You are on page 1of 28

Coordination Chemistry Reviews 256 (2012) 2944–2971

Contents lists available at SciVerse ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

Syntheses, structures and properties of structurally characterised complexes of


imide-based ligands
Matthew G. Cowan, Sally Brooker ∗
Department of Chemistry and the MacDiarmid Institute for Advanced Materials and Nanotechnology, University of Otago, PO Box 56, Dunedin, 9054, New Zealand

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2945
2. Synthesis of imide ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2946
3. Alkali and alkaline earth metal complexes of imides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2946
3.1. Summary of alkali and alkaline earth metal complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2946
4. Iron complexes of imides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2946
4.1. The discrete complexes [FeII/III (imide)2 ]0/+ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2946
4.2. Complexes assembled from [FeII/III (imide)2 ]0/+ building blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2948
4.3. Homometallic complexes derived from [FeIII (imide)Xn ] building blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2951
4.4. Heterometallic complexes of the anionic [FeIII (bpca)(CN)3 ] building block . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2952
4.4.1. Overview of heterometallic complexes of the anionic [FeIII (bpca)(CN)3 ] building block . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2952
4.4.2. Manganese . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2953
4.4.3. Copper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2954
5. Copper complexes of imides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2954
5.1. Overview of copper complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2954
5.2. The discrete complex [CuII (bpca)2 ] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2955
5.3. [CuII (bpca)2 ] used as a building block . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2955
5.4. CuII (imide)X·solvents complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2955
5.5. Discrete {[CuII (imide)]2 X} complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2957
5.6. Chain complexes built from [Cu(imide)]+ units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2960
5.7. Copper complexes with unusual imides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2961
5.8. Heterometallic complexes of [Cu(imide)]+ units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2961
6. Other first row transition metal ion complexes of imides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2963
6.1. Titanium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2963
6.2. Chromium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2963
6.3. Manganese . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2964
6.4. Cobalt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2964
6.5. Nickel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2965
6.6. Zinc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2966
7. Rhodium, rhenium and platinum complexes of imides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2966
8. Lanthanide complexes of imides. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2968
8.1. Overview of lanthanide complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2968
8.2. Lanthanide complexes of imide ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2968

Abbreviations: 1,2-dtcr, 1,2-dithiocroconate; 1,3-dtcr, 1,3-dithiocroconate; 1,3-dtsq, 1,3-dithiosquarate; 4,4 -bipy, 4,4 -bipyridine; 4,4 -dmbipy, 4,4 -dimethylpyridine;
5,5 -dmbipy, 5,5 -dimethylpyridine; AC, alternating current; CFSE, crystal field stabilisation energy; DC, direct current; DA, diacetamide; Dca, dicyanamide; DFT,
density functional theory; DIEDO, diiodoethylenedioxotetrathiavalene; DIET, diiodoethylenedithotetrathiavalene; DMF, dimethylformamide; DMSO, dimethylsul-
foxide; EDTA, ethylenediaminetetraacetic acid; EPR, electron paramagnetic resonance; H, exchange Hamiltonian; Hbpca, bis(2-pyridylcarbonyl)amide; Hbpcam,
bis(2-pyrimidylcarbonyl)amide; Hbmpca, bis(4-methyl-2-pyridylcarbonyl)amide; HbpBuca, bis(4-tBu-2-pyridyl-carbonyl)amide; H3 bcpca, bis(6-carboxypyridine-2-
carbonyl)amine; hfac, hexafluoroacetylacetonate; Hpypzca, N-(2-pyrazylcarbonyl)-2-pyridinecarboxamide; Hdpzca, N-(2-pyrazylcarbonyl)-2-pyrazinecarboxamide; Hpqca,
(2-pyridylcarbonyl)-2-quinolylcarbonyl)diimide; J, coupling constant; Ln, lanthanide ion; Ma, racemic mandelate; Mpa, racemic ␣-methoxyphenylacetate; NHE, normal
hydrogen electrode; S, spin multiplicity; Salcy, N,N -(1,2-cyclohexanediylethylene)bis(salicylideneiminato) dianion; SCE, saturated calomel electrode; SCM, single chain
magnet; SCO, spin crossover; SMM, single molecule magnet; SSCE, silver/silver chloride electrode; tcm, tricyanomethanide; Tp, tris(pyrazolyl)hydroborate; tpy, terpyridine;
TPyT, 2,4,6-tris(2-pyridyl)-1,3,5-triazine; TPymT, 2,4,6-tris(2-pyrimidyl)-1,3,5-triazine; UV–Vis, ultraviolet visible.
∗ Corresponding author. Tel.: +64 3 479 7919; fax: +64 3 479 7906.
E-mail address: sbrooker@chemistry.otago.ac.nz (S. Brooker).

0010-8545/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ccr.2012.06.004
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2945

8.3. Co-crystallisation of transition metal imide complexes with lanthanide ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2968
8.4. Lanthanide complexes of transition metal/imide metalloligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2969
9. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2970
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2970
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2970

a r t i c l e i n f o a b s t r a c t

Article history: This review details the syntheses, structures and properties of imide-based ligands (phthalimide ana-
Received 31 March 2012 logues have been excluded) and structurally characterised coordination complexes of them. The surveyed
Received in revised form 12 June 2012 imide complexes, including 3D-metal-organic frameworks and 1D-polymers formed from monometallic
Accepted 12 June 2012
building block complexes, exhibit a wide range of interesting properties, including single chain and single
Available online 20 June 2012
molecule magnetism, spin crossover, reversible electrochemistry and luminescence.
© 2012 Elsevier B.V. All rights reserved.
Keywords:
Imide ligand
Monometallic
Trimetallic
Polymetallic
Secondary building units
Heterometallic
Homometallic
Transition metal ion
Lanthanide ion
Single chain magnetism
Spin crossover
Electrochemistry
Structure

1. Introduction Complexes of Hbpca-based imide ligands often feature multiple


secondary/packing interactions in the solid state. These arise due
Since the first structurally characterised imide complexes to the presence of (a) the imide group, which is usually involved in
were reported in 1976 by Lerner and Lippard [1,2] a wealth of hydrogen bonding [3] and/or carbonyl–carbonyl interactions [4],
imide coordination chemistry has been revealed. The main focus and (b) the flat, highly conjugated aromatic rings, which are com-
of these studies has been on complexes of the ligand bis(2- monly involved in parallel offset stacking, C H· · ·␲ and anion· · ·␲
pyridylcarbonyl)amide (Hbpca), however a range of related ligands interactions [5,6]. These packing interactions make imide ligands
has emerged in recent years (Fig. 1). Ligands based on the bpca attractive for producing and manipulating the properties of mag-
anion have two possible binding sites. The N3 donor set consists netically interesting materials.
of the deprotonated, negatively charged imide nitrogen atom and The highly conjugated, and hence flat, nature of the deproto-
the two aromatic nitrogen atoms, whilst the O2 set comprises the nated imide ligands has also led to the predictable formation of
two imide oxygen atoms (Fig. 2). Most complexes feature either a large number of mononuclear complexes of octahedral metal
an N3 or an O2 binding mode, but the combined N3 :␮-O2 binding ions, [MII/III (bpca)2 ]0/+ , which feature the N3 ligand binding mode.
mode is also common, especially in mixed metal 3d–4f complexes Many of these have been designed to be used in the assembly of
(Fig. 2). Other rare binding modes, seen only in one study each, are 1-, 2- and 3-D polymeric complexes which consist of mononu-
described later (Figs. 18, 29 and 38, Sections 5.4, 5.7 and 6.4). clear building blocks bridged by transition metal or lanthanide ions.

O O O O O O O O
R R N N
N N N N N
H H H H
N N N N N N N N

Hbpca R = CH3, Hbmpca Hbpcam Hpypzca


R = tBu, HtBu-bpca
R = C5H5, HPh-bpca

O O O O O O O O

N N N N
H H H H
N N N N N N

diacetamide (DA)
N N O OR' R'O O
Hpqca / Hbqca
R' = H, H3bcpca
R' =CH3, Hbcpc
Hbbpca

Fig. 1. Imide ligands which have been used to produce coordination complexes.
2946 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

[21]. In a similar fashion, copper(II) acetate has been used to


O O O O
combine two amines, ultimately generating copper(II) imide com-
N plexes (Fig. 3, route iii) [22]. The proposed mechanism for this
N
process involves initial oxidation of one amine to the corre-
N M N
M sponding primary amide, which then reacts with a second amine
N3
generating ammonia and the secondary amide, with the latter sub-
sequently oxidised to the imide. This route was used to produce
[CuII (bpca)(OAc)(H2 O)]·H2 O (61%), [CuII (bqca)(OAc)(H2 O)] (58%)
and [CuII (pqca)(OAc)(H2 O)]3 [CuII 2 (OAc)4 (EtOH)2 ]1.5 (52%), which
M M'
O O O O contain the coordinated pqca anion (Fig. 1). In all three cases the
imide ligands were then isolated quantitatively by removal of cop-
N N per(II) using sodium EDTA.
H
N N N M N The oxidation of methylene groups situated between aromatic
rings and amide functionalities has also produced imide complexes
O2 N3:μ - O2 (Fig. 3, route ii) [23–25]. Specifically, the amide ligand N-(2-
picolyl)picolinamide was treated with sodium methoxide and salts
Fig. 2. Top left: The search fragment used to identify the material included in this of iron(II) and cobalt(II), producing the complexes [FeIII (bpca)2 ]BF4
review. Top right: Coordination to the deprotonated N3 donor set (N3 ) is the most
(23–45%) and [CoIII (bpca)2 ]ClO4 (50–75%) [24]. The authors deter-
common binding observed. Bottom left: Coordination to the oxygen atoms (O2 )
while the imide nitrogen remains protonated is commonly observed for alkali and mined that oxidation by atmospheric oxygen proceeded over
alkaline earth metal ions. Bottom right: A combination of both modes (N3 :␮-O2 ) is around 10 h and could be accelerated by the addition of hydro-
commonly observed for systems where M is a transition metal ion and M is a tran- gen peroxide. Similarly, in this research group electrochemistry
sition metal ion or lanthanide ion. Additional but rare binding motifs are shown in was used to observe the partial air oxidation of the amide complex
Figs. 18, 29 and 38.
[CoIII (L1M )2 ]+ (L1M = N-(2-pyridylmethyl)pyrazine-2-carboxamide)
to the analogous imide (Fig. 1) complex [CoIII (pypzca)2 ]+ , however
Highlights include single chain [7,8] and single molecule magnets that complex was not isolated [25]. In principle this method is only
[9], metal-organic frameworks [10] and materials which display limited by the synthetic availability of the initial amide ligand and
electrochemical reversibility [11], spin crossover [12] and lumines- could be used to prepare a wide variety of imide ligands.
cence [13,14]. Synthetic methods for generating imide ligands without requir-
This review summarises the preparation and coordination ing a metal ion have also emerged. Amination of anhydrides, and
chemistry of imide ligands reported before June 2011 (Fig. 1). The oxidation of amides using Dess–Martin Periodinane (DMP) [26],
review material was identified by means of a Cambridge Structural have been used, but these methods are yet to be applied to the
Database (Version 5.33)[15,16] search for the fragment shown in synthesis of ligands such as those described herein. Other metal-
Fig. 2, followed by exclusion of all structures that (a) did not coor- free routes utilise either the combination of two acid chlorides with
dinate via either the imide nitrogen and/or the oxygen atom of the ammonia (Fig. 3, route v) [27] or the combination of an acid chloride
imide or (b) were phthalimide derivatives. and an amide (Fig. 3, route vi) [13,28]. The wide varieties of starting
materials, straightforward synthesis and large scales possible with
2. Synthesis of imide ligands these methods make them extremely attractive for generating a
wide variety of imide ligands. Furthermore, the ability to choose
The synthesis of the imide ligands Hbpca and Hbpcam (Fig. 1) both an amide and acid chloride starting material allows the highly
was first reported in 1976. They formed unexpectedly during an controllable preparation of previously uncommon non-symmetric
investigation into complexes of 2,4,6-tris(2-pyridyl)-1,3,5-triazine imide ligands.
(TPyT) and the pyrimidine analogue (TPymT) (Fig. 3, route i) [1].
Despite the stability of TPyT and TPymT in boiling water and
3. Alkali and alkaline earth metal complexes of imides
6 mol L−1 hydrochloric acid, hydrolysis rapidly occurred in the
presence of transition metal ions, particularly copper(II). The result-
3.1. Summary of alkali and alkaline earth metal complexes
ing imide ligands (Fig. 1) were coordinated to the copper ions,
producing [CuII (bpca)2 ] and [CuII (bpcam)2 ] [1,2]. The free ligand
A large number of simple diacetamide (DA, Fig. 1) complexes
Hbpca was then isolated by treating the copper(II) complexes
with the general formula [M(DA)2–5 ]X1–2 ·(H2 O)0–2 were prepared
with sodium EDTA. Subsequent attempts to coordinate TPyT to
with Li+ , Na+ , K+ , Mg2+ , Ca2+ , Sr2+ and Ba2+ and counter-ions includ-
rhodium(III) and ruthenium(II) generated [RhIII (bpca)2 ]PF6 and
ing NO3 − , ClO4 − , Cl− , Br− and I− [29]. Attempted complexations
[RuII (TPyT)2 ](PF6 )2 [17]. From these results it was determined
were unsuccessful when the metal halide bond was calculated as
that an electron-withdrawing effect, rather than induced angular
greater than 63% ionic character [30] or where the cation:anion
strain, was the major factor responsible for metal ion promoted
radius ratio was more than 0.78. Infrared studies of these complexes
hydrolysis. Recent studies found that under appropriate reac-
suggested that the imide nitrogen atom remained protonated and
tion conditions copper complexes of triazines can be produced
that the neutral DA ligand coordinated via the imide oxygen atoms.
[18] and that the hydrolysis mechanism can even be reversed
This was subsequently confirmed by a number of structural studies
to generate triazine ligands from imine starting materials [19].
(Table 1).
Most recently, copper(II)-assisted hydrolysis of 1,3,5-tris(4-alkyl-
2-pyridyl)triazine has been used to produce further imide ligands,
including bis(4-methyl)-2-pyridylcarbonyl)amine (Hbmpca) [20] 4. Iron complexes of imides
and a further series of substituted Hbpca derivatives [8].
Cis- and trans-Pt(NH3 )2 Cl2 have been used as both an ammo- 4.1. The discrete complexes [FeII/III (imide)2 ]0/+
nia source and a scaffold to facilitate the transformation of two
equivalents of 2-pyridinecarbaldehyde into the corresponding A series of mononuclear ‘building block’ com-
amide, which is then oxidised by exposure to air to produce plexes, [FeII (bpca)2 ]·H2 O, [FeIII (bpca)2 ](NO3 )·1.67H2 O and
the square planar imide complex [PtII (bpca)Cl] (Fig. 3, route iv) [FeIII (bpca)2 ](ClO4 )·2CH3 CN, was reported in 1993 [12]. The
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2947

TPyT: Y = CH; R,Rii = H


TPymT: Y = N; R,Rii = H

R R''

Y N

N N
R Y Y R
N
N N

R'' R''

1. CuII or RhIII
(i)
2. Na2EDTA

O O O
O O
(vi) (ii)
Cl X NH2 R Y Y R X N
X N H
N N H 1. MII N N
N N
2. O2
R' R'
R' / R'' R'/ R'' HL1M: X = N

(v) (iii)
1. PCl5 1. [Cu(OAc)2]
O 2. NH3 2. Na2EDTA
(iv) H 2N
2 OH 2
1. [PdCl2(NH3)2] N
N

Fig. 3. Reported synthetic routes for obtaining Hbpca and related ligands (R = H, CH3 ; t Bu; Ph; X = CH, N; Y = CH, N); (R = H, COOH, bipyridine). (i) Hydrolysis of symmetric
triazine systems using CuII or RhIII followed by its removal with EDTA. (ii) Oxidation of amide based ligands using CoII or FeII , note however that only the MIII complexes have
been isolated, not the pure ligand. iii) The combination of amines using copper(II) acetate. This has been used to produce symmetric and non-symmetric ligands. (iv) The
combination of aldehydes using trans- or cis- [PtCl2 (NH3 )2 ] (v). Condensation of symmetric carboxylic acids with ammonia via in situ generation of the acid chloride. (vi)
Preparation of the acid chloride followed by reaction with an amide, allowing symmetric or non-symmetric ligands to be produced.

Table 1
Selected information on structurally characterised complexes of DA (Fig. 1): space group, metal ion coordination number and geometry, metal ion to imide oxygen atom
bond lengths.

Complex Space group Coordination Geometry M–Oimide (Å) Reference


number

[Li(DA)2 NO3 ] P-1 5 Trigonal bipyramid 1.987, 2.637 [31]


2.032, 2.068a
[Na(DA)2 Br] Cmca 4 Square planar 2.332, 2.314b [32]
[Na(DA)2 ClO4 ] P21 /c 6 Octahedral 2.385, 2.314 [33]
2.358, 2.337
2.505, 2.492
[K(DA)2 ]I P-1 4 Square planar 2.717, 2.642 [34]
2.661, 2.730
[Mg(DA)2 (H2 O)2 ](ClO4 )2 ·2DA P-1 6 Octahedral 2.041, 2.048c [35]
[Ca(DA)4 ]Br2 C2/c 8 Antiprismatic 2.405, 2.461b [36]
2.425, 2.373b
[Ca(DA)4 ](ClO3 )2 ·H2 O P2/c 8 Square antiprism 2.405, 2.407b [37]
2.402, 2.449b
[Ca(DA)4 ](ClO4 )2 ·DA P2/c 8 Antiprismatic 2.382, 2.449 [38]
2.430, 2.394
2.408, 2.439
2.423, 2.389
[Sr(DA)4 H2 O](ClO4 )2 P21 /m 9 Mono-capped square 2.544, 2.626b [39]
antiprism 2.692, 2.552
2.591, 2.591a
[Ba(DA)5 ](ClO4 )2 P21 /c 10 Symmetrically bicapped 2.80, 2.82 [40]
square antiprism 2.83, 2.91
2.73, 2.84
2.79, 2.89
2.81, 2.74
a
The M–X bond distance of the non-imide donor has not been included.
b
Imide ligands which generate a symmetry equivalent set of O2 donors.
c
Imide ligands which generate two symmetry equivalent sets of O2 donors.
2948 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

The complexes [FeII/III (bpca)2 ]0/+ appear to be air stable as solids


and in solution, and [FeII (bpca)2 ]·H2 O has been used as a mild
reducing agent [11]. The cyclic voltammogram of [FeII (bpca)2 ]
in chloroform shows a quasi-reversible peak at 0.35 V vs a sil-
ver/silver chloride electrode (SSCE). Electrochemical studies of
[FeIII (bpmca)2 ](ClO4 ) in nitromethane show that it undergoes a
reversible one electron process (∼−0.2 V vs Ag/Ag+ ), at a more neg-
ative potential than the [FeII/III (bpca)2 ]0/+ couple (∼0.0 V vs Ag/Ag+
in nitromethane). This indicates that, as expected, the inclusion
of electron donating methyl groups in the imide ligands stabilises
iron(III) relative to the iron(II) oxidation state [20], which may be
why [FeII (bpmca)2 ] has not been reported.

4.2. Complexes assembled from [FeII/III (imide)2 ]0/+ building blocks


Fig. 4. The crystal structures of [FeII (bpca)2 ]·H2 O (left) and cation of
[FeIII (bpca)2 ](ClO4 )·2CH3 CN (right). Solvent molecules, counterions and hydrogen
atoms have been omitted for clarity. The mononuclear unit [FeII (bpca)2 ] has been combined with
Source: These figures were generated from data obtained from the CCDC as published [FeIII (H2 O)6 ](ClO4 )3 to prepare black crystals of the discrete
originally in references [12] and [23], respectively. trimetallic complex [FeII (ClO4 )2 [FeIII (bpca)2 ]2 ](ClO4 )2 ·4(CH3 NO2 )
and of the polymeric single chain magnet originally
reported as {[FeII (ClO4 )2 [FeIII (bpca)2 ]](ClO4 )·2.5(CH3 NO2 )·
complexes were prepared in a mixture of water and acetone 0.5(CH2 Cl2 )·0.5(H2 O)}∞ (Fig. 5) [11,43], and later reported as
by a 2:1 combination of Hbpca with iron(II) sulphate, iron(III) the ·3CH3 NO2 solvate which was investigated as a single molecule
nitrate and iron(II) perchlorate, respectively. Addition of sodium magnet (Tables 2 and 3) [8,42,44]. For an introduction to the field
hydroxide increased the pH of the reaction solution, promoting of single molecule and single chain magnets please see reference
coordination and precipitation of the complexes from solution. [41]. The nature of the reaction product was controlled by using a
The complex [FeIII (bpca)2 ](ClO4 ) has also been produced by 2:1 or 1:1 ratio of [FeII (bpca)2 ] to [FeIII (H2 O)6 ](ClO4 )3 , respectively.
air oxidation of bis(2-picolyl)amine in the presence of iron(II) During the synthesis a redox reaction occurs whereby [FeII (bpca)2 ]
perchlorate [23]. Complexes have also been prepared from is oxidised by [FeIII (H2 O)6 ](ClO4 )3 and the resulting chains are
methyl, tert-butyl and phenyl disubstituted ligands (Fig. 1), constructed from alternating [FeIII (bpca)2 ]+ and [FeII (ClO4 )2 ] units
producing [FeIII (bpmca)2 ](ClO4 ) [20], [FeIII (t Bu-bpca)2 ]ClO4 and (Fig. 5).
[FeIII (Ph-bpca)2 ]ClO4 [8]. The discrete trimetallic complex consists of two [FeIII (bpca)2 ]+
The analysis of the crystal structures of these complexes is com- units bridged via the imide oxygen atoms by one [FeII (ClO4 )2 ] unit.
plicated by their spin crossover behaviour (for an introduction The central high spin iron(II) ion sits on an inversion centre and
to the field of spin crossover please see reference [41]). Electron has an octahedral coordination sphere, formed by four imide oxy-
paramagnetic resonance (EPR) and magnetic susceptibility mea- gen atoms in a square plane and capped axially by two ClO4 anions.
surements on the iron(II) complex [FeII (bpca)2 ]·H2 O show that it The axial FeII –Operchlorate bonds (2.253 Å) are significantly longer
is mostly low spin but has a residual paramagnetic fraction of 0.18 than the FeII –Oimide bonds (2.057, 2.043 Å). The bond lengths of the
(2.09 BM) at room temperature which drops to 0.10 (1.53 BM) at iron(III) centre are consistent with a low spin electronic configu-
20 K [12]. The nature of the gradual spin crossovers of the two struc- ration (Table 2). While a study of the properties of this discrete
turally characterised iron(III) complexes, [FeIII (bpca)2 ]X, is anion complex appears to have been overlooked, a significant amount of
dependent. Specifically, the eff of [FeIII (bpca)2 ]NO3 ·1.67H2 O is attention has been devoted to the polymeric complex, as described
3.00 BM at room temperature, dropping to 2.20 BM at 240 K, then below.
to 1.80 BM, fully low spin, at 20 K. In contrast, the spin crossover of From a structural perspective, the polymeric complex appears
[FeIII (bpca)2 ]ClO4 ·CH3 CN is incomplete, the eff is 4.4 BM (51% HS) to be a simple extension of the discrete complex. However it has
at 290 K and gradually drops to 3.3 BM at 20 K (25% HS). a 1:1 ratio of the components (rather than the 2:1 ratio of the dis-
The crystal structure of [FeII (bpca)2 ]·H2 O (293 K, Fig. 4) crete complex) and it is crystallographically different (P21 /n vs P-1),
is typical of [MII/III (imide)2 ]0/+ complexes, with the iron(II) with a glide plane generating adjacent repeat units of the poly-
bound in two approximately perpendicular N3 pockets [12]. mer. Despite this, the average bond lengths around the high spin
The two adjacent, constricted, 5-membered chelate rings of iron(II) centre are similar (FeII –Oimide 2.052 Å and FeII –Operchlorate
the bpca anion do not fully ‘wrap around’ the iron(II) ion. 2.140 Å), to those in the discrete complex (FeII –Oimide 2.050 Å and
Hence the trans intra-ligand Npy –FeII –Npy angles are ‘pinched’ FeII –Operchlorate 2.162 Å or 2.253 Å). On average the FeIII –N bond
(163.9◦ and 165.1◦ ), rather than being close to 180◦ . The length around the low spin iron(II) centre (1.953 Å) is similar to
geometries of the predominantly low spin iron(III) complexes that in the discrete complex (Table 2). As a result of the sec-
[FeIII (bpca)2 ](NO3 )·1.67H2 O (293 K), [FeIII (bpca)2 ](ClO4 )·2CH3 CN ondary coordination of the building block complex, the imide C O
(150 K) and [FeIII (bpmca)2 ](ClO4 )·H2 O·0.5(acetone) (100 K) are bond distances lengthen from 1.208 to 1.243 Å. The chains are
likewise distorted with similar intra-ligand Npy FeIII Npy angles magnetically isolated with no intermolecular interactions observed
(163.7–164.4◦ ). The bond lengths were reported as being consis- between them.
tent with low spin iron in all of these structures (Table 2). In all The polymeric chain complex, with alternating low spin iron(III)
cases the Fe–Nimide bonds are slightly shorter than the Fe–Npyridine centres and high spin iron(II) centres, displays the properties of a
bonds. The two ligand planes, and hence also the two planes com- single chain magnet (SCM). It was reported in 2005 as being unique
prising the imide nitrogen and two imide oxygen atoms of each because the SCM properties arise from a twisted arrangement of
ligand, are approximately perpendicular to one another. This has easy plane axes along the chain axis, producing a novel type of one-
significant consequences for the magnetic pathways in complexes dimensional Ising system with easy axis anisotropy over the whole
where such mononuclear units are bridged by other metal centres chain which is responsible for the slow reversal of magnetisation,
through the imide oxygen atoms (see later). rather than the typical bistability of the constituent ions [42,45].
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2949

Table 2
Structurally characterised complexes containing the [FeII/III (imide)2 ]0/+ moiety (NB. only the first entry contains Fe in the +2 state within this moiety): selected bond lengths
(Å) and angles (◦ ) pertaining to that particular iron centre (i.e. the one bound within this moiety) in the complex.

Complex Temperature (K) Fe–Nimide Fe–Npyridine Npy –Fe–Npy Npy –Fe–Nimide Reference
(Å) (Å) (◦ ) (◦ )

[FeII (bpca)2 ]·H2 O 293 1.918–1.929 1.929–1.957 165.1 81.7/83.4 [12]


163.9 82.5/81.4
[FeIII (bpca)2 ](NO3 )·1.67H2 O 293 1.908–1.908 1.943–1.971 163.7 80.9/83.2 [12]
164.4 82.4/82.0
[FeIII (bpca)2 ](ClO4 )·2CH3 CN 150 1.915–1.926 1.962–1.980 164.2 82.2/82.1 [23]
163.3 81.7/81.6
III
([Fe (bpmca)2 ](ClO4 )·H2 O·0.5(CH3 )CO)2 100 1.901–1.921 1.963–1.971 164.5 82.4/82.2 [20]
163.1 81.5/89.9
1.895–1.924 1.954–1.963 166.2 83.4/82.8
164.8 82.2/82.2
[FeII (ClO4 )2 (FeIII (bpca)2 )2 ](ClO4 )2 ·4CH3 NO2 213 1.869–1.944 1.960–1.980 162.3 83.8/83.8 [11]
167.5 81.0/81.2
{[Fe (ClO4 )2 (Fe (bpca)2 )](ClO4 )·2.5CH2 Cl2 ·0.5CH3 NO2 ·0.5H2 O}∞
II III
190 1.900–1.905 1.969–1.985 164.6 82.3/82.3 [11]
164.6 82.2/82.4
{[FeII (ClO4 )2 (FeIII (bpca)2 )](ClO4 )·3CH3 NO2 }∞ 200 1.907–1.916 1.967–1.984 164.4 82.5/81.7 [8,42]
164.2 82.8/81.6
{[FeII (ClO4 )2 (FeIII (bmpca))](ClO4 )·2CH3 NO2 ·H2 O}∞ 100 1.879–1.912 1.949–1.969 164.2 82.3/82.4 [8]
164.8 82.2/82.3
{[FeII (ClO4 )2 (FeIII (t Bu-bpca)2 )](ClO4 )·3.5CH3 NO2 }∞ 100 1.907–1.909 1.959–1.980 163.0 81.5/82.1 [8]
164.4 82.0/82.4
{[Fe (ClO4 )2 (Fe (Ph bpca)2 )](ClO4 )·4 EtNO2 }∞
II III
100 1.909–1.913 1.959–1.980 163.8 81.7/82.4 [8]
164.2 82.3/82.3

Fig. 5. The crystal structures of discrete [FeII (ClO4 )2 (FeIII (bpca)2 )2 ](ClO4 )2 ·4CH3 NO2 (left) and polymeric {[FeII (ClO4 )2 (FeIII (bpca)2 )](ClO4 )·2.5CH2 Cl2 ·0.5CH3 NO2 ·0.5H2 O}∞
(right), generated from the same building blocks. Solvent molecules, hydrogen atoms and uncoordinated anions have been omitted for clarity.
Source: These figures were generated from data obtained from the CCDC as published originally in reference [11].

The unique environment provided by the imide building block com- field was applied; rather, these sextets correspond to the iron(II)
plexes was the key to obtaining the SCM properties. Specifically, and iron(III) centres feeling quasi-static hyperfine fields (of 192 and
the ␲-donor character and planar arrangement of the imide oxy- 335 kOe, respectively) due to the paramagnetic relaxation (rever-
gen atoms lifts the degeneracy of the t2g orbitals, splitting them sal of spin) in the chain complex occurring at a rate slower than
into eg and b2g orbitals, quenching the orbital angular momentum that of the Mossbauer timescale (10−7 s). This result, in conjunction
of the high spin (S = 2) iron(II) ions. Consequently, there is easy- with the magnetic data, was used to estimate a blocking tempera-
plane-type anisotropy through the plane of the four imide oxygen ture of 1.3 K and activation energy barrier of 27 K. In further work,
atom donors (Fig. 5). The individual easy-plane spins are coupled Muon spectroscopy, again without an applied external field, was
to the adjacent low spin (S = 1/2) [FeIII (bpca)2 ]+ units in the chain used to directly observe the ground spin state of the SCM [46].
in a ferrimagnetic arrangement. These experiments were used to observe the growth of short-range
Mössbauer spectra collected at and above 20 K, with no spin ordering and showed that slow reversal of the spin occurred
externally applied magnetic field, show the expected doublets cor- below 6 K and at 0.3 K the FeII spin fluctuations were almost entirely
responding to the high spin iron(II) and low spin iron(III) centers suppressed.
(Fig. 6, bottom left). However, on further cooling, two six-line sig- More recent work has shown that the magnetic properties
nals emerge at 3.7 K (Fig. 6, bottom right). No external magnetic can be altered by removal and re-adsorption of nitromethane

Table 3
Blocking temperature and selected angles between specified pairs of calculated mean planes for a series of four single chain magnets [8].

Complex Fe· · ·Fe intrachain  a (◦ )  b (◦ )  c (◦ ) TB (K)a

{[FeII (ClO4 )2 [FeIII (bpca)2 ]](ClO4 )·3CH3 NO2 }∞ 5.187(2), 5.195(2) 0.5(3) 80.4(2) 84.7 2.3
{[FeII (ClO4 )2 [FeIII (bmpca)2 ]](ClO4 )·2CH3 NO2 ·H2 O}∞ 5.2155(9)–5.2346(9) 3.1(2), 3.5(2) 84.97 (6), 89.35(6) 89.1 2.3
{[FeII (ClO4 )2 [FeIII (t Bu-bpca)2 ]](ClO4 )·3.5CH3 NO2 }∞ 5.229(1)–5.256(1) 3.7(2) 80.2(1) 83.3 2.8
{[FeII (ClO4 )2 [FeIII (Ph-bpca)2 ]](ClO4 )·4 EtNO2 }∞ 5.201(2)–5.252(2) 17.6(3), 17.9(3) 84.6 (2), 84.4 (2) 86.3 2.0

 a : The two planes comprised of three consecutively numbered imide oxygen atoms around the high spin iron(II) centre e.g. O1, O2, O3 and O2, O3, O4.  b : The two
O C N C O planes around the low spin iron(III) centre.  c : The two planes defined by the O4 imide oxygen atoms around two adjacent high spin iron(II) centres.
a
Blocking temperatures were measured at 997 Hz.
2950 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Fig. 6. Properties of the SCM {[FeII (ClO4 )2 (FeIII (bpca)2 )](ClO4 )·3CH3 NO2 }∞ . Top left: The temperature dependence of eff (per dimetallic repeating unit) for a powder sample
() and (inset) for an oriented single crystal in a dc field aligned either along the chain () or perpendicular to the chain (䊉). Solid lines are the theoretical curves. Top right:
Out of phase (m  ) ac magnetic susceptibility vs temperature in a 3.0 Oe ac field oscillating at the indicated frequencies with no applied dc field. Solid lines are a guide for
the eyes. Bottom: 57 Fe Mössbauer spectra acquired at the indicated temperatures.
Source: Reprinted with permission from reference [42]. Copyright 2005 American Chemical Society.

from the crystals of this chain complex, {[(FeIII (bpca)2 )


[FeII (ClO4 )2 ](ClO4 )·3(CH3 NO2 )}∞ [44]. Thermogravimet-
ric analysis at 30 ◦ C and X-ray powder diffraction showed
that all three nitromethane solvent molecules per
[FeII (ClO4 )2 (FeIII (bpca)2 )](ClO4 ) unit can be reversibly removed
and regained (repeated for three cycles). The drying, which causes
small structural changes, has a significant effect on the magnetic
behaviour of the complex, increasing the blocking temperature
by 21% (Fig. 7). Proposed explanations for the significant effect of
drying on the magnetic properties included an alteration to the
easy axis of the SCM. As the easy axis arises from the orthogonal
arrangement of the adjacent easy planes of individual iron(II)
ions (it lies approx. along the intersection of these planes), slight
deviations from that arrangement (as could be caused by solvent
loss) will strongly affect the magnetic behaviour. Alternatively
this structural reorganisation could cause a modification to the
antiferromagnetic interaction between the (S = 2) iron(II) and
(S = 1/2) iron(III) ions within the ferrimagnetically arranged chain,
Fig. 7. The plots of the out of phase m  vs T under zero dc applied
leading to the significant changes in observed magnetic behaviour. field using 5, 25, 100, 500 and 1000 Hz frequencies for {[(FeIII (bpca)2 )
An analogous SCM, {[FeII (ClO4 )(H2 O)FeIII (bmpca)2 ](ClO4 )2 · [FeII (ClO4 )2 ]](ClO4 ).3CH3 NO2 }∞ (left) and {[(FeIII (bpca)2 )[FeII (ClO4 )2 ]](ClO4 )}∞
4CH3 NO2 }∞ , was generated from the methyl substituted complex (right).
[FeIII (bmpca)2 ](ClO4 ) (Figs. 8 and 1, Tables 2 and 3) [20]. There Source: Reprinted with permission from reference [44]. Copyright 2007 Elsevier.
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2951

Fig. 8. Crystal structures of the building block [FeIII (bmpca)2 ](ClO4 )·acetone·2H2 O (left) and SCM {[FeII (ClO4 )(H2 O)[FeIII (bmpca)2 ]](ClO4 )2 ·4CH3 NO2 }∞ (right). Note that
solvent molecules, hydrogen atoms and uncoordinated anions have been omitted for clarity.
Source: These figures were generated from data obtained from the CCDC as published originally in reference [20].

is a significant structural difference at the iron(II) centres: one


of the two coordinated perchlorate ions has been replaced by a
water molecule. However, this difference does not result in any
intermolecular contacts significant enough to compromise the SCM
behaviour. Another difference is that the asymmetric unit contains
two sets of each building block moiety, not just one set as in the
previous chain complex. The complex acts as a SCM below 3 K,
although, in contrast to the original SCM, the interaction between
the S = 2 iron(II) and S = 1/2 iron(III) centres within the chain is
noticeably weakened and the activation energy barrier (estimated
at 19 K) is 8 K lower than in the original SCM.
Recent work has extended this series of SCMs to include
t Bu-bpca and Ph-bpca derivatives (Fig. 1) [8]. The building block

[FeIII (imide2 )]ClO4 complexes have been reported, as have the


Fig. 9. The structure of the [MnII (hfac)2 FeII (bpca)2 MnII (hfac)2 ] complex. Note that
polymeric chain complexes {[FeII (ClO4 )(H2 O)[(FeIII (t Bu-bpca)2 )]]
solvent molecules, hydrogen atoms and uncoordinated anions have been omitted
(ClO4 )2 ·(CH3 NO2 )3.5 }∞ and {[FeII (ClO4 )(H2 O)]0.5 [FeII (H2 O)2 ]0.5 for clarity.
[FeIII (Ph-bpca)2 ](ClO4 )2.5 ·(CH3 CH2 NO2 )}∞ (Tables 2 and 3). These Source: This figure was generated from data obtained from the CCDC as published
complexes were prepared in an analogous manner to those pre- originally in reference [47].
viously reported and also display SCM behaviour. The distortions
caused by the various substituents are reflected in the blocking being diamagnetic iron(II) (S = 0) rather than paramagnetic iron(III)
temperatures of this series (H, 2.3 K; CH3 , 2.3 K; t Bu, 2.8 K; Ph, (S = 1/2).
2 K; Table 3). The larger substituents generated greater Fe· · ·Fe
interchain separations, specifically 10.011(2), 9.3555(1), 10.686 4.3. Homometallic complexes derived from [FeIII (imide)Xn ]
(2) and 11.774(3) Å for the chain complexes produced with the building blocks
bpca, bmpca, t Bu-bpca and Ph-bpca imide ligands, respectively.
The t Bu-bpca chain complex can lose and regain CH3 NO2 solvent, Building block complexes of the formula [FeIII (imide)Xn ]
retaining crystallinity during solvent loss. However, unlike for the include [FeIII (bpca)Cl2 (H2 O)], [FeIII (bpca)Cl2 (EtOH)],
analogous complex of the bpca anion, the process is not completely [Fe (bpca)Cl2 (CH3 OH)] and [Fe (bpca)(CN)3 ]− [12,48,49]. These
III III
reversible. The Ph-bpca chain complex became amorphous upon complexes are all 6-coordinate with the bpca anion coordinated
drying and lost SCM behaviour. This was the first example of an in the N3 fashion and the remaining binding sites occupied by
SCM transitioning between super-paramagnetic and paramagnetic Cl− and solvents or by CN− . The chloride or cyanide co-ligands
states. The study concluded that substitution at the 4-position of result in high and low-spin electronic configurations, respectively
the pyridine ring does not appear to affect the magnetic behaviour (Table 4).
of these chain complexes by electronic effects, but rather by the The unit [FeIII (bpca)Cl2 (EtOH)] [48] has been used to sequen-
structural alterations that result from the substitution (Table 3). tially generate an anionic tri-iron(III) complex, crystallised as the
The [FeII (bpca)2 ] unit has also been used to bridge two hexa-iron(III) Na2 [FeIII 3 (␮3 -O)(bpca)2 Cl4 (␮2 -OEt)2 ]2 (Fig. 10), then
[M (hfac)2 ] units, where MII = Fe or Mn and hfac = hexafluoroac-
II
a mixed valent hepta-iron complex {FeII [FeIII 3 (␮3 -O)(bpca)2 Cl4 -
etylacetonate [47]. In the resulting discrete trimetallic (␮2 -OEt)2 ](EtOH)2 } [50]. The tri-iron(III) subunit consists of two
[MII (hfac)2 FeII (bpca)2 MII (hfac)2 ] complexes (Fig. 9), the cen- [FeIII (bpca)Cl] units connected via a [FeIII 3 (␮3 -O)(␮2 -OEt)2 Cl2 ]
tral [FeII (bpca)2 ] unit bridges the terminal metal ions via the moiety by two ␮2 -bridging ethoxide oxygen atoms and a ␮3 -
imide oxygen atoms and contains a low spin iron(II) ion. In both bridging oxide. Sodium ions then link two tri-iron(III) subunits
complexes the outer iron(II) or manganese(II) centres are in the via the imide oxygen atoms, forming the hexa-iron(III) struc-
high spin state and these outer metal ions are magnetically inde- ture, Na2 [FeIII 3 (␮3 -O)(bpca)2 Cl4 (EtO)2 ]2 . Spin frustration occurs
pendent of one another at room temperature. Zero field splitting between the three iron(III) centres in each tri-iron subunit, leading
of the iron(II) centres occurs below 50 K. The lack of magnetic to an overall spin of S = 3/2 per tri-iron(III) cluster.
communication, when compared to the SCM’s discussed above, The mixed valent hepta-iron cluster is made up of two tri-
could be attributed to the bridging low spin iron imide complex iron(III) clusters connected through an iron(II) centre. The central,
2952 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Table 4
A selection of bond angles and distances for the FeIII (bpca) moiety in [FeIII (bpca)Xn (solvent)]0/3− and [FeIII (bpca)2 ]X complexes.

Complex Temperature (K) Npy –FeIII –Npy (◦ ) Average FeIII –Npy (Å) Reference
III
[Fe (bpca)Cl2 (H2 O)] 208 154.2 2.120 [12]
[FeIII (bpca)Cl2 (EtOH)] 193 154.0 2.123 [48]
[FeIII (bpca)Cl2 (CH3 OH)] 223 154.0 2.119 [49]
[FeIII (bpca)(CN)3 ]PPh4 295 165.7 1.959 [49]
[FeIII (bpca)2 ]NO3 293 163.7, 164.4 1.958 [12]
[FeIII (bpca)2 ]ClO4 143 163.3, 164.2 1.969 [23]
Na2 [FeIII 3 (␮3 -O)(bpca)2 Cl4 (EtO)2 ]2 200 153.5, 153.0 2.137 [50]
[FeII (FeIII 3 (␮3 -O)(bpca)2 Cl4 (EtO)2 )(EtOH)2 ] 193 150.0, 155.4 2.131 [50]

Fig. 10. The sequential formation of [FeIII (bpca)Cl2 (EtOH)] (top left), then the anionic tri-iron(III) cluster, crystallised as Na2 [FeIII 3 (␮3 –O)(bpca)2 Cl4 (EtO)2 ]2 (top right), then
the mixed valent hepta-iron complex [FeII (FeIII 3 (␮3 -O)(bpca)2 Cl4 (EtO)2 )(EtOH)2 ] (bottom). Solvent molecules, hydrogen atoms and uncoordinated anions have been omitted
for clarity.
Source: These figures were generated from data obtained from the CCDC as published originally in references [48] and [50].

high spin iron(II) centre, Fe(4), is in an octahedral O6 environment (diiodoethylenedithotetrathiavalene) and DIEDO (diiodoethylene-
made up of four imide oxygen atoms in a square plane and two dioxotetrathiavalene) [51]. Crystal structures of the salts reveal
axial ethanol molecules (FeII O = 2.090(3)–2.113(4) Å). The ethanol that the organic and inorganic units are packed in alternate layers.
molecules make hydrogen bond interactions with the adjacent The layers are connected by interactions between the iodine atoms
uncoordinated imide oxygen atoms (O3 and O4). Experimentally of DIET and DIEDO and the oxygen atoms of coordinated bpca
the hepta-iron complex has a ground state spin of S = 12/2, higher anions and nitrogen atoms of the coordinated cyanide (Fig. 11).
than that expected from ferromagnetic coupling of two tri-iron(III) The salt formed with DIET acted as a semi-conductor at room tem-
units (S = 3/2) and a high spin iron(II) centre (S = 4/2) (STotal = 10/2). perature. The salt formed with DIEDO is weakly metallic at room
As there are no intermolecular interactions observed in the crystal temperature, with large antiferromagnetic coupling between the
structure it was concluded that the magnetic interaction responsi- radial DIEDO units, but it becomes a semiconductor below 270 K.
ble for generating this unusually high ground spin state lies within
the hepta-iron complex, perhaps due to a change in the ground- 4.4. Heterometallic complexes of the anionic [FeIII (bpca)(CN)3 ]
state spin within the tri-iron(III) subunit. building block
The anionic building block [FeIII (bpca)(CN)3 ]− has a slightly
distorted octahedral geometry around the low spin iron(III) 4.4.1. Overview of heterometallic complexes of the anionic
centre and was first crystallised with a PPh4 + counterion [49]. [FeIII (bpca)(CN)3 ] building block
The complex has been used to produce the charge transfer The mono-anionic [FeIII (bpca)(CN)3 ] building block, described
salts (DIET)2 [FeIII (bpca)(CN)3 ] and (DIEDO)2 [FeIII (bpca)(CN)3 ] in Section 4.3, has been combined with metal ion ‘linkers’ to pro-
(Fig. 11) by co-crystallisation with the radical organic units DIET duce a variety of discrete and polymeric networks. The coordinated
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2953

Fig. 11. The alternate organic and inorganic layers of the semiconductor formed from [FeIII (bpca)(CN)3 ] anions and DIET radicals, highlighting the connections between them
(dotted lines). Hydrogen atoms have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [51].

cyanides provide ‘spare’ nitrogen donors that can potentially bind The polymeric structure of {[FeIII (bpca)(CN)3 MnII (H2 O)3 ]-
to other metal ions, and, when combined with the propensity of [FeIII (bpca)(CN)3 ]·3H2 O}∞ is shown in Fig. 13 [49]. The
imide ligands to promote secondary interactions, this has led to a [FeIII (bpca)(CN)3 MnII (H2 O)3 + ]∞ cation repeats, by a centre of
variety of polymeric and mixed metal structural motifs. DFT cal- inversion, forming a double-stranded, cyanide-bridged, cationic
culations have indicated that the unpaired electron density on the chain, which is connected to the [FeIII (bpca)(CN)3 ] anions via a
iron(III) centre is held in a dxy orbital defined by the plane of the complex hydrogen bonding network involving the solvent water
three mer-coordinated cyanide ligands [49]. The structures and molecules. All of the iron(III) centres are low spin, and DFT calcula-
properties of these motifs are described below. tions suggest a combination of minor ferromagnetic and dominant
antiferromagnetic coupling between the low spin iron(III) and
4.4.2. Manganese high spin manganese(II) centres. Lack of an appropriate model
Combination of the [(FeIII (bpca)(CN)3 )] anion with man- for describing heterometallic chain complexes prevented a more
ganese(II) was done with the goal of making a one-dimensional detailed evaluation of the magnetic coupling.
ferrimagnetic chain of the low spin iron(III) (S = 1/2) and high Two one-dimensional zig-zag chain complexes,
spin manganese(II) (S = 5/2) centres. The discrete trimetallic com- [(MnIII (Salcy))(FeIII (bpca)(CN)3 )]∞ were prepared using
plex {[FeIII (bpca)(CN)3 ]2 MnII (CH3 OH)2 (H2 O)2 }·2H2 O (Fig. 12) was the (R, R) and (S, S) enantiomers of Salcy (N,N -(1,2-
obtained from slow evaporation of a water/methanol solution of cyclohexanediylethylene)bis(salicylideneiminato) dianion)
two equivalents of [(FeIII (bpca)(CN)3 )]Bu4 N and one equivalent of (Fig. 14) [53]. The repeat units of the chain are generated by
manganese(II) chloride tetrahydrate [52]. The manganese(II) centre translation along the a axis. Again, it is the two axial cyanide lig-
bridges the iron(III) building blocks via trans cyanide interactions, ands, not the imide oxygen atoms, of the anionic [FeIII (bpca)(CN)3 ]
rather than via the imide oxygen atoms. The asymmetric unit con- which bridge the [MnIII (Salcy)] cations, via the axial sites of the
tains two molecules, one with cis-coordinated water molecules manganese(III) ion (all four equatorial sites are occupied by the
and cis-coordinated methanol molecules (Fig. 12), the other has Salcy dianion). Both iron(III) and manganese(III) coordination
these trans-coordinated (Fig. 12). The magnetic data show that spheres are distorted octahedral. As expected, the manganese(III)
the iron(III) centres remain low spin following coordination to the centres present a Jahn–Teller distortion (TRR = 0.838; TSS = 0.830)
manganese(II) centre. The magnetic behaviour was modelled as [54]. Both enantiomeric forms display ferroelectric hysteresis, i.e.
antiferromagnetic coupling between the high spin manganese(II) hysteretic altering of electric polarisation in response to an applied
centre and each of the two low spin iron(III) centres (J = –3.28 cm−1 ) coercive field. Polarisation of the local distortions, which would
with no coupling occurring between the two terminal iron(III) cen- interfere in electronic communication along the chain, is proposed
tres H = −2JSMn (SFe1 + SFe2 ). as the reason for this observation.

Fig. 12. The structure of one of the two molecules in the asymmetric unit of the discrete trimetallic complex [(FeIII (bpca)(CN)3 )2 MnII (CH3 OH)2 (H2 O)2 ]·2H2 O. Note that the
secondary coordination occurs through the coordinated cyanides, rather than the imide oxygen atoms. Solvent molecules and hydrogen atoms have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [52].
2954 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Fig. 13. The structure of the cation of the infinite, heterometallic, double-stranded chain complex {[FeIII (bpca)(CN)3 MnII (H2 O)3 ][FeIII (bpca)(CN)3 ]· 3H2 O}∞ . Hydrogen atoms,
solvent molecules and [FeIII (bpca)(CN)3 ] anions have been excluded for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [49].

Fig. 14. The one-dimensional ‘zigzag’ chain complex (S, S)-{[MnIII (Salcy)][FeIII (bpca)(CN)3 ]}∞ . Solvent molecules and hydrogen atoms have been excluded for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [53].

4.4.3. Copper Attempts to model the magnetism as either dimers or chains did not
A bridging unit containing copper(II) has been used to pre- accurately account for the antiferromagnetic magnetic behaviour
pare the complex {[FeIII (bpca)(CN)3 ][CuII (H2 O)2 (CN)2 ]·1.5H2 O}∞ observed for the complex.
[52]. The [FeIII (bpca)(CN)3 ]− anion links the copper(II) centres
through an equatorial and an axial cyanide and the structure was 5. Copper complexes of imides
described as a one-dimensional chain. The shortest intramolecular
distances for Cu· · ·Cu, Fe· · ·Cu and Fe· · ·Fe are 7.263, 4.964, 6.702 Å, 5.1. Overview of copper complexes
respectively and the shortest intermolecular distances are 11.214,
7.205 and 7.938 Å, respectively. For the modelling of the tempera- The imide coordination chemistry of copper is dominated
ture dependent and field dependent magnetic measurements, the by copper(II) complexes of Hbpca. The mononuclear complex
system was treated as a uniform CuII /FeIII chain: ferromagnetic [CuII (bpca)2 ] has been reported, however unlike the analogous
intra- and interdimeric exchange constants, 7.9(3) and 1.03(2) cm−1 [FeII/III (bpca)2 ]0/+ it has not been used to produce a variety of coor-
respectively, were obtained. dination polymer materials. Instead, copper(II) imide complexes
A right-handed helical chain {[FeIII (bpca)(CN)3 ][CuII (bpca) mostly feature only one bound imide anion per copper centre, and
(H2 O)]·H2 O}∞ has been reported (Fig. 15) [52]. Surprisingly, circu- take two main forms: discrete mononuclear [CuII (bpca)X], where
lar dichroism measurements on 40 individual crystals from each of X− does not bridge to another complex; and discrete dinuclear
the two batches all exhibited a negative Cotton effect at max = 450 [CuII (bpca)XCuII (bpca)], where X2− bridges the two copper(II) cen-
and 625 nm, indicating an enantiomeric excess despite starting tres to form a dimer. Complexes of the form [CuII (bpca)X] have also
from achiral reagents. The chain structure is constructed from been used as building blocks to produce a variety of discrete and
[FeIII (bpca)(CN)3 ] anions and [CuII (bpca)(H2 O)] cations connected polymeric heterometallic complexes. As a consequence of the use
into dimers via a cyanide bridge (Cu(1) N6 1.953 Å) in the plane of copper(II) in imide synthesis (Section 1), imide complexes of cop-
of both bpca− ligands. A second, weaker interaction (Cu(1A)· · ·N5 per display the most variety in ligand design. Indeed many Hbpca
2.632 Å) connects these dimers into the helical chain structure. ligand variants are only reported as complexes of copper.
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2955

appropriate bridging metal ion, would be excellent building blocks


for facilitating electronic communication, with structures like those
exemplified above in the iron section of this review (Section 4.2).
However, the complex [CuII (bpca)2 ] has only been used as a build-
ing block for the construction of one extended complex (Section
5.3) [47]. This may be due to the combined effects of the Jahn–Teller
distortion on one of the fully conjugated bpca anions and the labil-
ity of copper(II) leading to disassembly of the [CuII (bpca)2 ] unit
in solution, preventing predictable and/or controlled construction
of larger assemblies. Another potential pitfall is that, as copper(II)
gains little crystal field stabilisation energy (CFSE) from being in an
octahedral geometry, during such secondary assembly reactions
transmetalation of copper(II) by metal ions such as iron(II) and
iron(III) can occur, especially as these are likely to have a low spin
electronic configuration and hence even greater CFSE [47].

5.3. [CuII (bpca)2 ] used as a building block

The only example of the [CuII (bpca)2 ] unit being used as a


building block is the complex [MnII (hfac)2 CuII (bpca)2 MnII (hfac)2 ],
which is isostructural to the iron(II) complex
[MnII (hfac)2 FeII (bpca)2 MnII (hfac)2 ] described earlier, in Sec-
tion 4.2 (Fig. 9) [47]. The copper(II) ion in the N6 pocket has an
average bond length of 2.134 Å compared to Fe–N = 1.958 Å in the
Fig. 15. The structure of the chiral chain complex
[(FeIII (bpca)(CN)3 )(CuII (bpca)(H2 O))·H2 O]∞ , oriented such that the chain runs low spin iron(II) analogue (Fig. 9). Unlike the rest of the man-
into and out of the page. The chain is made up of [Fe (bpca)(CN)3 ]− and
III ganese(II) and nickel(II) analogues (Sections 6.3 and 6.5), there was
[CuII (bpca)(H2 O)]+ dimers connected by a FeIII –CN–CuII bond which repeats by no interaction between the CuII and MnII centres. For the reasons
symmetry (dashed lines; Cu(1A)· · ·N(5) 2.632 Å). Solvent molecules and hydrogen discussed above (see Section 5.2), attempts to use [CuII (bpca)2 ] to
atoms have been excluded for clarity.
bridge [FeII (hfac)2 ] units led to transmetallation, with a low spin
Source: This figure was generated from data obtained from the CCDC as published
originally in reference [52].
iron(II) ion replacing the copper(II) ion and forming the all-iron(II)
imide complex.

5.2. The discrete complex [CuII (bpca)2 ]


5.4. CuII (imide)X·solvents complexes

The structure of [CuII (bpca)2 ]·H2 O (Fig. 16) [55] is similar to that
Many complexes of the formula [CuII (bpca)X] have resulted
of [FeII/III (bpca)2 ]0/+ , but it differs in that the geometry around the
from the hydrolysis of TPyT (Fig. 3), in the presence of differ-
six coordinate copper(II) ion is significantly distorted away from
ent co-ligands. However deliberate synthesis from the isolated
octahedral due to the Jahn–Teller effect. The Cu–N bond lengths
imide ligand offers the opportunity to choose the co-ligands that
are longer along the z axis (CuII –Npy = 2.29–2.30 Å) compared to the
will occupy the remaining coordination sites of the three coor-
x and y axes (CuII –Npy = 2.08, 2.09 Å). Hence, unlike in the iron(II)
dinate copper(II) ion in the [CuII (bpca)]+ moiety. The formulae
and iron(III) analogues (Table 2), one ligand is very different to
and selected bond distances and angles of the mononuclear com-
the other. EPR spectroscopy showed that a considerable amount
plexes reported to date are summarised in Table 5. In all cases
of the unpaired electron spin was transferred onto the ligands.
the bpca anion is meridionally bound through the N3 donor set.
This suggests that [CuII (bpca)2 ] units, in combination with an
The significant deviations from ideal square-based geometries, as
exemplified here by the trans Npy -Cu-Npy angles of 153.0–164.8◦
are due to the small bite angle (77.0–82.6◦ ) of the two adjacent
5-membered chelate rings imposed by the bpca anion (Table 5).
The coordination numbers and geometries of the copper(II) centres
in these monometallic [CuII (imide)X] complexes vary: 4 coordi-
nate are distorted square planar, 5 coordinate are distorted square
based pyramids, and 6 coordinate are distorted octahedra. The
examples selected for discussion are: [CuII (bpca)(tptz)](CF3 SO3 )
[56], [CuII (bpca)(Cl)(NCO)] [27] and [CuII (bpca)(H2 O)2 ]NO3 [57]
(Fig. 17). These include uncoordinated anions, coordinated anions
and a hydrogen bonded network, respectively.
As the bpca anion is fully conjugated it is expected to remain
planar when coordinated. However, in all of these [CuII (bpca)]+
complexes the bpca anion is better described as being “hinged”
along the CuII Nimide bond, comprising two planar ‘halves’ which
intersect at an angle of 6–8◦ [58,59]. Each half is made up of one
pyridine ring, the adjacent carbonyl group and the imide nitrogen
atom. When the ligand is considered as being made up of two indi-
vidual planes the atoms within the two halves of the ligand deviate
Fig. 16. A ball and stick representation of [CuII (bpca)2 ]·H2 O. Note that solvent
molecules and hydrogen atoms have been excluded for clarity. by, at most, 0.012 Å from their respective planes, whereas when
Source: This figure was generated from data obtained from the CCDC as published the plane is comprised of the whole ligand, several atoms deviate
originally in reference [55]. significantly from that plane (by, at most, 0.689 Å [59]).
2956 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Table 5
A list of monometallic [CuII (imide)X] complexes containing the [CuII (bpca)]+ or [CuII (bpcam)]+ moiety and selected bond lengths and angles pertaining to the coordination
of the imide ligand around the copper(II) ion.

Complex CuII –Nimide (Å) CuII –Npy (Å) Npy –CuII –Npy (◦ )a Npy –CuII –Nimide (◦ )a Reference
II b,c
[Cu (bpca)2 ]·H2 O 1.960, 1.997 2.084–2.305 160, 153 80/80, 77/78 [55]
[CuII (bpca)(C5 NH5 CONH2 )(ClO4 )] 1.931 2.018/2.017 162.0 82.0/82.1 [56]
[CuII (bpca)(C5 NH5 CONH2 )(CF3 SO3 )] 1.935 2.016/2.022 163.5 81.5/82.4 [59]
[CuII (bpca)(TpyT)](CF3 SO3 )c 1.956 2.047/2.055 162.8 81.3/81.5 [59]
[CuII (bpca)(H2 O)2 ](NO3 ) 1.938 1.994/1.996 163.8 81.87/81.98 [57]
[CuII (bpca)(Cl)(NCO)] 1.952 2.022/2.026 161.9 80.7/81.5 [27]
[CuII (bpca)(Br)]d 1.928 2.011/2.011 164.2 82.1/82.1 [18]
[CuII (bpca)(I)]d 1.936 2.020/2.020 164.5 82.5/82.5 [18]
[CuII (bpca)(2Cl C6 H4 CO2 )(H2 O)] 1.947 2.012/2.014 161.9 81.5/81.7 [61]
[CuII (bpca)(3Cl C6 H4 CO2 )]d 1.928 2.000/2.002 164.8 82.3/82.5 [61]
[CuII (bpca)(4Cl C6 H4 CO2 )(DMF)] 1.950 2.012/2.018 163.3 81.89/81.65 [61]
[CuII (bpca)(C5 NH5 CO2 )]·2(H2 O) 1.952 2.046/2.047 159.9 80.87/80.94 [62]
[CuII (bpca)(CH3 CO2 )(H2 O)2 ]·H2 Oc 1.947 1.998/2.000 162.0 81.6/82.3 [60]
[CuII (bpca)(NCS)(H2 O)] 1.94 2.00/2.01 162.0 81.5/81.8 [58]
[CuII (bpcam)(H2 O)3 ]NO3 c 1.938 2.002/2.020 163.5 81.57/82.01 [2]
[CuII (bpcam)(H2 O)2 ]ClO4 ·3H2 O 1.940 2.011/2.011 162.6 81.77/81.85 [63]
[CuII (bpcam)(tcm)(H2 O)] 1.935 2.035/2.036 162.6 81.20/81.60 [64]
[CuII (bpcam)(tcm)(H2 O)]·2H2 O 1.937 2.019/2.027 161.9 81.47/81.60 [64]
[CuII (bpcam)(CN)(H2 O)] 1.950 2.036/2.036 160.4 80.51/80.78 [65]
[CuII (bpcam)(N3 )(H2 O)] 1.954 2.031/2.027 160.2 80.70/81.30 [65]
[CuII (bpcam)(NCS)(H2 O)] 1.936 2.019/2.028 161.1 81.03/81.39 [65]
[CuII (bpcam)(dca)(H2 O)] 1.941 2.024/2.045 161.4 81.09/81.12 [65,66]
[CuII (bpcam)(ma)] 1.931 1.999/2.005 163.3 82.6/82.0 [67]
[CuII (bpcam)(mpa)(H2 O)]·0.5H2 O·0.5CH3 OH 1.933 2.009/2.014 163.5 81.7/81.75 [67]

Unless otherwise noted the geometry of the copper(II) ion is square pyramidal.
a
Intraligand.
b
Evident Jahn–Teller distortion.
c
Six-coordinate complex.
d
Four-coordinate complex.

The crystal packing of these complexes often shows head to molecules (Fig. 17) [58,59,61,62]. An unusual feature is present
tail carbonyl–carbonyl interactions [60], stacking of pyridine rings in the crystal packing of [CuII (bpca)(H2 O)2 ](NO3 )·2H2 O [57] and
[27,61] and extensive hydrogen bonding networks between the [CuII (bpca)(3Cl C6 H4 CO2 )] [61]. In [CuII (bpca)(H2 O)2 ](NO3 )·2H2 O,
imide oxygen atoms and coordinated and uncoordinated solvent in the sixth potential coordination site of the copper(II) centre, a

Fig. 17. The structures of [CuII (bpca)(tptz)](CF3 SO3 ) (left), [CuII (bpca)(H2 O)2 ]NO3 ·2(H2 O) (right) and [CuII (bpca)(Cl)(NCO)] (bottom). Anions and non-acidic hydrogen atoms
have been omitted for clarity.
Source: These figures were generated from data obtained from the CCDC as published originally in references [27], [57] and [59].
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2957

Fig. 19. The structure of [CuII 2 (bpca)2 (H2 O)2 (SO4 )]·H2 O where the sulphate anion
forms a three atom bridge between the copper(II) centres. Hydrogen atoms not
involved in hydrogen bonding (dashed lines) have been excluded for clarity.
Fig. 18. The structure of [CuII (bpca)(3Cl C6 H4 CO2 )] depicting a weak side-on imide
Source: This figure was generated from data obtained from the CCDC as published
C O· · ·CuII interaction between adjacent complexes. Hydrogen atoms have been
originally in reference [63].
omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published
originally in reference [61].
for the coordinated bpca anion (6–8◦ ). The complexes have the
same propensity to form extended hydrogen bonded networks
through the imide oxygen atoms; interestingly the uncoordinated
very weak side-on interaction (3.302 Å) occurs with the imide oxy-
pyrimidine nitrogen atoms are not involved.
gen atom of an adjacent complex. In [CuII (bpca)(3Cl C6 H4 CO2 )]
(Fig. 18) a similar side-on carbonyl copper(II) interaction occurs
[Cu(1) O(2A) = 2.7189 Å], with the oxygen atom occupying the api- 5.5. Discrete {[CuII (imide)]2 X} complexes
cal site of a square based pyramid. While the imide oxygen atoms
are commonly observed to coordinate to a single metal centre in The synthesis of discrete dinuclear {[CuII (imide)]2 X} complexes
an in-plane bidentate fashion, these monodentate side-on interac- has been performed both directly from a mixture containing cop-
tions provide another type of weak coordination mode for imide per(II) nitrate and TPyT (Fig. 1) and from the labile building block
type ligands. Another notable feature of these side-on CuII · · ·Oimide [CuII (imide)(H2 O)2 ]NO3 , by the addition of a divalent anion salt.
interactions is the deviation of the imide oxygen atom, i.e. O(2) in The coordinated water can be substituted by the bridging dian-
Fig. 18, out of the plane of the adjacent pyridine ring towards the ion X2− under very mild reaction conditions, i.e. short periods
copper(II) centre, Cu(1B), by 0.391 Å in this case. of stirring at room temperature. Therefore the majority of these
UV–vis studies have been used to show that the complex bridged complexes have been prepared by either slow evapora-
[CuII (bpca)(C5 NH5 CONH2 )(ClO4 )] becomes [CuII (bpca)(H2 O)3 ]+ in tion of a reaction solution or slow diffusion together of solutions
solution. The lability of the co-ligands in copper(II)/imide com- containing the reactants. Particular attention has been paid to mag-
plexes is an important consideration when designing molecular netostructural relationships in the resulting dimetallic complexes:
building blocks for higher order structures [56]. With regard to the the [CuII (imide)]+ moiety is a useful cationic building block to
stability of these complexes, studies examining the acid–base equi- examine the nature of the magnetic pathways provided by the var-
libria of the complex [CuII (bpca)(H2 O)2 ](NO3 ) concluded that the ious dianionic linkers. In all cases the unpaired electron density is
compound remained intact in the pH range 3–7, however outside determined to be lying in an equatorial dx2 −y2 metal centred orbital
this pH range either protonation or deprotonation of the complex where the x and y axes (the equatorial plane of the copper(II) cen-
occurs. Comparison of the UV–vis spectra at pH values of 1.5 and 7 tre) are approximately defined by the imide and aromatic nitrogen
showed that the [CuII (bpca)]+ unit remained intact so protonation atoms of the imide ligand. In all but three of the complexes exam-
of the complex did not result in the imide ligand dissociating [57]. ined here (Figs. 20 and 26), the dianion provides the fourth donor
Electrochemical studies of the complex in this plane.
[CuII (bpca)(2Cl C6 H4 CO2 )(H2 O)] showed an irreversible reduc- Tetrahedral dianions bridge the [CuII (imide)]+ units in
tion, at −0.78 V in acetonitrile or −0.434 V in dimethylformamide, two ways, either via three atoms (i.e. CuII –O–A–O–CuII ),
vs NHE [61]. This was assigned as CuII → CuI , and the irreversibility as in the cases of [CuII 2 (bpca)2 (H2 O)2 (AO4 )]·H2 O, A = W
was attributed to the CuI complex being unstable, resulting in [69], Mo [69], and [CuII 2 (bpcam)2 (H2 O)2 (AO4 )]·H2 O, A = S
decomposition. The highly negative reduction potential showed (Fig. 19) [63], or via a single oxygen atom (CuII –O–CuII )
that, as expected, the copper(II) oxidation state is significantly in the cases of [CuII 2 (bpca)2 (H2 O)3 (AO4 )]·H2 O where
stabilised by coordination to the imide anion, in comparison to A = S or Cr (Fig. 20) [70]. In contrast, in the complex
free CuII (E0 Cu(II) /Cu(I) = 0.153 V) [68]. [CuII (bpcam)(H2 O)2 ][CuII (bpcam)(H2 O)(SO4 )], sulphate is
Copper complexes of Hbpcam were also originally produced observed to coordinate but not bridge [63]. In this case, the
from the hydrolysis of the corresponding triazine (Fig. 3). Indeed coordination mode obtained appears to depend on the source of
the complex [CuII (bpcam)(H2 O)3 ]NO3 was the first reported sulphate: either CuSO4 (both) or Na2 SO4 (non-bridging).
structure of an imide complex [1,2]. The structural characteristics In the complexes where the [CuII (bpca)]+ units are linked by
of Hbpcam complexes are similar to those of Hbpca complexes three atom bridges provided by tungstate, molybdate or sulphate,
(Table 5). When coordinated, the bpcam anion is also best described the copper(II) ions adopt a square pyramidal geometry with an
as two planes, however the dihedral angle between these planes oxygen atom of the anion and the imide-N3 pocket making up
(3–13◦ ) varies more from complex to complex than was the case the equatorial plane, which is capped by a coordinated water
2958 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

sulphate and chromate complexes, respectively, and the shortest


intermolecular CuII · · ·CuII distances are 6.923 and 6.971 Å, respec-
tively. The antiferromagnetic coupling propagated through the
single atom bridges is weak due to the bridging anion occupying
an equatorial site on the square pyramidal copper centre, but an
axial site on the octahedral copper centre (sulphate J = −0.6 cm−1 ;
chromate J = −1.6 cm−1 ). (H = −JS1 ·S2 ).
The organic acids oxalate, squarate and croconate have also been
used as linkers between [CuII (imide)]+ units. The pair of structurally
characterised oxalate complexes [(CuII (bpca))2 (H2 O)2 (C2 O4 )] [71]
and the dihydrate [(CuII (bpca))2 (H2 O)2 (C2 O4 )]·2H2 O [70] (Fig. 21)
feature octahedral copper(II) centres, with half of the complex
in the asymmetric unit and the other half generated by a cen-
tre of inversion. The imide oxygen atoms of the coordinated bpca
anion are heavily involved in intermolecular hydrogen bonding.
In both structures the planar bis-bidentate oxalate anion binds
to each of the two copper(II) centres via both an axial and equa-
torial position, bridging them (intramolecular CuII · · ·CuII 5.631 vs
5.422 Å for the dihydrate). In both cases the plane of the bridg-
ing oxalate anion is almost at right angles (80.6◦ vs 83.6◦ for
the hydrate) to the equatorial plane of the copper(II) ion. Hence
it is not surprising that both complexes display only a very
weak ferromagnetic interaction between the copper(II) centres,
Fig. 20. The structure of [CuII 2 (bpca)2 (H2 O)3 (CrO4 )]·H2 O. Non-acidic hydrogen
atoms have been omitted for clarity.
with the non-hydrated form displaying the slightly stronger cou-
Source: This figure was generated from data obtained from the CCDC as published
pling (J = + 1.1 vs +1.0 cm−1 ). (H = −JS1 ·S2 + S1 DS2 ). The analogous
originally in reference [70]. complex [(CuII (bpcam))2 (H2 O)2 (C2 O4 )] adopts a similar structure,
however the oxalate and copper equatorial planes are some-
what less orthogonal to one another (69.2◦ ) and the CuII · · ·CuII
molecule at the apical position (Fig. 19) [69]. The anions provide distance is slightly longer (5.677 Å). Weak ferromagnetic cou-
three atom bridges between the two copper(II) centres (CuII · · ·CuII pling (J = +0.75 cm−1 ) was also observed for this analogue [72].
distances of 5.572, 5.550 and 5.494 Å, respectively). The resulting (H = –JS1 ·S2 ).
discrete dicopper(II) complexes are further connected through a In the dimetallic complex [(CuII (bpca))2 (H2 O)2 (C4 O4 )] (Fig. 22)
hydrogen bonding network involving all available oxygen atoms the equatorial sites of the two copper centres are bridged by a pla-
except O(6). Magnetic coupling between the pair of copper(II) cen- nar squarate anion. As in the oxalate structures, only half of the
tres varies widely with anion, from weak ferromagnetic coupling complex is in the asymmetric unit and the other half is generated
for tungstate (J = 0.557 cm−1 ), to antiferromagnetic coupling for by a centre of inversion.
molybdate (J = −77.2 cm−1 ), to weak antiferromagnetic coupling for The plane of the anion bridge is close to perpendicular (85.5◦ )
sulphate (J = –7.0 cm–1 ) (H = −JS1 ·S2 ). to the equatorial plane of the copper(II) centre. The observed
The dimetallic complexes [CuII 2 (bpca)2 (H2 O)3 (SO4 )]·H2 O and CuII · · ·CuII distance is 7.833 Å, far longer than in the oxalate
[CuII 2 (bpca)2 (H2 O)3 (CrO4 )]·H2 O are isostructural [70]. Both com- complex, as expected for this longer bridge. Unsurprisingly,
prise a square pyramidal and an octahedral copper centre bridged magnetic studies show there is no coupling via the squarate
by a single oxygen atom of the oxo-anion, with the whole dicop- bridge [73]. In contrast, in the structurally characterised ana-
per(II) molecule in the asymmetric unit (Fig. 20). Again, the oxygen logue [(CuII (bpcam))2 (H2 O)4 (C4 O4 )]·10H2 O, whilst the metal ions
atoms are involved in a large number of hydrogen bonds. The are held at the same separation (7.819 Å), the copper equatorial
intramolecular CuII · · ·CuII distances are 3.747 and 3.660 Å for the plane and that of the squarate are further from being orthogonal

Fig. 21. The oxalate bridged complex [CuII 2 (bpca)2 (H2 O)2 (C2 O4 )]·2H2 O. Non-acidic hydrogen atoms have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [70].
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2959

Fig. 22. The structure of two [CuII (bpca)(H2 O)]+ units linked by a squarate anion. Non-acidic hydrogen atoms have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [73].

to one another (67.8◦ ). Weak ferromagnetic coupling constants of An attempt to use 4,4 -bipyridine (4,4 -bipy) as the
J = +0.75 cm−1 (anhydrous) and J = +1.3 cm−1 (·10 H2 O) are observed linker produced the dimetallic complex [CuII 2 (bpca)2 (4,4 -
[72]. (H = −JSA ·SB ). bipy)(H2 O)2 ](ClO4 )2 [75] in which 4,4 -bipyridine acts as a
In the complex [CuII 2 (bpca)2 (C5 O5 )]·3H2 O (Fig. 23) the entire monodentate ligand rather than a bridge. The monometallic
dimetallic complex is in the asymmetric unit [74]. Both the square units are instead linked via a weak side-on interaction (2.838 Å)
pyramidal, Cu(2), and octahedral, Cu(1), copper(II) centres again between CuII and an imide oxygen atom from an adjacent unit.
feature an N3 -coordinated imide ligand in the equatorial plane. Similar molecules reported in the same work were successfully
The planar croconate anion is triply bridging. It coordinates to linked through 4,4 -bipyridine, highlighting the lack of predictabil-
Cu(1) and Cu(2) in a bis-bidentate fashion, occupying the remain- ity when including multiple possible coordination instructions in
ing equatorial [1.961(2), 1.973(2) Å] and an axial [axial 2.477(3), a supramolecular system. The thermal behaviour of the complex,
2.585(3) Å] site on both centres, providing a four atom bridge a perchlorate salt, was studied in anticipation of a phase transition
between them [Cu(1)· · ·Cu(2) 6.824 Å]. Further, the remaining cro- occurring, however it exploded at 339 ◦ C.
conate oxygen atom coordinates to the axial site of a neighbouring The magnetic interactions of the dithiosquarate complexes
dinuclear unit, Cu(1A) [2.496(3) Å]; this bridge generates a 1D [CuII 2 (bpca)2 (1,2-dtsq)(H2 O)]·2H2 O and [CuII 2 (bpca)2 (1,3-
chain structure overall. Once again the croconate bridge is almost dtsq)]·2H2 O have been studied and analysed using DFT calculations
orthogonal [Cu(1) 81.3◦ , Cu(2) 85.5◦ ] to the equatorial planes of [76]. Two very different magnetic pathways are possible through
the copper(II) centres, and provides a pathway for weak antifer- the 1,2-dtsq dianion which bridges the two [CuII (bpca)]+ units
romagnetic coupling (J = −9.63 cm−1 ) between the two copper(II) in a non-symmetric fashion (Fig. 24). The direct single atom
centres. (H = −JS1 ·S2 ). The individual chains are linked by a weak S(1) bridge from the axial position of Cu(1) to the equato-
side-on carbonyl· · ·copper interaction [Cu(2)· · ·O(3 ) = 2.932(3) Å]. rial plane of Cu(2) is responsible for ferromagnetic coupling
The structure of the analogous Hbpcam complex was not obtained J = +34.2 cm−1 . There is also a four atom bridge, S(1) C C S(2),
due to poor X-ray diffraction by the crystals, however the complex providing another possible, but likely weak, magnetic path-
displays much weaker antiferromagnetic coupling (J = –0.45 cm−1 ) way between the equatorial positions on both Cu(1) and Cu(2)
[72]. (H = –JSA ·SB ). The authors suggest that the structural influ- (Fig. 24). A weak intermolecular bridge [Cu(1)· · ·S2A 3.207 Å]
ence of the methyl groups alters the magnetic pathway and could propagates a weak antiferromagnetic interaction (J = –5.0 cm–1 );
even have resulted in an ionic material with no covalent connection H = –J(S1 ·S2 + S1a ·S2a ) – J(S1 ·S1a ). The equatorial planes of the
between the [CuII (bpcam)]+ units. two imide units and the 1,2-dtsq anion are all approximately

Fig. 23. The dicopper(II) [Cu2 (bpca)2 (C5 O5 )]·3H2 O units are further connected by an interaction involving the fifth croconate oxygen atom [Cu(1a)–O(10) = 2.496(3) Å],
generating a 1D-chain complex overall. Note that hydrogen atoms and solvent molecules have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [74].
2960 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Fig. 24. In the structure of [CuII 2 (bpca)2 (1,2-dtsq)(H2 O)]·2H2 O two


[CuII (bpca)]+ units are bridged by 1,2-dtsq in a non-symmetric fash-
ion. Hydrogen atoms and solvent molecules have been omitted for
Fig. 26. In the structure of [CuII 2 (bpca)2 (H2 O)2 CuII (1,2-dtcr)2 ]·2H2 O two
clarity. Cu(1)–S(1) = 2.913(1), Cu(1)–S(2) = 2.280(1), Cu(2)–S(1) = 2.324(1),
[CuII (bpca)(H2 O)]+ units are bridged by one [CuII (1,2-dtcr)2 ]2− . Non-acidic
Cu(2)–O(4) = 3.241, Cu(1)–S(2a) = 3.207 Å.
hydrogen atoms and solvent molecules have been excluded for clarity.
Source: This figure was generated from data obtained from the CCDC as published
Source: This figure was generated from data obtained from the CCDC as published
originally in reference [76].
originally in reference [77].

orthogonal to one another [planeN1 ,N2 ,N3 , to planeN4 ,N5 ,N6 = dimetallic complex generates a loosely connected chain, so the
84.5◦ , planeN1 ,N2 ,N3 , to plane1,2−dtsq = 79.4◦ ; planeN4 ,N5 ,N6 , to data can also be modelled as a magnetic chain with intramolecular
plane1,2−dtsq = 85.7◦ ]. (J1 = –33.6 cm−1 ) and weaker intermolecular (J2 = –2.2 cm−1 ) anti-
In the complex produced using a 1,3-dtsq anion to bridge ferromagnetic coupling.
[CuII (bpca)]+ units, the copper centres have distorted square pyra- The dianionic copper(II) 1,2-dithiocroconate (1,2-dtcr) com-
midal geometries and only half the complex is in the asymmetric plex [CuII (1,2-dtcr)2 ]2− has been used to bridge two [CuII (imide)]
unit, with the other half generated by an inversion centre. The units, forming [CuII 2 (bpca)2 (H2 O)2 CuII (1,2-dtcr)2 ]·2H2 O (Fig. 26)
two [CuII (bpca)]+ units are connected through all four donors of [77]. The square pyramidal copper(II) centres in the two
the planar 1,3-dtsq anion. The two sulphur donors [S(1), S(1a)] [CuII (bpca)(H2 O)]+ units are related by a centre of inversion. Again
occupy the remaining equatorial site, whilst the oxygen donors the equatorial plane is occupied by the imide ligand, and in this
[O(3), O(3a)] occupy an axial site, on each of the copper centres case a water molecule. The axial sites are bridged by a three atom,
(Fig. 25) [76]. In contrast to [CuII 2 (bpca)2 (1,2-dtsq)(H2 O)]·2H2 O, S(2)-Cu(2)-S(2a), bridge between Cu(1) and Cu(1a) [Cu(1)· · ·Cu(1a)
the equatorial planes of the copper(II) centres are now parallel, 8.378 Å; Cu(1)· · ·Cu(2) 4.189 Å]. The square plane of the Cu(2) centre
but offset, due to symmetry [planeN1 ,N2 ,N3 , to planeN4 ,N5 ,N6 = 0◦ ; in the bridging dianion is almost parallel (11.4◦ ) to the equatorial
offset by 4.892 Å], and are not as close to orthogonal to the dian- plane of Cu(1) but is considerably offset from it. No coupling was
ionic bridge [[planeN1 ,N2 ,N3 , to plane1,3-dtsq = 73.4◦ ]. Despite the observed between these copper(II) centres.
relatively long Cu(1)· · ·Cu(1a) separation (7.212 Å), when modelled
as a dimer of copper(II) ions the planar 1,3-dtsq anion provides 5.6. Chain complexes built from [Cu(imide)]+ units
antiferromagnetic coupling of J1 = –33.6 cm−1 . A weak side-on
carbonyl-copper interaction [Cu(1)· · ·O(1b) = 2.90 Å] to an adjacent In contrast to the use of dianionic linkers with [CuII (imide)]+
units, which usually produces discrete complexes as discussed in
Section 5.5 [the key exception being the 1D chain formed using cro-
conate (Fig. 23)], the use of monoanionic linkers exclusively results
in the production of polymeric chain complexes.
The chains are made up of [CuII (imide)]+ units where the imide
again provides three equatorial nitrogen donors and defines the
location of the dx2 −y2 orbital which contains the unpaired electron
density [78].
The chain complex {[CuII (bpca)](ClO4 )}∞ (tetragonal, P41 ),
revealed a previously unseen structural motif for Hbpca where the
coordination sphere of the square pyramidal metal ion is com-
prised of the N3 pocket from one imide and two imide oxygen
atoms of the imide of an adjacent complex, generating the 1D
chain (Fig. 27) [56]. Within the chain, adjacent N3 planes are almost
orthogonal to one another [88.7◦ ], the Cu· · ·Cu distance is 5.32 Å and
the bpca anion mediates a antiferromagnetic interaction of J = 10 K
[79]. (H = JS1 ·S2 ). The chains are magnetically isolated [smallest
Fig. 25. In the structure of [CuII 2 (bpca)2 (1,3-dtsq)]·2H2 O two [CuII (bpca)]+ units intermolecular Cu· · ·Cu 8.07 Å]. The absence of ␲-stacking inter-
are bridged by 1,3-dtsq. A weak interaction (Cu(1)· · ·O3 = 2.590 Å) has been shown
actions is somewhat surprising given the availability of aromatic
as dashed lines. Hydrogen atoms and solvent molecules have been excluded for
clarity. rings. The chain was crystallised from water but is not hydrated.
Source: This figure was generated from data obtained from the CCDC as published Nevertheless it was shown, by UV–vis studies, to dissociate into
originally in reference [76]. [CuII (bpca)(H2 O)2 ]+ in aqueous solution [56].
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2961

Fig. 27. The imide based chain complex {[CuII (bpca)](ClO4 )}∞ . Hydrogen atoms and uncoordinated anions have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [56].

The tricyanomethanide (tcm) anion has, in addition to the dis- (Fig. 29). The result is a previously unseen binding mode, with only
crete complexes reported in Section 5.5, been used to generate two of the nitrogen donors of the N3 imide set binding to one square
the chain complexes [CuII (tcm)(bpca)]∞ and [CuII (tcm)(bpcam)]∞ pyramidal copper(II) centre, whilst the third nitrogen donor of this
[64,65]. The copper(II) centres of these complexes have a square set, along with the second of these two bipy donor atoms, binds to
pyramidal geometry which is composed of the bpca anion and two the second square pyramidal copper(II) centre. In order to achieve
tcm ligands (Fig. 28). Each anionic tcm ligand acts as a 1,5-bridge this, the imide group is twisted out of coplanarity to an uncharac-
between the axial and equatorial sites on adjacent copper centres, teristically large degree: the two individual N C O planes of the
generating a zig-zag structural motif (Fig. 28). The equatorial-axial imide form an angle of 46.12◦ . The resulting helical structure is
nature combined with the length of the anionic tcm bridge results stabilised by face-to-face ␲-stacking interactions.
in very weak antiferromagnetic coupling between the copper(II)
centres [J = −0.64 and −0.3 cm−1 for the bpca and bpcam chain 5.8. Heterometallic complexes of [Cu(imide)]+ units
complexes, respectively] (H = − J˙ i Si · Si+1 ).
Complexes of the formula [CuII (bpca)X]∞ , where X = CN− , N3 − , Discrete square tetrametallic complexes have been pro-
Cl− , Br− or benzilate, form 1D chain structures [67,79,80]. These duced by combining [CuII (imide)]+ units with the anionic
chains are composed of square planar [CuII (bpca)X] building blocks linkers [FeIII (CN)6 ]3− [82], [FeIII (CN)3 (Tp)]− (Tp = tris(pyrazolyl)
which are bridged through an axial site by the X of an adjacent hydroborate) [83] and [FeIII (5,5 -dmbipy)(CN)4 ]− (5,5 -
molecule. The axial-equatorial nature of the interaction again leads dmbipy = 5,5 -dimethylpyridine) [84].
to weak antiferromagnetic coupling between the square pyramidal In the two rectangular complexes [CuII 2 (bpcam)2 FeIII 2 -
copper(II) centres, with the exception of X = CN− which shows no (CN)6 ](PPh4 )4 ·4H2 O and [CuII 2 (bpca)2 FeIII 2 (CN)6 ](AsPh4 )4 (Fig. 30,
interaction at all. Table 6), by inversion symmetry the two square pyramidal
copper(II) centres and two octahedral iron(III) centres alternate
5.7. Copper complexes with unusual imides around the rectangle (one of each is present in the asymmet-
ric unit) [82]. Each [FeIII (CN)3 ]3− unit links unequally to an
Copper(II) promoted reactions have resulted in the unexpected equatorial and an axial site on the adjacent [CuII (imide)]+ units
formation of several imide complexes, including quinoline contain- via the nitrogen atoms of two cis cyanide ligands (Table 6).
ing [CuII (bqca)(H2 O)(OAc)] and [CuII (pqca)(H2 O)(OAc)] [22] and Due to the axial-equatorial bridging of the [CuII (imide)]+
2,2 -bipyridine containing [CuII 2 (bbpca)2 ](PF6 )2 (Fig. 1) [81]. The units, the complex [CuII 2 (bpcam)2 FeIII 2 (CN)6 ](PPh4 )4 is
ligands Hbqca and Hpqca have been isolated by removal of cop- modelled as a rectangle rather than a square. Hence the
per(II) using EDTA4− (Fig. 3) [22]. In a separate study, the complex magnetic interactions were modelled using the Hamiltonian
[CuII 2 (bbpca)2 ](PF6 )2 ·EtOH·2CH3 CN resulted from hydrolysis of a H = −J1 ·[SFe(1) ·SCu(1) + SFe(1a) ·SCu(1a) ] − J2 ·[SFe(1) ·SCu(1a) + SFe(1a) ·SCu(1) ].
triazine complex [81]. The Hbbpca ligand has not been isolated. This results in different strengths for the ferromagnetic coupling
The complex has a helical structure due to the unusual coordina- between the low spin iron(III) centre and Cu(1) (J = 3.7 cm–1 )
tion mode of this imide ligand. Specifically, due to the presence of vs Cu(1a) (J = 7.0 cm–1 ). The difference is due to the degree of
a 2,2 -bipyridine moiety, this ligand is a capable of being ditopic orthogonal overlap achieved along one side of the rectangle

Fig. 28. The chain complex [CuII (tcm)(bpcam)]∞ . Hydrogen atoms have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [65].
2962 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Fig. 29. The structure of [CuII 2 (bbpca)2 ](PF6 )2 ·2CH3 CN·CH3 CH2 OH oriented to highlight the unusual N2 /N binding mode of the imide anion and the resulting helical twist.
Hydrogen atoms, uncoordinated anions and solvents have been excluded for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [81].

Table 6
The distances between copper(II) and iron(III) centres along the edges of the rectangular heterometallic complexes.

Complex Fe· · ·Cu (Å)a Fe· · ·Cu (Å)b Cu· · ·Fe· · ·Cu (◦ ) Fe· · ·Cu· · ·Fe (◦ )
II III
[Cu 2 (bpcam)2 Fe 2 (CN)6 ](PPh4 )4 4.822 4.991 77.2 102.8
[CuII 2 (bpca)2 FeIII 2 (CN)6 ](AsPh4 )4 5.041 5.211 89.9 90.1
{(Tp)2 [FeIII (CN)3 ]2 [CuII (bpca)]2 }·4H2 O 5.023 5.217 90.3 89.7
[CuII 2 (bpca)2 (H2 O)2 FeIII 2 (5,5 -dmbipy)2 (CN)8 ][CuII (bpca)FeIII (5,5 -dmbipy)(CN)4 ]·4H2 O 4.973 5.363 91.91 88.1
a
Cyanide bridge in equatorial plane of the copper centre.
b
Cyanide bridge in axial position of the copper centre.

between the magnetic (equatorial) orbital of the copper(II) motif (Table 6) and hence the same arrangement of magnetic
centre and the t2g magnetic orbital of the S = 1/2 iron(III) orbitals. However, the magnetic properties were modelled with a
centre. The temperature dependent magnetic properties of Hamiltonian only accounting for one exchange parameter, rather
[CuII 2 (bpca)2 FeIII 2 (CN)6 ](AsPh4 )4 were not reported. than the two exchange parameters reported for the other exam-
The structurally similar complex {[(Tp)2 FeIII (CN)3 ]- ples. In this case, the analysis showed weak ferromagnetic coupling
II
2 [Cu (bpca)]2 }·4H2 O has the same, almost square, structural (J = 1.38 cm−1 ; H = −2 J[SFe(1) (SCu(1) + SCu(2) ) + SFe(2) (SCu(1) + SCu(2) )]
[83]. AC measurements performed between 1.8 and 10 K showed
that the complex does not undergo magnetic ordering or slow spin
relaxation.
Ferromagnetic coupling is also observed for the co-crystallised
complex [CuII 2 (bpca)2 (H2 O)2 FeIII 2 (5,5 -dmbipy)2 (CN)8 ]-
II III 
[Cu (bpca)Fe (5,5 -dmbipy)(CN)4 ]2 ·4H2 O (Table 6). The cal-
culated coupling was smaller in the rectangular fragment
[CuII 2 (bpca)2 (H2 O)2 FeIII 2 (5,5 -dmbipy)2 (CN)8 ] (J1 = 2.64 cm−1 ;
H = −2J1 [SFe(1) ·(SCu(1) + SCu(1a) ) + SFe(1a) ·(SCu(1) + SCu(1a) )] − 2 × 2J2 SFe2
·SCu2 ), where the (CuII (bpca))+ unit is bridged both axially and
equatorially, than in the linear fragment [CuII (bpca)FeIII (5,5 -
dmbipy)(CN)4 ]2 (J2 = 5.40 cm−1 ), where the (CuII (bpca))+ unit is
linked equatorially to an iron(III) centre [84].
Repositioning of the methyl groups on the bipyri-
dine moiety led instead to the one-dimensional chain
complex {[CuII (bpca)][FeIII (4,4 -dmbipy)(CN)4 ]}∞ (Fig. 31, 4,4 -
dmbipy = 4,4 -dimethyl-pyridine) [84]. The copper(II) centres
adopt a trigonal bipyramidal geometry, with two of the three trig-
onal plane sites occupied by bridging cyanide groups, each from
a separate [FeIII (4,4 -dmbipy)(CN)4 ]− moiety, and the third site
occupied by the imide nitrogen atom. Adjacent complexes in the
chain are generated by translation, resulting in a zig-zag pattern
in which the copper(II) and low spin iron(III) centres are fer-
Fig. 30. In the rectangular complex [CuII 2 (bpcam)2 FeIII 2 (CN)6 ](PPh4 )4 each cop- n−1
per(II) centre has two different CuII · · ·FeIII contacts along the edge of the rectangle. romagnetically coupled (J = 4.82cm−1 ; H = −J i=1 SAi · SAi+1 ).
Hydrogen atoms, uncoordinated anions and solvent molecules have been omitted A small intermolecular ferromagnetic coupling was also
for clarity. calculated (zJ = 0.029 cm–1 ), using the van Vleck equation.
Source: This figure was generated from data obtained from the CCDC as published
Field dependent magnetisation studies at 1.8 K revealed no
originally in reference [82].
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2963

Fig. 32. The structure of [CrIII (phen)(C2 O4 )2 CuII (bpca)(H2 O)]·2H2 O, a discrete
chromium(III)/copper(II) complex incorporating the [CuII (bpca)]+ unit. Hydrogen
atoms and solvent molecules have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published
Fig. 31. The chain complex [CuII (bpca)FeIII (4,4 -dmbipy)(CN)4 ]∞ which has both
originally in reference [86].
intra and intermolecular ferromagnetic coupling. Hydrogen atoms have been omit-
ted for clarity.
Source: This figure was generated from data obtained from the CCDC as published Ph Ph
originally in reference [84]. O O
TiCl2
O O O O
evidence for slow relaxation and hence no single chain magnet O O
behaviour. Ph Ph
+ O N
In the discrete trinuclear complex {[CuII (bpca)(H2 O)]- O N R
[AgI (CN)2 ][CuII (bpca)(H2 O)]} the two square pyrami- R
dal [CuII (bpca)(H2 O)]+ units are linked equatorially by
the NC–AgI –CN bridge. However, in the chain complex Fig. 33. A general scheme for asymmetric Diels–Alder and 1,3-dipolar cycloaddition
reactions catalysed by the chiral titanium imide complex.
{AuI [CuII (bpca)(NC)2 ]·H2 O}∞ the gold(I) ions form two types
of links, one bridging the equatorial sites and the other bridging
axial sites of the square pyramidal copper(II) ions in adjacent enantiomeric excess observed for the reactions. Elucidation of this
[CuII (bpca)(NC)2 ]− units [85]. Relatively long CuII · · ·CuII separa- intermediate structure helped the authors determine the mecha-
tions, of 15.022 Å (AgI link) and 10.085 Å (AuI equatorial-link), nism of the catalytic reaction. Similar reaction intermediates have
result. Not surprisingly, given these long separations, a magnetic been isolated by Hinterman and Seebach during a modification of
measurement of the gold(I) linked chain showed only very weak the Evans auxiliary, also used to produce chiral products [88].
antiferromagnetic coupling [J = −0.045 cm−1 ; H = −2J˙(Si · Si+1 )].
The mono-anionic linkers [CrIII (bipy)(C2 O4 )2 ]− and 6.2. Chromium
[CrIII (phen)(C2 O4 )2 ]− are in principle able to bridge [CuII (bpca)]+
units by coordination through both oxalate ligands. In prac- The complex [CrIII (bpca)2 ]ClO4 was prepared in 5.1% yield
tice, however, they instead form the discrete dimetallic by hydrolysis of TPyT (Fig. 3) [11]. Electrochemical studies in
complexes, [CrIII (bipy)(C2 O4 )2 CuII (bpca)(H2 O)]·2.5H2 O and
[CrIII (phen)(C2 O4 )2 CuII (bpca)(H2 O)]·2H2 O (Fig. 32) [86]. In both
complexes the [CuII (bpca)]+ units are connected through an axial
Cu-O bond involving a single oxygen atom of the oxalate, com-
pleting the square pyramidal coordination sphere of the copper(II)
centre, and resulting in CrIII · · ·CuII distances of 5.283 and 5.407 Å,
respectively. No magnetic interaction was observed between
the octahedral CrIII and square pyramidal CuII centres, unlike
the ferromagnetic coupling observed for the complexes where a
bis-bidentate oxalate bridges two octahedral copper(II) centres
(Section 5.5, Fig. 21).

6. Other first row transition metal ion complexes of imides

6.1. Titanium

A titanium complex containing a non-symmetric imide ligand


has been reported as the active form of a chiral alkoxytitanium cata-
lyst for Diels–Alder and 1,3-dipolar cycloaddition reactions (Fig. 33)
Fig. 34. The structure of the imide-containing titanium chiral-catalyst intermediate.
[87]. During the reaction the imide substrate acts as a ligand, coor- Hydrogen atoms have been omitted for clarity.
dinating to titanium through the imide oxygen atoms (Fig. 34), Source: This figure was generated from data obtained from the CCDC as published
generating a chiral 6-coordinate complex and facilitating the high originally in reference [87].
2964 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Fig. 35. The structure of [MnII (bpca)2 ]·H2 O. Note that hydrogen atoms and the sol-
vent molecule have been omitted for clarity. Note that hydrogen atoms and solvent
molecules have been excluded for clarity.
Fig. 36. The structure of [CoIII (bpca)2 ]ClO4 ·H2 O. Hydrogen atoms and the solvent
Source: This figure was generated from data obtained from the CCDC as published
molecule have been omitted for clarity.
originally in reference [89].
Source: This figure was generated from data obtained from the CCDC as published
originally in reference [11].
acetonitrile showed reversible processes at –0.74 and –1.47 V
vs SSCE, both of which were assigned as metal based. The com- is concentrated on the imide nitrogen atom in [MnII (bpca)2 ]
plex was combined with iron(II) perchlorate to prepare the but delocalises across the imide group upon coordination to the
chain complex {[(CrIII (bpca)2 )FeII (ClO4 )2 ]ClO4 ·CH3 NO2 ·H2 O}∞ . [MnII (hfac)2 ] units. It was proposed that the increased Mn-Nimide
The differently solvated chain complex, distances (compared to other late first row MII ) reduced the d␲–p␲
{[CrIII (bpca)2 FeII (ClO4 )2 ]ClO4 ·1.5CH3 NO2 ·2.5H2 O}∞ , was pre- interactions between the manganese(II) and the coordinated bpca
pared in nitromethane by combining [CrIII (bpca)2 ]ClO4 and anion, leading to the observed weak antiferromagnetic coupling
iron(III) perchlorate, using catechol as an in situ reducing agent, as between the manganese(II) centres in the trinuclear complex
purple crystals in 68% yield [11]. As for the analogous FeIII/II chains (J = –0.35 cm−1 ; H = −2J(SMn1 · SMn2 + SMn1 · SMn3 )) [90].
(Section 4.2), antiferromagnetic coupling is observed between
the chromium(III) S = 3/2 and high spin iron(II) S = 2 centres 6.4. Cobalt
(Curie–Weiss model: CrIII /FeII  = −12.5(4) K vs FeIII/II  = −15.3(6) K
[11]), in contrast to the ferromagnetic coupling observed in the The cobalt complex [CoIII (bpca)2 ]ClO4 ·solvent (Fig. 36) was
NiII /FeII chains discussed below (Section 6.5). This is believed to be reported in 2002 by two research groups [11,24]. In one case
due to both CrIII and low spin FeIII lacking electron density in the eg [24], [CoIII (bpca)2 ]ClO4 ·CH3 OH was prepared in 70% yield by air
orbitals, in stark contrast to NiII , which prevents propagation of fer- oxidation of the corresponding amide ligand in methanol solu-
romagnetic coupling with the adjacent high spin FeII ions. Whereas tion with cobalt(II) perchlorate and sodium methoxide. Addition
both the FeIII/II and NiII /FeII chains are ferrimagnetically coupled of hydrogen peroxide increased the yield to 75%. The structure
overall, this CrIII /FeII chain is not: the authors proposed that this of [CoIII (bpca)2 ]ClO4 ·CH3 OH is slightly compressed compared to
was due to the similar spin multiplicities of the chromium(III) and [FeIII (bpca)2 ]ClO4 , with average M–N bond lengths of 1.943 Å com-
high spin iron(II) centres and the weak coupling between them pared to 1.978 Å.
[11]. A similar amide → imide oxidation was observed in this research
group: air oxidation of the amide complex [CoIII (L1M )2 ]BF4 was pro-
6.3. Manganese posed to have formed a mixture of the mixed amide-imide complex
[CoIII (L1M )(pypzca)]+ and the all imide complex [CoIII (pypzca)2 ]+ in
The preparation of [MnII (bpca)2 ]·H2 O (Fig. 35) was reported in solution (Fig. 1), based on electrochemical evidence (Fig. 37) [25].
1990 [89], when it was studied for zero-field splitting parameters. Independently of the first group, [CoIII (bpca)2 ]ClO4 ·H2 O (Fig. 36)
The complex was prepared as red-brown crystals by mixing Hbpca was also prepared by reaction of Hbpca and cobalt(II) per-
and manganese(II) nitrate in acetone and water, using sodium chlorate in a nitromethane/water mixture in 82% yield [11].
hydroxide to deprotonate Hbpca. The structure is similar to those of Electrochemical studies of the complex showed a metal based
other reported [MII (imide)2 ] complexes. EPR studies of the complex reduction at −0.28 V vs SSCE (CoIII → CoII ) and a second reduc-
and of a sample of [ZnII (bpca)2 ] doped with 1% of the complex were tion at −1.66 V vs SSCE. The second reduction was assigned as the
performed at 130 and 298 K. These gave the axial |D| = 0.073 cm−1 CoII → CoI reduction [11], however recent results have suggested
and equatorial |E| = 0.017 cm−1 zero field splitting parameters. The it may be ligand based [91]. The mononuclear [CoIII (bpca)2 ]ClO4
low A (hyperfine) value obtained is consistent with significant was then used to generate the discrete trimetallic struc-
unpaired electron density on the ligands. ture {[FeII (ClO4 )2 ][CoIII (bpca)2 )2 ](ClO4 )2 (CH3 NO2 )4 } and the chain
In the discrete trinuclear complex {[MnII (bpca)2 ][MnII (hfac)2 ]2 } complex {[FeII (ClO4 )2 ][CoIII (bpca)2 ]ClO4 (CH3 NO2 )(H2 O)}∞ . The
the bridging [MnII (bpca)2 ] unit binds to two [MnII (hfac)2 ] moieties diamagnetic cobalt(III) centres separate the high spin iron(II) ions
via the imide oxygen atoms. As a result the Mn–Nimide (2.196 and in the chain, so the iron(II) ions are magnetically independent at
2.203 Å) bond distances are elongated compared to those seen room temperature. However, between 250 and 50 K a weak fer-
in the free [MnII (bpca)2 ] moiety (2.179 and 2.169 Å). This was romagnetic interaction of  = +0.2 (3) K was estimated using the
taken to suggest that the negative charge of the imide system Curie–Weiss model.
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2965

10 µA

Fig. 39. The graphite like structure of {[FeII [NiII (bpca)2 ]1.5 ](ClO4 )2 ·4.5CH3 Cl·
3CH3 OH·10H2 O}∞ . Non-coordinated anions and solvent molecules have been omit-
ted for clarity.
Source: This figure was generated from data obtained from the CCDC as published
originally in reference [10].

complex {[FeII [NiII (bpca)2 ]1.5 ](ClO4 )2 ·4.5CH3 Cl·3CH3 OH·10H2 O}∞
and the discrete, deep violet, trimetallic complex
{[NiII (bpca)2 ][FeII (ClO4 )2 (H2 O)2 ][NiII (bpca)2 ]·4H2 O} were pro-
duced by combining a 2:1 ratio of [NiII (bpca)2 ] and [FeII (H2 O)6 ]ClO4
in a nitromethane/chloroform mixture, the former complex
crystallising from this mixture and the latter complex crys-
-1.4 -1.2 -1 -0.8 -0.6 -0.4 -0.2 0 tallising when water was added to the reaction solution. In
comparison to the one-dimensional chain structures described
V earlier (Section 4.2), where the iron(II) centre bridges two
[M(imide)2 ] units, the iron(II) centre in the polymeric complex
Fig. 37. Cyclic voltammograms of (top) [CoIII (L1E )2 ](BF4 )·0.5H2 O; (middle) fresh
[CoIII (L1M )2 ](BF4 )·0.25H2 O; and (bottom) aged [CoIII (L1M )2 ](BF4 ) that was believed
{[FeII [NiII (bpca)2 ]1.5 ](ClO4 )2 ·4.5CH3 Cl·3CH3 OH·10H2 O}∞ bridges
to also contain [CoIII (L1M )(pypzca)]+ and [CoIII (pypzca)2 ]+ . Scan rate: 200 mV s−1 . three [NiII (bpca)2 ] units, resulting in a graphite-like structure
Reference: 0.01 M AgNO3 /Ag. (Fig. 39). Adjacent layers are related by mirror symmetry so
Source: Reprinted with permission from reference [25]. Copyright 2009 Wiley. are made up of all  then all  isomers of these enantiomeric
iron(II) complexes. These layers are separated by uncoordinated
Sodium and potassium ions have also been used to bridge perchlorate anions and solvent molecules, which also occupy
[CoIII (bpca)2 ]+ units via the imide oxygen atoms [92]. This was the pore spaces. Both anions and solvent can be exchanged, for
achieved using a variety of co-ligands including [Co(NCS)4 ]2− , hexafluorophosphate anions and bromoform, respectively. The
NCS− , N3 − and O2− . Four binding modes for the imide oxygen atoms discrete complex was considered to be a precursor to the polymeric
were observed, as shown in Fig. 38 (left to right); the common complex, and the addition of water to the reaction mixture was
N3 :␮-O2 mode (Fig. 2); an unusual mode where separate cations used to stabilise it, allowing its isolation.
are bound to each imide oxygen atom, providing a five atom bridge; The magnetic properties of these complexes were reported
a mode where the imide oxygen atoms coordinate in the common in a subsequent paper [93]. In the discrete trinuclear com-
N3 :␮-O2 mode to one metal centre and another is involved in a fur- plex, {[NiII (bpca)2 ][FeII (ClO4 )2 (H2 O)2 ][NiII (bpca)2 ]}·4H2 O,
ther (weaker) interaction; and finally another unusual mode where the high spin iron(II) centre bridges two [NiII (bpca)2 ] units.
the pair of imide oxygen atoms doubly bridge the two metal centres. Magnetic studies revealed weak ferromagnetic coupling
between the NiII and FeII centres (J = 0.56(1) cm−1 ; H =
6.5. Nickel −2J(SNi · SFe2 + SFe2 · SNi3 )). From the structural analysis of the poly-
mer, {[FeII [NiII (bpca)2 ]1.5 ](ClO4 )2 ·4.5CH3 Cl·3CH3 OH·10H2 O}∞ , the
The neutral building block [NiII (bpca)2 ] has also been bond lengths indicate that the iron(II) centre is in the high
bridged by iron(II) centres via the imide oxygen atoms spin state (FeII –Oaverage = 2.07 Å). This is consistent with the
[10,93]. In the initial study the polymeric, dark purple, magnetic moment from 300 to 40 K. Below 40 K zero-field

M M
M M M M
O O O O O O O O
M
N N N N
N Co N N Co N N Co N N Co N

Fig. 38. The binding modes observed for [CoIII (bpca)2 ]+ units linked using M Na+ or K+ in the presence of various co-ligands. The first N3 :␮-O2 (Fig. 2) is common, but the
rest are unique to this study.
2966 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Fig. 40. The general structure of ([FeII X2 [NiII (bpca)2 ]](ClO4 )0–2 ·solvents)∞ (X = CH3 CH2 OH or CH3 CH2 CH2 OH) chain complexes where the coordinated solvents are arranged
cis around the iron(II) centre. Hydrogen atoms, uncoordinated anions and solvents have been omitted for clarity.
Source: This figure (X = CH3 CH2 OH) was generated from data obtained from the CCDC as published originally in reference [93].

splitting causes the magnetic moment to decrease signifi- 7. Rhodium, rhenium and platinum complexes of imides
cantly. The structure of an isomorphous iron(II) chain complex
{[FeII [FeII (bpca)2 ]3 ](ClO4 )4 ·5CHCl3 ·2CH3 OH·7H2 O}∞ was also The first Hbpca complex of rhodium, pale yellow
reported, however in that case the yield and purity was limited by [RhIII (bpca)2 ]PF6 , was synthesised in a successful attempt to
the tendency of [FeII (bpca)2 ] to oxidise, so the magnetic properties determine the hydrolysis mechanism of TPyT ligands (Fig. 3 route
were not reported. (i)) [17]. The geometry of [RhIII (bpca)2 ]PF6 was described as a
A series of chain complexes {[FeII X2 [NiII (bpca)2 ]](ClO4 )n ·solven- distorted octahedron, with two bpca anions each coordinated
ts}∞ where X = methanol, ethanol, 1,3-propandiol, thiocyanate and meridionally via the N3 donors. As with the first row tran-
bromide/water were also reported [93]. Despite the variation of the sition metal complexes, the bond lengths to the imide nitrogen
charge of the complexes, due to coordination of solvent vs anions, (1.993–1.996 Å) are compressed in comparison to the pyridyl bonds
the structures are very similar. The key structural difference is (2.015–2.048 Å). Likewise the two adjacent 5-membered chelate
the arrangement of the 2 X moieties in the coordination sphere of rings prevent the ligand providing a 180◦ Npyridine –Rh–Npyridine
the iron(II) centre: cis X for ethanol (Fig. 40) and 1,3-propandiol; angle (intraligand: 162.7◦ and 162.9◦ ). The electrochemistry of the
trans X for the others (Fig. 41). The cis vs trans arrangements of complex was studied in dry dimethylformamide and the quasi-
X result in two different orientations for the adjacent, iron(II) reversible reduction of RhIII → RhI was observed as a two electron
bridged, imide units: either approximately orthogonal for cis process at −1.15 V vs Ag/AgCl. Unlike many other RhIII complexes,
(Fig. 40) or approximately co-planar for trans (Fig. 41). Studying perhaps due to the delocalised nature of the bpca anion, the
the temperature dependent magnetic properties of X = methanol reduction did not resolve into two one electron processes at high
and n = 2 revealed a ferromagnetic interaction between nickel(II) scan rates (20 V s−1 ). Irreversible reductions at −1.44 and −1.84 V
and the high spin iron(II) centre with a best fit of  = +1.6(3) K, vs Ag/AgCl were assigned as ligand-based processes.
estimated using the Curie–Weiss law. Despite the one step two electron reduction in dimethylfor-
The discrete trimetallic complexes {[NiII (bpca)2 ][MII (hfac)2 ]2 } mamide, RhIII → RhI , of [RhIII (bpca)2 ]PF6 , a later study reported
where MII = Mn or Fe are structurally analogous to those dis- the preparation of the greenish-brown rhodium(II) complex,
cussed in Sections 4.2, 5.3 and 6.3 [47]. However, in contrast [RhII (bpca)2 ]·3H2 O (Fig. 42) [94]. This was achieved by performing
to the previous examples, weak ferromagnetic interactions hydrolysis of TPyT with rhodium(III) chloride in 1:1 ethanol:water
were observed between the nickel(II) (S = 1) and manganese(II) under an argon atmosphere, followed by washing away the ionic
(S = 5/2; J = +0.13(2) cm−1 ) or iron(II) (S = 2; J = 0.37(1) cm–1 ) cen- [RhIII (bpca)2 ]PF6 and column chromatography of the remaining
tres, respectively (H = −2J(SM · SNi2 + SNi2 · SM )). In this family residue which afforded the product in a yield of 20%. The mixed
of {[M1II (bpca)2 ][M2II (hfac)2 ]2 } complexes, the dominant spin ligand complex [RhII (bpca)(tpy)]Cl·8H2 O (14%) was obtained
exchange pathway propagates through the eg orbitals of M1II . using a similar method, starting from [RhIII (tpy)Cl3 ]. The bite
These are occupied for M1II = Ni, and are orthogonal to the eg angle of the bpca anion (162.4–163.2◦ ) was similar for both
orbitals on M2II : this combination gives rise to ferromagnetic rhodium(II) complexes (Fig. 42). Results from EPR spectroscopy
coupling. of the rhodium(II) complexes showed that the unpaired electron
is located in the dx2 −y2 orbital. The complexes have an axially
compressed octahedral geometry (RhII –Nimide 1.994, RhII –Npyridine
6.6. Zinc
2.040 Å), typical of [M(bpca)2 ] complexes (other than in the case of
The diamagnetic zinc complex [ZnII (bpca)2 ]·H2 O was prepared Jahn–Teller elongated M = CuII ). The unusual stability of these low
for use as a dopant in EPR studies of [MnII (bpca)2 ]·H2 O. However, spin (S = 1/2) rhodium(II) complexes was attributed to the absence
no physical data regarding [ZnII (bpca)2 ]·H2 O were reported [89]. of an electron in the dz2 orbital significantly reducing the lability

Fig. 41. The general structure of a series of ([FeII X2 [NiII (bpca)2 ]](ClO4 )0–2 ·solvents)∞ (X = CH3 OH, NCS− and Br/H2 O) where the coordinated solvents/anions are arranged
trans around the iron(II) centre. Hydrogen atoms, uncoordinated anions and solvents have been omitted for clarity.
Source: This figure (X = CH3 OH) was generated from data obtained from the CCDC as published originally in reference [93].
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2967

Fig. 44. The cyclic voltammogram of [RhIII (bpca)(tpy)][PF6 ]2 in dry dimethylfor-


mamide when under an argon or a carbon dioxide atmosphere.
Source: Reprinted with permission from reference [95]. Copyright 1998 American
Fig. 42. The crystal structure of the rhodium(II) complex, [RhII (bpca)2 ]·3H2 O. Chemical Society.
Hydrogen atoms have been excluded for clarity.
Source: This figure was generated from data obtained from the CCDC as published
originally in references [94].
potentials of –1.44 and –1.24 V vs SSCE, respectively (Fig. 44)
of these complexes. Electrochemical studies of [RhII (bpca)2 ]·3H2 O [95]. Addition of small amounts of water (2.5%) to act as a
in acetonitrile showed a reversible one electron process at −1.13 V proton source significantly increased the effectiveness of these
vs SCE, assigned as RhII → RhI , along with the ligand-based pro- complexes as catalysts. Under optimal conditions, at a poten-
cesses observed for [RhIII (bpca)2 ]+ (Fig. 43). For [RhIII (bpca)2 ]PF6 tial of -1.49 V vs SSCE, [RhIII (bpca)(tpy)][PF6 ]2 ·CH3 CN produced
a two electron reduction assigned as RhIII → RhI is observed at formate at a rate of 9.5 mol molcatalyst −1 h−1 with a charge effi-
approximately the same potential (Fig. 43). ciency of 78.4%. [RhIII (bpca)(2-picolinamdie)Cl][PF6 ]·H2 O required
The rhodium(III) complexes [RhIII (bpca)(tpy)][PF6 ]2 ·CH3 CN and a smaller applied voltage (−1.31 V vs SSCE), however per-
[RhIII (bpca)(2-picolinamide)Cl][PF6 ]·H2 O were also synthesised formed at a poorer rate, 7.6 molformate molcatalyst −1 h−1 and with
via hydrolysis of [RhIII (tptz)Cl3 ] [95]. The crystal structure of a lower charge efficiency (68.4%). These rates and efficiencies
[RhIII (bpca)(tpy)][PF6 ]2 ·CH3 CN allowed a structural comparison of were similar to those reported for other rhodium(III) com-
the coordinated tridentate bpca anion and tpy ligands. The bpca plexes of bipyridyl ligands and higher than those of iridium(III)
anion had a slightly smaller bite angle (80.18(9)◦ vs 81.31(9)◦ ) complexes.
and remained more planar, with the exception of one imide When the rhenium(V) amine complex [ReV 2 (␮-O)Cl2 (tris(2-
oxygen atom, than the tpy ligand. For the bpca anion there was pyridylmethylamine)2 ]2+ was exposed to air for several weeks
only a 3.1(2)◦ angle between the pyridyl rings whereas for tpy the ligand oxidised to the imide, losing a 2-pyridylmethylamine
the outer pyridyl rings were at angles of +4.64(2)◦ and +8.03(2)◦ arm in the process, to form the imide complex [ReIII 2 (␮-
relative to the central pyridyl ring. Electrochemical studies of O)O2 Cl2 (bpca)2 ]·CH3 CN (Fig. 45) [96]. This was the first example of
these complexes in dimethylformamide showed an irreversible an O = ReIII –O–ReIII O bridge in a complex with a tridentate ligand.
two electron reduction, RhIII → RhI , at –0.76 and –0.98 V vs SSCE No further properties of the complex were discussed.
respectively, followed by a one electron ligand-based process at
–1.53 V vs SSCE. Along with another two rhodium(III) complexes,
solutions of [RhIII (bpca)(tpy)][PF6 ]2 ·8H2 O and [RhIII (bpca)(2-
picolinamide)Cl][PF6 ] were able to reduce carbon dioxide to
formate (determined using chromoprotic acid) under applied

Fig. 43. The cyclic voltammograms of [RhII (bpca)2 ] and [RhIII (bpca)2 ]+ in acetoni- Fig. 45. The crystal structure of [Re2 (␮-O)O2 Cl2 (bpca)2 ]·CH3 CN. Hydrogen atoms
trile. and uncoordinated solvent molecules have been omitted for clarity.
Source: From reference [94] – Reproduced by permission of The Royal Society of Source: This figure was generated from data obtained from the CCDC as published
Chemistry. originally in reference [96].
2968 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Fig. 46. Coordination sphere around the europium ion (left) within the structure of the [EuIII 3 (bcpca)3 (H2 O)6 ]·3DMSO·12H2 O complex (right). Non-acidic hydrogen atoms
and uncoordinated solvent molecules have been omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [13].

8. Lanthanide complexes of imides [LnIII (bcpca)(H2 O)2 ] unit, both of which sit above this plane. The
lanthanide ion itself sits 0.855 (Eu) and 0.846 (Gd) Å above this N3
8.1. Overview of lanthanide complexes plane. The coordination of the pair of bridging imide oxygen atoms
from the adjacent complex is complimentary, such that the crystal-
The incorporation of lanthanide (Ln) ions opens up many fur- lographic three-fold axis generates a triangle of these units (Fig. 46).
ther possibilities, including the study of luminescence [97] and of Comparison of the solid state and solution luminescence spectra
magnetic properties resulting from 3d–4f interactions [41]. Three of [EuIII 3 (bcpca)3 (H2 O)6 ] led to the conclusion that the triangular
classes of imide lanthanide complexes have been reported, those structure was maintained in solution. The expected series of 5 D0 →
in which the lanthanide ion: (a) is bound in the N3 pocket, (b) 7 F transitions was observed, with an overall quantum yield of 37%.
x
is bound by the ␮-O2 of a transition metal [M(imide)2 ]0/+ com- Both complexes were evaluated [14] for potential use as medical
plex, (c) is co-crystallised with a transition metal [M(imide)2 ]0/+ magnetic resonance imaging (MRI) agents [98]. Magnetic stud-
complex. These three classes are presented and discussed in the ies of the [GdIII 3 (bcpca)3 (H2 O)6 ]·3.5DSMO·12H2 O triangle complex
following three sections. In the first of these, the basic design of showed that no interaction occurred between the GdIII (S = 7/2)
Hbpca has been adapted to produce pentadentate imide ligands centres [14]. The molecular relaxivity of the complex was studied
(Fig. 1) capable of binding the lanthanide ions, which prefer higher by NMR spectroscopy on a 9:1 water/dimethylformamide solu-
coordination numbers and carboxylate donors (Section 8.2). The co- tion and was calculated to be 28.5 s−1 mM−1 at 20 MHz, with a
crystallised systems are described for completeness (Section 8.3). density of relaxivity of 18.8 g mol–1 s–1 mM–1 , a reasonably good
Transition metal imide complexes have also proven effective for value. The relaxivity of the complex was limited by the rela-
bridging lanthanide ions through the imide oxygen atoms (Section tively slow exchange rate of the coordinated water molecules
8.4; Fig. 2, N3 :␮O2 ). These distinct structural outcomes appear to (0.3 × 106 s−1 ). As part of the study, the thermodynamic stability
depend on both the choice of co-ligands and the charge of the imide of [EuII 3 (bcpca)3 (H2 O)6 ] was also evaluated from a titration with
complex. EDTA4− . The results showed that, whilst not stable enough for med-
ical applications [98], the complexes were reasonably stable at a
range of physiological pH values and in the presence of competing
8.2. Lanthanide complexes of imide ligands
ligands [14].
The ligand bis(6-carboxypyridine-2-carbonyl)amine (H3 bcpca,
Fig. 1) has been used to produce the only examples of an imide
binding to a lanthanide through the N3 donor set (Fig. 2) [13,14]. 8.3. Co-crystallisation of transition metal imide complexes with
The methyl ester of H3 bcpca hydrolyses by the combination of lanthanide ions
coordination to the Lewis acidic lanthanide ion and the pres-
ence of 3.6 equivalents of sodium hydride, in DMF, producing the Mononuclear imide complexes such as low spin [FeIII (bpca)2 ]+
triangular complexes [EuIII 3 (bcpca)3 (H2 O)6 ]·3DMSO·12H2 O and could reasonably be expected to coordinate to lanthanide
[GdIII 3 (bcpca)3 (H2 O)6 ]·3.5DMSO·12H2 O (Fig. 46), after recrystalli- ions through the imide oxygen atoms. While this is pos-
sation by water diffusion into a DMSO solution. One third of the sible (Section 8.4), there have also been reports of co-
triangle, a [LnIII (bcpca)(H2 O)2 ] moiety, is present in the asymmet- crystallisation products, namely [FeIII (bpca)2 ][EuIII (NO3 )4 (H2 O)2 ]
ric unit, with the other two thirds of the triangle generated by a [99] and [FeIII (bpca)2 ][ErIII (NO3 )3 (H2 O)4 ]NO3 [100]. In these com-
three-fold axis. The imide ligand has a pentadentate pocket made plexes only weak hydrogen bonds connect the discrete transition
up of the usual N3 donor set plus two oxygen atoms from the metal and lanthanide complexes. Magnetically the d and f electron
pendant carboxylate groups. The coordination sphere of the lan- systems of the discrete complexes do not interact and the imide
thanide ion is completed by one coordinated water molecule sited units act to physically and magnetically isolate the lanthanide ions.
below the N3 plane of the bcpca anion, and a coordinated water Magnetic studies and ab intio calculations were used to determine
molecule and a pair of bridging imide oxygen atoms from an adjacent and model the easy axis of magnetisation of the lanthanide ions.
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2969

(FeIII S = 1/2, DyIII S = 5/2), which occurs via the  skeleton of the
ligand rather than the delocalised system.
The neutral imide building blocks [NiII (bpca)2 ] and [FeII (bpca)2 ]
have been used as bridging moieties to generate the isostructural
discrete trinuclear species {[DyIII (hfac)3 ]2 [MII (bpca)2 ]}·CHCl3 with
MII = Ni or Fe (Fig. 48), similar to those seen in previous Sections 4.2,
5.3, 6.3 and 6.5 but with an additional hfac ion bound per Ln, from
a chloroform solution [101]. Unlike in the above dinuclear complex
[FeIII (bpca)(␮-bpca)DyIII (NO3 )4 ] (Fig. 47), in these two cases the
imide complex bridges two eight-coordinate lanthanide ions, using
both sets of imide oxygen atoms to do so. This raises the question of
whether the nature of the co-ligands on the lanthanide ion or, more
intuitively, the charge on the central imide unit, is responsible for
the difference in terminal vs bridging coordination by the imide
complex. Further examples are required before this point can be
confidently addressed.
Within these trinuclear complexes the Dy· · ·Dy separations are
11.31 (M = NiII ) and 11.46 Å (M = FeII ). Although no weak interac-
Fig. 47. The crystal structure of [FeIII (bpca)(␮-bpca)DyIII (NO3 )4 ]·H2 O·4.5CH3 NO2 .
tions are observed between the discrete trinuclear complexes, the
Hydrogen atoms, uncoordinated anions and solvent molecules have been omitted shortest Dy· · ·Dy separations are actually intermolecular (NiII 9.84,
for clarity. FeII 9.53 Å). These two complexes provide an excellent case study to
Source: This figure was generated from data obtained from the CCDC as published investigate how the orbital pathways provided by the nickel(II) vs
originally in reference [9]. iron(II) bridges affect communication between the two lanthanide
ions. In contrast to the diamagnetic iron(II) bridge, the paramag-
netic nickel(II) bridge (S = 1), with occupied eg orbitals, provides
8.4. Lanthanide complexes of transition metal/imide a pathway for ferromagnetic coupling between the two dyspro-
metalloligands sium ions (each S = 5/2). Surprisingly, this ferromagnetic interaction
halved the activation barrier required for reversal of magnetisa-
In the dinuclear 3d–4f complex [FeIII (bpca)(␮- tion. This was ascribed to the orthogonal orientation of the bridging
bpca)DyIII (NO3 )4 ]·H2 O·4.5CH3 NO2 , prepared in a imide oxygen atoms resulting in cancellation of the Ising anisotropy
methanol/nitromethane mixture, the N3 :␮-O2 binding mode between the two individual dysprosium(III) ions.
(Fig. 2) occurs with the low spin [FeIII (bpca)2 ]+ cation coordinated The neutral imide building block [NiII (bpca)2 ] has also been used
to the dysprosium(III) ion via both of the imide oxygen atoms to bridge lanthanide ions through both sets of imide oxygen atoms
(Fig. 47) [9]. The remainder of the 10-coordinate dysprosium(III) to produce three 1-dimensional chain complexes with the formula
coordination sphere is completed by four bidentate nitrate anions. {[LnIII (NO3 )(H2 O)3 ][NiII (bpca)2 )](NO3 )2 ·3H2 O}∞ where LnIII = Gd,
Unlike the series of transition metal complexes discussed above, Tb or Dy, from a dichloromethane/ethanol solution (Fig. 49) [102].
the [Fe(bpca)2 ]+ only coordinates through one of the two sets of Here the nine-coordinate lanthanide ion binds four imide oxy-
imide oxygen atoms available, resulting in a dinuclear, rather than gen atoms, three water molecules and one bidentate nitrate anion.
a trinuclear or polymeric, complex. Magnetisation measurements Hence, unlike the previous examples where all of the nitrate anions
between 0.4 and 4.2 K showed hysteresis loops and steps in the are coordinated, this structure has two uncoordinated nitrate
magnetisation. AC susceptibility measurements showed slow anions per repeating unit. Unlike the discrete systems, which have
relaxation of magnetisation consistent with SMM behaviour. An no uncoordinated anions, there are several intermolecular inter-
ab-initio study of the ligand field, spin-orbit and exchange effects actions between the chains that are mediated by the numerous
of the 3d–4f system was used to model the 3d–4f interaction uncoordinated anions present. These include a significant hydrogen

Fig. 48. The structure of the discrete trimetallic complex [(DyIII (hfac)3 }2 )(FeII (bpca)2 )]·CHCl3 . Hydrogen atoms and uncoordinated solvent molecules have been omitted for
clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [101].
2970 M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971

Fig. 49. The LnII Gd example of the isostructural series {[LnIII (NO3 )(H2 O)3 ][NiII (bpca)2 )](NO3 )2 ·3H2 O}∞ . Hydrogen atoms and uncoordinated solvent molecules have been
omitted for clarity.
Source: This figure was generated from data obtained from the CCDC as published originally in reference [102].

bonding network involving both coordinated and uncoordinated limited exploration of metal ions other than iron and copper, there
anions and solvent molecules. Further intermolecular connections is clearly plenty of scope for further development. Indeed it is clear
between adjacent chains occur via stacking interactions between from this review that considerable potential exists for the design,
pyridyl rings. Despite the observation of ferromagnetic exchange development and investigation of new generations of imide based
within the chain, the numerous intermolecular interactions pre- ligands and coordination complexes with exciting properties. Such
cluded the observation of SCM behaviour. studies are currently underway in many of the groups referenced
Interestingly, in the similar chain complexes above, as well as in this group [91,104], and in others [105].
{[(LnIII (NO3 )3 ][MII (bpca)2 ]·xH2 O}∞ (LnIII = Gd, Tb, Dy, Ho, Y(Ni
II
only) and M = Ni, Zn) all of the nitrate counter-ions are coordinated Acknowledgements
to the ten-coordinate lanthanide centres, the only difference in
preparation being the use of methanol/chloroform as the solvent We are grateful to the University of Otago for funding this
[103]. For this system, no intermolecular interactions are reported research (including the award of a Postgraduate Scholarship to
between the chains. Consequently the chains are magnetically MGC) and to Dr David McMorran (University of Otago) for proof-
isolated, and the three chains where MII = Ni and LnIII = Tb, Dy or Ho reading an early version of this review.
exhibit slow relaxation of magnetisation in a DC field at tempera-
tures between 1.8 and 10 K (field induced single molecule magnet References
behaviour). Where the lanthanide ions are separated by diamag-
netic [ZnII (bpca)2 ] bridges, the magnetic properties are similar to [1] E.I. Lerner, S.J. Lippard, J. Am. Chem. Soc. 98 (1976) 5397.
[2] E.I. Lerner, S.J. Lippard, Inorg. Chem. 16 (1977) 1546.
those of isolated Ln(III) ions. Nevertheless these zinc/lanthanide
[3] G.R. Desiraju, Acc. Chem. Res. 35 (2002) 565.
complexes also undergo slow relaxation when a DC field is applied [4] F.H. Allen, C.A. Baalham, J.P.M. Lommerse, P.R. Raithby, Acta Crystallogr. Sect.
(field induced single molecule magnet behaviour). B B54 (1998) 320.
[5] (a) C.A. Hunter, J.K.M. Sanders, J. Am. Chem. Soc. 112 (1990) 5525;
(b) C.R. Martinez, B.L. Iverson, Chem. Sci. 3 (2012) 2191.
9. Concluding remarks [6] T.J. Mooibroek, C.A. Black, P. Gamez, J. Reedijk, Cryst. Growth Des. 8 (2008)
1082.
[7] T. Kajiwara, M. Nakano, Y. Kaneko, S. Takaishi, T. Ito, M. Yamashita, A.
The synthesis, structure and properties of the twelve imide lig- Igashira-Kamiyama, H. Nojiri, Y. Ono, N. Kojima, J. Am. Chem. Soc. 127 (2005)
ands and over 150 structurally characterised imide complexes of 10150.
those ligands reported before June 2011 have been reviewed. The [8] T. Kajiwara, H. Tanaka, M. Nakano, S. Takaishi, Y. Nakazawa, M. Yamashita,
Inorg. Chem. 49 (2010) 8358.
imide ligands are most commonly observed to bind through an N3 [9] M. Ferbinteanu, T. Kajiwara, K.-Y. Choi, H. Nojiri, A. Nakamoto, N. Kojima, F.
donor set which includes the deprotonated imide nitrogen atom, Cimpoesu, Y. Fujimura, S. Takaishi, M. Yamashita, J. Am. Chem. Soc. 128 (2006)
but are also observed to form bonds through the pair of imide oxy- 9008.
[10] A. Kamiyama, T. Noguchi, T. Kajiwara, Angew. Chem. Int. Ed. 39 (2000) 3130.
gen atoms, and sometimes via both. Other binding modes have also [11] T. Kajiwara, R. Sensui, T. Noguchi, A. Kamiyama, T. Ito, Inorg. Chim. Acta 337
been observed on rare occasions, so overall imide ligands are clearly (2002) 299.
established as an extremely versatile ligand type. [12] S. Wocadlo, W. Massa, J.-V. Folgado, Inorg. Chim. Acta 207 (1993) 199.
[13] S. Zebret, N. Dupont, G. Bernardinelli, J. Hamacek, Chem. Eur. J. 15 (2009) 3355.
Both discrete and polymeric, homometallic and heterometallic, [14] S. Zebret, E. Torres, E. Terreno, L. Guénée, C. Senatore, J. Hamacek, Dalton
complexes have been produced and the influence of weak inter- Trans. 40 (2011) 4284.
actions and inclusion of simple co-ligands has produced a wide [15] F.H. Allen, Acta Crystallogr. Sect. B 58 (2002) 380.
[16] F.H. Allen, S.A. Bellard, M.D. Brice, B.A. Cartwright, A. Doubleday, H. Higgs,
variety of structures. To date, studies of these materials have pre- T. Hummelink, B.G. Hummelink-Peters, O. Kennard, W.D.S. Motherwell, J.R.
dominantly focussed on the investigation of magnetic properties, Rodgers, D.G. Watson, Acta Crystallogr. Sect. B 35 (1979) 2331.
with particular highlights being the observation of single molecule [17] P. Paul, B. Tyagi, M.M. Bhadhade, E. Suresh, Dalton Trans. (1997) 2273.
[18] X.-P. Zhou, D. Li, S.-L. Zheng, X. Zhang, T. Wu, Inorg. Chem. 45 (2006) 7119.
and single chain magnetism. Other useful properties observed for
[19] S.K. Padhi, V. Manivannan, Inorg. Chem. 45 (2006) 7994.
these complexes include reversible redox processes and the elec- [20] H. Tanaka, T. Kajiwara, Y. Kaneko, S. Takaishi, M. Yamashita, Polyhedron 26
trocatalytic reduction of carbon dioxide; (gradual) spin crossover; (2007) 2105.
luminescence; anion and solvent exchange; and semi-conductor [21] P.J.S. Miguel, M. Roitzsch, L. Yin, P.M. Lax, L. Holland, O. Krizanovic, M. Lutter-
beck, M. Schürmann, E.C. Fusch, B. Lippert, Dalton Trans. (2009) 10774.
behaviour. Given the wealth of properties reported to date, despite [22] R. Sahu, S.K. Padhi, H.S. Jena, V. Manivannan, Inorg. Chim. Acta 363 (2010)
the very limited number of imide ligands employed and the 1448.
M.G. Cowan, S. Brooker / Coordination Chemistry Reviews 256 (2012) 2944–2971 2971

[23] H. Kooijman, S. Tanase, E. Bouwman, J. Reedijk, A.L. Spek, Acta Crystallogr., [68] R.C. Weast, M.J. Astle, W.H. Beyer, CRC Handbook of Chemistry and Physics,
Sect. C: Cryst. Struct. Commun. C62 (2006) m510. 64 ed., CRC Press, 1983.
[24] J.M. Rowland, M.M. Olmstead, P.K. Mascharak, Inorg. Chem. 41 (2002) 2754. [69] M.-L. Liu, W. Gu, Z.-P. Ma, P. Zhu, Y.-Q. Gao, X. Liu, J. Coord. Chem. 61 (2008)
[25] R.M. Hellyer, D.S. Larsen, S. Brooker, Eur. J. Inorg. Chem. (2009) 1162. 3476.
[26] K.C. Nicolaou, C.J.N. Mathison, Angew. Chem. Int. Ed. 117 (2005) 6146. [70] M.L. Calatayud, I. Castro, J. Sletten, F. Lloret, M. Julve, Inorg. Chim. Acta
[27] H. Chowdhury, S.H. Rahaman, R. Ghosh, S.K. Sarkar, H.-K. Fun, B.K. Ghosh, J. 300–302 (2000) 846.
Mol. Struct. 826 (2007) 170. [71] I. Castro, J. Faus, M. Julve, M. Mollar, A. Monge, E. Gutierrez-Puebla, Inorg.
[28] X.T. Li, C. Zhan, Y. Wang, J. Yao, Chem. Commun. (2008) 2444. Chim. Acta 161 (1989) 97.
[29] P.S. Gentile, T.A. Shankoff, J. Inorg. Nucl. Chem. 27 (1965) 2301. [72] D. Cangussu, H.O. Stumpf, H. Adams, J.A. Thomas, F. Lloret, M. Julve, Inorg.
[30] L. Pauling, Nature of the Chemical Bond, Cornell University Press, Ithaca, New Chim. Acta 358 (2005) 2292.
York, 1948. [73] I. Castro, J. Faus, M. Julve, Y. Journaux, J. Sletten, Dalton Trans. (1991) 2533.
[31] D.D. Bray, Inorg. Chim. Acta 111 (1986) L39. [74] I. Castro, J. Sletten, J. Faus, M. Julve, Y. Journaux, F. Lloret, S. Alvarez, Inorg.
[32] J.P. Roux, J.C.A. Boeyens, Acta Crystallogr., Sect. B: Struct. Sci. 25 (1969) 1700. Chem. 31 (1992) 1889.
[33] P.S. Gentile, J.G. White, D.D. Cavalluzzo, Inorg. Chim. Acta 18 (1976) 183. [75] S.S. Massoud, E. Druel, M. Dufort, R. Lalancette, J. Kitchen, J. Grebowicz, R.
[34] J.P. Roux, J.C.A. Boeyens, Acta Crystallogr., Sect. B: Struct. Sci. 25 (1969) 2395. Vicente, U. Mukhopadhyay, I. Bernal, F.A. Mautner, Polyhedron 28 (2009)
[35] P.S. Gentile, J.G. White, M.P. Dinstein, D.D. Bray, Inorg. Chim. Acta 21 (1977) 3849.
141. [76] I. Castro, M.L. Calatayud, J. Sletten, F. Lloret, J. Cano, M. Julve, G. Seitz, K. Mann,
[36] J.P. Roux, J.C.A. Boeyens, Acta Crystallogr., Sect. B: Struct. Sci. 26 (1970) 526. Inorg. Chem. 38 (1999) 4680.
[37] P.S. Gentile, A.P. Ocampo, Inorg. Chim. Acta 29 (1978) 83. [77] J. Sletten, M. Julve, F. Lloret, I. Castro, G. Seitz, K. Mann, Inorg. Chim. Acta 250
[38] J.P. Roux, Acta Crystallogr., Sect. B: Struct. Sci. 62 (1976) 1171. (1996) 219.
[39] P.S. Gentile, M.P. Dinstein, J.G. White, Inorg. Chim. Acta 19 (1976) 67. [78] J.V. Folgado, E. Escrivà, A. Beltrán-Porter, D. Beltrán-Porter, Transition Met.
[40] P.S. Gentile, J. White, S. Haddad, Inorg. Chim. Acta 13 (1975) 149. Chem. 12 (1987) 306.
[41] S. Brooker, J.A. Kitchen, Dalton Trans. (2009) 7331. [79] H. Casellas, F. Costantino, A. Mandonnet, A. Caneschi, D. Gatteschi, Inorg. Chim.
[42] T. Kajiwara, M. Nakano, Y. Kaneko, S. Takaishi, T. Ito, M. Yamashita, A. Igashira- Acta 358 (2005) 177.
Kamiyama, H. Nojiri, Y. Ono, N. Kojima, J. Am. Chem. Soc. 127 (2005) 10150. [80] J.V. Folgado, E. Coronado, D. Beltrán-Porter, A. Fuertes, C. Miravitlles, Dalton
[43] T. Kajiwara, T. Ito, Mol. Cryst. Liq. Cryst. 335 (1999) 73. Trans. (1988) 3041.
[44] Y. Kaneko, T. Kajiwara, H. Yamane, M. Yamashita, Polyhedron 26 (2007) 2074. [81] E.A. Medlycott, G.S. Hanan, Chem. Commun. (2007) 4884.
[45] T. Kajiwara, H. Tanaka, M. Yamashita, Pure Appl. Chem. 80 (2008) 2297. [82] L.M. Toma, R. Lescouëzec, D. Cangussu, R. Llusar, J. Mata, S. Spey, J.A. Thomas,
[46] T. Kajiwara, I. Watanabe, Y. Kaneko, S. Takaishi, M. Enomoto, N. Kojima, M. F. Lloret, M. Julve, Inorg. Chem. Commun. 8 (2005) 382.
Yamashita, J. Am. Chem. Soc. 129 (2007) 12360. [83] W. Liu, C.-F. Wang, Y.-Z. Li, J.-L. Zuo, X.-Z. You, Inorg. Chem. 45 (2006) 10058.
[47] A. Kamiyama, T. Noguchi, T. Kajiwara, T. Ito, Inorg. Chem. 41 (2002) 507. [84] H.-R. Wen, C.-F. Wang, Z.-Y. Du, J.-L. Zuo, Inorg. Chim. Acta 361 (2008)
[48] T. Kajiwara, T. Ito, Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 56 (2000) 2901.
22. [85] M.-L. Liu, L.-Z. Zhang, X.-P. Sun, Z.-P. Ma, W. Gu, X. Liu, J. Coord. Chem. 61
[49] R. Lescouëzec, J. Vaissermann, L.M. Toma, R. Carrasco, F. Lloret, M. Julve, Inorg. (2008) 2266.
Chem. 43 (2004) 2234. [86] R. Lescouezec, G. Marinescu, J. Vaissermann, F. Lloret, J. Faus, M. Andruh, M.
[50] T. Kajiwara, T. Ito, Angew. Chem. Int. Ed. 39 (2000) 230. Julve, Inorg. Chim. Acta 350 (2003) 131.
[51] L. Ouahab, F. Setifi, S. Golhen, T. Imakubo, R. Lescouezec, F. Lloret, M. Julve, R. [87] K.V. Gothelf, R.G. Hazell, K.A. Jorgensen, J. Am. Chem. Soc. 117 (1995) 4435.
Swietlik, C. R. Chimie 8 (2005) 1286. [88] T. Hintermann, D. Seebach, Helv. Chim. Acta (1998) 2093.
[52] H.-R. Wen, C.-F. Wang, J.-L. Zuo, Y. Song, X.-R. Zeng, X.-Z. You, Inorg. Chem. [89] D. Marcos, J.-V. Folgado, D. Beltrán-Porter, M.T.D. Prado-Gambardella, S.H.
45 (2006) 582. Pulcinelli, R.H. de Almeida-Santos, Polyhedron 9 (1990) 2699.
[53] H.-R. Wen, Y.-Z. Tang, C.-M. Liu, J.-L. Chen, C.-L. Yu, Inorg. Chem. 48 (2009) [90] T. Kajiwara, T. Ito, Dalton Trans. (1998) 3351.
10177. [91] M.G. Cowan, S. Brooker, Dalton Trans. 41 (2012) 1465.
[54] B.J. Hathaway, Struct. Bond. 57 (1984) 55. [92] R. Sahu, V. Manivannan, Inorg. Chim. Acta 363 (2010) 4008.
[55] D. Marcos, R. Martinez-Maňez, Inorg. Chim. Acta 159 (1989) 11. [93] A. Kamiyama, T. Noguchi, T. Kajiwara, T. Ito, CrystEngComm 5 (2003) 231.
[56] A. Cantarero, J.M. Amigo, J. Faus, M. Julve, T. Debaerdemaeker, Dalton Trans. [94] P. Paul, B. Tyagi, A.K. Bilakhiya, M.M. Bhadhade, E. Suresh, Dalton Trans. (1999)
(1988) 2033. 2009.
[57] I. Castro, J. Faus, M. Julve, J.M. Amigo, J. Sletten, T. Debaerdemaeker, Dalton [95] P. Paul, B. Tyagi, A.K. Bilakhiya, M.M. Bhadhade, E. Suresh, G. Ramachandraiah,
Trans. (1990) 891. Inorg. Chem. 37 (1998) 5733.
[58] G. Madariaga, F.-J. Zuniga, T. Rojo, Acta Crystallogr., Sect. C: Cryst. Struct. [96] H. Sugimoto, T. Takahira, T. Yoshimura, M. Shiro, M. Yamasaki, H. Miyake, K.
Commun. 47 (1991) 1632. Umakoshi, Y. Sasaki, Inorg. Chim. Acta 337 (2002) 203.
[59] J. Faus, M. Julve, J.M. Amigo, T. Debaerdemaeker, Dalton Trans. (1989) 1681. [97] G.R. Motson, J.S. Fleming, S. Brooker, Adv. Inorg. Chem. 55 (2004) 361.
[60] J.V. Folgado, E. Martínez-Tamayo, A. Beltrán-Porter, D. Beltrán-Porter, A. [98] P. Caravan, J.J. Ellison, T.J. McMurry, R.B. Lauffer, Chem. Rev. 99 (1999) 2293.
Fuertes, C. Miravitlles, Polyhedron 8 (1989) 1077. [99] M. Ferbinteanu, F. Cimpoesu, T. Kajiwara, M. Yamashita, Solid State Sci. 11
[61] J. Borrás, G. Alzuet, M. González-Álvarez, F. Estevan, B. Macías, M. Liu- (2009) 760.
González, A. Castiñeiras, Polyhedron 26 (2007) 5009. [100] M. Ferbinteanu, T. Kajiwara, F. Cimpoesu, K. Katagiri, M. Yamashita, Polyhe-
[62] Y.-Q. Zheng, W. Xu, F. Lin, G.-S. Fang, J. Coord. Chem. 59 (2006) 1825. dron 26 (2007) 2069.
[63] D. Cangussu, H.O. Stumpf, F. Lloret, M. Julve, V. Gonzalez, H. Adams, J.A. [101] F. Pointillart, K. Bernot, R. Sessoli, D. Gatteschi, Chem. Eur. J. 13 (2007) 1602.
Thomas, Inorg. Chim. Acta 358 (2005) 1113. [102] A.M. Madalan, K. Bernot, F. Pointillart, M. Andruh, A. Caneschi, Eur. J. Inorg.
[64] C. Yuste, D. Cangussu, H. Adams, J.A. Thomas, F. Lloret, M. Julve, Polyhedron Chem. (2007) 5533.
27 (2008) 2577. [103] W. Zhang, F. Zhao, S. Gao, Sci. China Ser. B-Chem. 50 (2007) 308.
[65] D. Cangussu, L.M. Toma, H.O. Stumpf, H. Adams, J.A. Thomas, F. Lloret, M. Julve, [104] M.G. Cowan, J. Olguín, S. Narayanaswamy, J.L. Tallon, S. Brooker, J. Am. Chem.
Polyhedron 27 (2008) 559. Soc. 134 (2012) 2892.
[66] B. Vangdal, J. Carranza, F. Lloret, M. Julve, J. Sletten, Dalton Trans. (2002) 566. [105] L. Carlucci, G. Ciani, S. Maggini, D.M. Proserpio, R. Sessoli, F. Totti, Inorg. Chim.
[67] P. Halder, E. Zangrando, T.K. Paine, Polyhedron 29 (2010) 434. Acta 376 (2011) 538.

You might also like