. Phys., 1976, 25, 87.
203 J. Bacsa, D. Eve and K. R. Dunbar, Acta Crystallogr., Sect. C,
2005, 61, m58.
204 U. Garcia-Couceiro, O. Castillo, A. Luque, G. Beobide and
P. Roman, Inorg. Chim. Acta, 2004, 357, 339.
205 U. Garcia-Couceiro, O. Castillo, A. Luque, J. P. Garcia-Teran,
G. Beobide and P. Roman, Cryst. Growth Des., 2006, 6, 1839.
206 J. Y. Lu, M. A. Lawandy, J. Li, T. Yuen and C. L. Lin, Inorg.
Chem., 1999, 38, 26952704.
207 T. Yuen, C. L. Lin, T. W. Mihalisin, M. A. Lawandy and J. Li,
J. Appl. Phys., 2000, 87, 6001.
208 R. Feyerherm, A. Loose, M. A. Lawandy and J. Li, J. Phys.
Chem. Solids, 2002, 63, 71.
209 J. A. Lukin, S. Simizu, N. S. van der Ven and S. A. Friedberg,
J. Magn. Magn. Mater., 1995, 140144, 1669.
210 I. Sledzinska, A. Murasik and P. Fisher, J. Phys. C: Solid State
Phys., 1988, 21, 5273.
211 Z. Duan, Y. Zhang, B. Zhang and D. Zhu, Inorg. Chem., 2008,
47, 9152.
212 H. Kumagai and M. Kurmoo, unpublished results.
213 J. P. Garcia-Teran, O. Castillo, A. Luque, U. Garcia-Couceiro,
P. Roman and F. Lloret, Inorg. Chem., 2004, 43, 5761.
214 O. Castillo, A. Luque, P. Roman, F. Lloret and M. Julve, Inorg.
Chem., 2001, 40, 5526.
215 O. Castillo, A. Luque, F. Lloret and P. Roman, Inorg. Chim.
Acta, 2001, 324, 141.
216 M. B. Hursthouse, M. E. Light and D. J. Price, Angew. Chem.,
Int. Ed., 2004, 43, 472.
217 E. Coronado, J. R. Gala n-Mascaro s and C. Marti-Gastaldo,
J. Am. Chem. Soc., 2008, 130, 14987.
218 M.-H. Zeng, W.-X. Zhang, X.-Z. Sun and X.-M. Chen, Angew.
Chem., Int. Ed., 2005, 44, 3079.
219 Y.-Y. Zhang, M.-H. Zeng, Y. Qi, S.-Y. Sang and Z.-M. Liu,
Inorg. Chem. Commun., 2007, 10, 33.
220 D. J. Price, A. K. Powell and P. T. Wood, J. Chem. Soc., Dalton
Trans., 2000, 3566.
221 S. M. Humphrey and P. T. Wood, J. Am. Chem. Soc., 2004, 126,
1323613237.
222 D. S. Kim, P. M. Forster, G. D. de Delgado, S.-E. Park and
A. K. Cheetham, Dalton Trans., 2004, 3365.
223 Z. Honda, K. Katsumata, A. Kikkawa and K. Yamada, Phys.
Rev. Lett., 2005, 95, 087204.
224 M. Matsuda, S. Wakimoto, K. Kakurai, Z. Honda and
K. Yamada, Phys. Rev. B, 2007, 75, 012405.
225 A. Kikkawa, K. Katsumata, Z. Honda and K. Yamada, J. Magn.
Magn. Mater., 2007, 310, 1239.
226 F. Zigan and R. Rothbauer, Neues Jahrb. Mineral Monatsh, 1967,
137; J. B. Parise, K. Leinenweber, D. J. Weidner, K. Tan and
R. B. Von Dreele, Am. Mineral., 1994, 79, 193.
227 T. Takada, Y. Bando, M. Kiyama, H. Myamoto and T. Sato,
J. Phys. Soc. Jpn., 1966, 21, 2726.
228 M. Sorai, A. Kosaki, H. Suga and S. Seki, J. Chem. Thermodyn.,
1969, 1, 119.
229 M. Kurmoo, MetalOrganic and Organic Molecular Magnets, ,
Spec. Publ. R. Soc. Chem., Cambridge, 2000, vol. 252, p. 185.
230 M. Ben Salah, S. Vilminot, M. Richard-Plouet, G. Andre ,
T. Mhiri and M. Kurmoo, Chem. Commun., 2004, 2548.
231 M. Ben Salah, S. Vilminot, G. Andre , M. Richard-Plouet,
T. Mhiri, S. Takagi and M. Kurmoo, J. Am. Chem. Soc., 2006,
128, 7972.
232 M. Kurmoo, H. Kumagai, S. M. Hughes and C. J. Kepert, Inorg.
Chem., 2003, 42, 6709.
233 P. M. Forster, M. M. Tafoya and A. K. Cheetham, J. Phys.
Chem. Solids, 2004, 65, 11.
234 A. Rujiwatra, C. J. Kepert and M. J. Rosseinsky, Chem.
Commun., 1999, 2307.
235 A. Rujiwatra, C. J. Kepert, J. B. Claridge, M. J. Rosseinsky,
H. Kumagai and M. Kurmoo, J. Am. Chem. Soc., 2001, 123,
10584.
236 P. Day, Acc. Chem. Res., 1988, 211, 250.
237 H. Aruga Katori and K. Katsumata, Phys. Rev. B, 1996, 54,
R9620.
238 M. Kurmoo and S. Takagi, unpublished results.
239 M. Kurmoo, Chem. Mater., 1999, 11, 3370.
240 M. Kurmoo, J. Mater. Chem., 1999, 9, 2595.
241 M. Kurmoo, Mol. Cryst. Liq. Cryst., 2000, 341, 395.
242 X.-M. Zhang, X.-H. Zhang, H.-S. Wu, M.-L. Tong and S. W. Ng,
Inorg. Chem., 2008, 47, 7462.
243 A. Mesbah, A. Carton, L. Aranda, T. Mazet, F. Porcher and
M. Francois, J. Solid State Chem., 2008, 181, 3229.
244 M. Kurmoo, H. Kumagai, M. A. Green, B. W. Lovett,
S. J. Blundell, A. Ardavan and J. Singleton, J. Solid State Chem.,
2001, 159, 343.
245 M. Kurmoo and H. Kumagai, Mol. Cryst. Liq. Cryst., 2002, 376,
555.
246 B. W. Lovett, S. J. Blundell, H. Kumagai and M. Kurmoo, Synth.
Met., 2001, 121, 1814.
247 M. Kurmoo, Philos. Trans. R. Soc. London, Ser. A, 1999, 357,
3041.
248 Y.-Z. Zheng, M.-L. Tong, W.-X. Zhang and X.-M. Chen, Chem.
Commun., 2006, 165.
249 Y.-L. Wang, D.-Q. Yuan, W.-H. Bi, X. Li, X.-J. Li, F. Li and
R. Cao, Cryst. Growth Des., 2005, 5, 1849.
250 F.-C. Liu, Y.-F. Zeng, J. Jiao, X.-H. Bu, J. Ribas and
S. R. Batten, Inorg. Chem., 2006, 45, 2776.
251 M.-H. Zeng, S. Gao and X.-M. Chen, Inorg. Chem. Commun.,
2004, 7, 864.
252 M.-H. Zeng, X.-L. Feng, W.-X. Zhang and X.-M. Chen, Dalton
Trans., 2006, 5294.
253 P. G. Slade, E. W. Rodoslovich and M. Raupach, Acta Crystallogr.,
Sect. B, 1971, 27, 2432.
254 R. P. Eckberg, W. E. Hateld and B. D. Losee, Inorg. Chem.,
1974, 13, 740.
255 F. L. Pratt, P. J. Baker, S. J. Blundell, T. Lancaster, M. A. Green
and M. Kurmoo, Phys. Rev. Lett., 2007, 99, 017202/1.
256 S. J. Blundell, Comtemp. Phys., 1999, 40, 175.
257 H. Kawamura, J. Phys.: Condens. Matter, 1998, 10, 4707.
258 C. Livage, C. Egger and G. Fe rey, Chem. Mater., 2001, 13, 410.
259 C. Livage, C. Egger and G. Fe rey, Chem. Mater., 1999, 11, 1546.
260 C. Livage, C. Egger, M. Nogues and G. Fe rey, J. Mater. Chem.,
1998, 8, 2743.
261 L.-S. Long, X.-M. Chen, M.-L. Tong, Z.-G. Sun, Y.-P. Ren,
R.-B. Huang and L.-S. Zheng, J. Chem. Soc., Dalton Trans., 2001,
2888.
1378 | Chem. Soc. Rev., 2009, 38, 13531379 This journal is c The Royal Society of Chemistry 2009
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online
262 H. Kumagai, Y. Oka, K. Inoue and M. Kurmoo, J. Phys. Chem.
Solids, 2004, 65, 55.
263 Y. Oka, H. Kumagai, K. Inoue and M. Kurmoo, Mol. Cryst. Liq.
Cryst., 2002, 379, 265.
264 H. Kumagai, N. Kyritsakas, Y. Oka, K. Inoue and M. Kurmoo,
Mol. Cryst. Liq. Cryst., 2002, 379, 217.
265 H. Kumagai, Y. Oka, K. Inoue and M. Kurmoo, J. Chem. Soc.,
Dalton Trans., 2002, 3442.
266 I. Pantenburg and U. Ruschewitz, Z. Anorg. Allg. Chem., 2002,
628, 1697.
267 F. S. Delgado, M. Hernandez-Molina, J. Sanchiz, C. Ruiz-Pe rez,
Y. Rodriguez-Martin, T. Lopez, F. Lloret and M. Julve,
CrystEngComm, 2004, 6, 106.
268 E. W. Lee, Y. J. Kim and D.-Y. Jung, Inorg. Chem., 2002, 41, 501.
269 C. Livage, C. Egger, M. Nogues and G. Fe rey, C. R. Acad. Sci.
Paris, 2001, 4, 221.
270 M. Kurmoo, P. J. Baker and S. J. Blundell, unpublished results.
271 M. Nakakuma, J. Yamamoto, S. Takagi, M. Mito, H. Deguchi,
M. Yoshihiro, N. Kyritsakas and M. Kurmoo, J. Magn. Magn.
Mater., 2007, 310, 1435.
272 M. E. Lines, J. Phys. Chem. Solids, 1970, 31, 101.
273 S. J. Rettig, R. C. Thompson, J. Trotter and S. Xia, Inorg. Chem.,
1999, 38, 1360.
274 H. Tamaki, Z. J. Zhong, N. Matsumoto, S. Kida, M. Koikawa,
N. Achiwa, Y. Hashimoto and H. Okawa, J. Am. Chem. Soc.,
1992, 114, 6974.
275 C. Mathoniere, C. J. Nuttall, S. G. Carling and P. Day, Inorg.
Chem., 1996, 35, 1201.
276 S. Decurtins, H. W. Schmalle, P. Schneuwly and H. R. Oswald,
Inorg. Chem., 1993, 32, 1888.
277 E. Coronado, J. R. Gala n-Mascaro s and C. Marti-Gastaldo,
J. Mater. Chem., 2006, 16, 2685.
278 Y.-Q. Sun, J. Zhang and G.-Y. Yang, Dalton Trans., 2003, 3634.
279 A. Rodriguez-Dieguez, J. Cano, R. Kiveka, A. Debdoubi and
E. Colacio, Inorg. Chem., 2007, 46, 2503.
280 E. Coronado, J. R. Gala n-Mascaro s and C. Marti-Gastaldo,
Inorg. Chem., 2007, 46, 8108.
281 E. Coronado, J. R. Gala n-Mascaro s, C. Marti-Gastaldo,
J. C. Waerenborgh and P. Gaczynski, Inorg. Chem., 2008, 47,
6829.
282 M. Kurmoo, unpublished results.
283 B. Zhang, Y. Zhang, J. Zhang, J. Li and D. Zhu, Dalton Trans.,
2008, 5037.
284 D. J. Price, A. K. Powell and P. T. Wood, Dalton Trans., 2003,
2478.
285 Zhu, W. Gu, F.-Y. Cheng, X. Liu, J. Chen, S.-P. Yan and
D.-Z. Liao, CrystEngComm, 2008, 10, 963.
286 Q.-X. Jia, Y.-Q. Wang, Q. Yue, Q.-L. Wang and E.-Q. Gao,
Chem. Commun., 2008, 4894.
287 X.-Y. Wang, H.-Y. Wei, Z.-M. Wang, Z.-D. Chen and S. Gao,
Inorg. Chem., 2005, 44, 572.
288 R. Kuhlman, G. L. Schimek and J. W. Kolis, Inorg. Chem., 1999,
38, 194.
289 Z. Wang, B. Zhang, H. Fujiwara, H. Kobayashi and M. Kurmoo,
Chem. Commun., 2004, 416.
290 Z. Wang, B. Zhang, M. Kurmoo, M. A. Green, H. Fujiwara,
T. Otsuka and H. Kobayashi, Inorg. Chem., 2005, 44, 12301237.
291 D. N. Dybtsev, H. Chun, S. H. Yoon, D. Kim and K. Kim,
J. Am. Chem. Soc., 2004, 126, 32.
292 M. Viertelhaus, P. Adler, R. Cle rac, C. E. Anson and
A. K. Powell, Eur. J. Inorg. Chem., 2005, 692.
293 M. Viertelhaus, C. E. Anson and A. K. Powell, Z. Anorg. Allg.
Chem., 2005, 631, 2365.
294 J. A. Rood, B. C. Noll and K. W. Henderson, Inorg. Chem., 2006,
45, 5521.
295 Z. Wang, Y. Zhang, T. Liu, M. Kurmoo and S. Gao, Adv. Funct.
Mater., 2007, 17, 1523.
296 Z. Wang, B. Zhang, Y. Zhang, M. Kurmoo, T. Liu, S. Gao and
H. Kobayashi, Polyhedron, 2007, 26, 2207.
297 Z. Wang, Y. Zhang, M. Kurmoo, T. Liu, S. Vilminot, B. Zhao
and S. Gao, Aust. J. Chem., 2006, 59, 617.
298 B. Zhang, Z. Wang, M. Kurmoo, S. Gao, K. Inoue and
H. Kobayashi, Adv. Funct. Mater., 2007, 17, 577.
299 J. A. Stride, M. Kurmoo and Z. Wang, Physica B: Condens.
Matter, 2006, 385386, 465.
300 H. B. Cui, K. Takahashi, Y. Okano, H. Kobayashi, Z. Wang and
A. Kobayashi, Angew. Chem., Int. Ed., 2005, 44, 6508.
301 H. Cui, Z. Wang, K. Takahashi, Y. Okano, H. Kobayashi and
A. Kobayashi, J. Am. Chem. Soc., 2006, 128, 15074.
302 M. Kurmoo, unpublished results.
303 Y.-L. Fu, J.-L. Ren, Z.-W. Xu and S. W. Ng, Acta Crystallogr.,
Sect. E, 2005, 61, m2395.
304 S. T. Bramwell and M. J. P. Gingras, Science, 2001, 94, 1495.
305 J. E. Greedan, J. Mater. Chem., 2001, 11, 37.
306 A. P. Ramirez, in Handbook of Magnetic Materials, ed. K. J. H.
Buschow, Elsevier Science, Amsterdam, 2001, vol. 13, p. 423.
307 D. Grohol, K. Matan, J.-H. Cho, S.-H. Lee, J. W. Lynn,
D. G. Nocera and Y. S. Lee, Nat. Mater., 2005, 4, 323.
308 Magnetic Systems with Competing Interactions: Frustrated Spin
Systems, ed. H. T. Diep, World Scientic, Singapore, 1994.
309 X.-Y. Wang, L. Gan, S.-W. Zhang and S. Gao, Inorg. Chem.,
2004, 43, 4615.
310 Z. Wang, B. Zhang, T. Otsuka, K. Inoue, H. Kobayashi and
M. Kurmoo, Dalton Trans., 2004, 2209.
311 Z. Wang, X. Zhang, S. R. Batten, M. Kurmoo and S. Gao, Inorg.
Chem., 2007, 46, 8439.
312 Z. Wang, B. Zhang, K. Inoue, H. Fujiwara, T. Otsuka,
H. Kobayashi and M. Kurmoo, Inorg. Chem., 2007, 46, 437.
313 P. Jain, N. S. Dalal, B. H. Toby, H. W. Kroto and
A. K. Cheetham, J. Am. Chem. Soc., 2008, 130, 10450.
314 M. Kurmoo, unpublished results.
315 M. Matsura, H. W. J. Blo te and W. J. Huiskamp, Physica, 1970,
50, 444.
316 K. Takeda, T. Haseda and M. Matsura, Physica, 1971, 52, 225.
317 J. Wang, L.-L. Zheng, C.-J. Li, Y.-Z. Zheng and M.-L. Tong,
Cryst. Growth Des., 2006, 6, 357.
318 D. J. Price, F. Lionti, R. Ballou, P. T. Wood and A. K. Powell,
Philos. Trans. R. Soc. London, Ser. A, 1999, 357, 3099.
319 M. Molinier, D. J. Price, P. T. Wood and A. K. Powell, J. Chem.
Soc., Dalton Trans., 1997, 4061.
320 C. N. R. Rao, S. Natarajan and R. Vaidhyanathan, Angew.
Chem., Int. Ed., 2004, 43, 1466.
321 S. Decurtins, Philos. Trans. R. Soc. London, Ser. A, 1999, 357,
30253040.
322 M. Hernandez-Molina, F. Lloret, C. Ruiz-Pe rez and M. Julve,
Inorg. Chem., 1998, 37, 4131.
323 W.-X. Zhang, W. Xue, J.-B. Lin, Y.-Z. Zheng and X.-M. Chen,
CrystEngComm, 2008, 10, 1770.
This journal is c The Royal Society of Chemistry 2009 Chem. Soc. Rev., 2009, 38, 13531379 | 1379
D
o
w
n
l
o
a
d
e
d

b
y

N
A
T
I
O
N
A
L

T
E
C
H
N
I
C
A
L

U
N
I
V
E
R
S
I
T
Y

O
F

A
T
H
E
N
S

o
n

2
4

F
e
b
r
u
a
r
y

2
0
1
3
P
u
b
l
i
s
h
e
d

o
n

2
4

F
e
b
r
u
a
r
y

2
0
0
9

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
8
0
4
7
5
7
J
View Article Online

Вам также может понравиться