+ All Categories
Home > Documents > METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

Date post: 14-Dec-2016
Category:
Upload: guillem
View: 219 times
Download: 5 times
Share this document with a friend
34

Click here to load reader

Transcript
Page 1: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

This article was downloaded by: [University of Wyoming Libraries]On: 05 October 2013, At: 10:22Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH,UK

Comments on InorganicChemistry: A Journal of CriticalDiscussion of the CurrentLiteraturePublication details, including instructions forauthors and subscription information:http://www.tandfonline.com/loi/gcic20

METAL-BASED MOLECULARCHAINS: DESIGN BYCOORDINATION CHEMISTRYGuillem Aromí aa Departament de Quimica Inorgànica, Universitatde Barcelona, Barcelona, Spain

To cite this article: Guillem Aromí (2011) METAL-BASED MOLECULAR CHAINS:DESIGN BY COORDINATION CHEMISTRY, Comments on Inorganic Chemistry: AJournal of Critical Discussion of the Current Literature, 32:4, 163-194, DOI:10.1080/02603594.2011.642086

To link to this article: http://dx.doi.org/10.1080/02603594.2011.642086

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all theinformation (the “Content”) contained in the publications on our platform.However, Taylor & Francis, our agents, and our licensors make norepresentations or warranties whatsoever as to the accuracy, completeness,or suitability for any purpose of the Content. Any opinions and viewsexpressed in this publication are the opinions and views of the authors, andare not the views of or endorsed by Taylor & Francis. The accuracy of theContent should not be relied upon and should be independently verified withprimary sources of information. Taylor and Francis shall not be liable for anylosses, actions, claims, proceedings, demands, costs, expenses, damages,and other liabilities whatsoever or howsoever caused arising directly or

Page 2: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

indirectly in connection with, in relation to or arising out of the use of theContent.

This article may be used for research, teaching, and private study purposes.Any substantial or systematic reproduction, redistribution, reselling, loan,sub-licensing, systematic supply, or distribution in any form to anyone isexpressly forbidden. Terms & Conditions of access and use can be found athttp://www.tandfonline.com/page/terms-and-conditions

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 3: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

METAL-BASED MOLECULAR CHAINS: DESIGN

BY COORDINATION CHEMISTRY

GUILLEM AROMI

Departament de Quimica Inorganica, Universitat deBarcelona, Barcelona, Spain

Molecular extended metal atom chains (EMACs) or ‘‘molecular

chains’’ are discrete molecules exhibiting arrays of closely spaced

metals in a nearly straight fashion. Such species are gaining momen-

tum as possible crucial components of molecular functional devices.

The synthetic strategies for the preparation of such species have

diversified over the years. The existing strategies rely either on the

formation of ‘‘unsupported’’ metal-metal bonds or depend entirely

on the ability of specifically designed ligands to promote the assembly

of metals in desired topologies. In this review, a summary of the meth-

ods employed for the preparation of molecular EMACs is presented,

and several important examples are described.

Keywords: coordination chemistry, electric, magnetism conductivity,

metallic chains, molecular wires, spintronics

1. INTRODUCTION

For a long time, coordination chemists have used their synthetic skills for

the preparation of so-called ‘‘metallic molecular wires’’ or molecular

extended metal atom chains (EMACs). These species may be defined

as discrete assemblies of closely spaced metals in the form of linear

strings. Such systems have recently attracted considerable interest in

the context of nanotechnology.[1,2] Indeed, molecular metal chains are

Address correspondence to Guillem Aromı, Departament de Quimica Inorganica,

Universitat de Barcelona, Diagonal 647, Barcelona 08028, Spain. E-mail: guillem.aromi@

qi.ub.es

Comments on Inorganic Chemistry, 32: 163–194, 2011

Copyright # Taylor & Francis Group, LLC

ISSN: 0260-3594 print

DOI: 10.1080/02603594.2011.642086

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 4: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

promising candidates to constitute the connecting ‘‘wires’’ in future

molecular devices. In this respect, one of their advantages over organic-

based, p-extended systems is that the conducting metal core is embedded

inside an organic shell formed by ligands that could act as an insulator,

thus preventing eventual lateral current leakages.[3] Some advances have

been made in probing the potential of this category of molecules for

technological purposes; for instance, the electric conductivity of individ-

ual molecules of several aligned metals has been tested, as well as their

ability to act as on=off switches.[4] These functions depend on the nature

of the metals and their mode of interaction, the latter being a direct conse-

quence of the molecular structure imposed by the ligands responsible for

the assembly. Most of the current relevant systems involve ligands

that ensure the formation of metal-metal bonds by locating the metals

sufficiently close to each other. In such species, the electron transport

efficiency is directly related to the electron delocalization within the chain

and thus on the electronic structure of the metals and the nature of the

interactions between them.[5] Other ligands have the metals connected

by mono- or diatomic bridges or by longer spacers. This serves to modu-

late the nature of the (possible) electronic exchange between metals.

Other important properties can also be influenced by the ligands, such

as the redox properties of the metallic centers or the magnetic interaction

between them. Another category of discrete metallic chains rely on the

establishment of ‘‘unsupported’’ metal-metal bonds. We begin this survey

on molecular chains with a small summary of this last category of

molecules, and subsequently revise more extensively, in a non-exhaustive

manner, the various families of molecular EMACs resulting directly from

the ligand structure. An outlook with the future challenges faced by this

field of synthetic coordination chemistry is included at the end.

2. MOLECULAR CHAINS FEATURING ‘‘UNSUPPORTED’’

M���M BONDS

Metal-metal interactions may suffice on their own to stabilize molecular

chains of metals. An early example is the RhI=RhII cation [Rh4(dicp)8

Cl]5þ (dicp¼ 1,3�diisocyanopropane), consisting of two dimetallic

‘‘paddlewheel’’ units, each exhibiting four dicp bridges, and linked to

each other exclusively via a (formally) RhII�RhII metal bond (with a dis-

tance of 2.775 A).[6] The distance between metals within dinuclear units

(2.932 A) is consistent with that expected for a d7d8 interaction, with a

164 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 5: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

(formal) bond order of one half. In fact, the [Rh4] cations are linked to

each other by Cl� bridges, forming infinite cationic chains. A group of

related tetranuclear complexes is that of pairs of dinuclear metal units,

exhibiting two bridging carboxylates each. The M2 units are connected

to each other solely through unsupported interactions involving the

metals. This assembly has been featured with Cu(I) in the formation of

[Cu4(OBz)4(tic)4] (tic¼ p�tolylisocyanide),[7] Ru(I) with [Ru4(tfmb)4

(CO)10] (Htfmb¼ 3,5�bis�trifluoromethylbenzoic acid),[8] Rh(I,II) in

[Rh4(O2CH)4(bpy)4](PF6)2,[9] and Pt(II) as [Pt4(O2iPr)4(NH3)8](ClO4)4.[10]

The Cu4 compound relies on weak interactions between the metal of one

Cu2 unit and the isocyanide carbon atom of the neighboring pair. The Ru4

system exhibits an interdimer Ru � � �Ru contact 2.9065 A long, in line with

unsupported bonds between Ru(I) metals. The unbridged Rh � � �Rh bond

is 2.7797 A in length, which reflects the particular stability of the Rh46þ

core. The d8-d8 unsupported interaction of the Pt4 complex leads to an

interdimer metal�metal distance of 3.2619 A. Pyridonate (Opy) has

also served as a bridging ligand within dinuclear metal units that in turn

ensemble into molecular chains through metal�metal bonds. Thus, the

dinuclear complex [Ir2(Opy)2(CO)4] was found to be, in solution, an equi-

librium between the head-to-head (HH) and the head-to-tail (HT) con-

figuration. Oxidation of this solution with I2 produces the mixed valence

IrI2IrII

2 chains HH,HH�[Ir4(Opy)4I2(CO)8] and HH,HT�[Ir4(Opy)4I2

(CO)8], the former being the thermodynamic product and the latter being

the kinetic one. Control of the reaction thermal conditions allows isolating

either one or the other, so that each of them has been crystallographically

characterized.[11] More remarkably, the use of a deficient amount of

oxidant produces a hexanuclear IrI4IrII

2 chain (average oxidation state

Figure 1. Representation of complex [Pt8(amd)8(NH3)16](NO3)10 (amd¼ acetamidate). Only

H atoms on the N-carriers of amd are represented. dPt���Pt (most ext.)¼ 2.880 and 2.900 A,

dNi���Ni (sec. ext.)¼ 2.900 and 2.778 A, dNi���Ni (cent.)¼ 2.934 A. (Color figure available online.)

METAL-BASED MOLECULAR CHAINS 165

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 6: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

of þ1.33) with formula HH,HT,HH�[Ir6(Opy)6I2(CO)12].[12] Several

related tetranuclear complexes involving pyridonate and other similar

bridging donors also exist with platinum (belonging to the group of

so-called ‘‘platinum blue’’ compounds), such as the complex salt contain-

ing the cation [Pt4(Opy)4(NH3)8]5þ,[13] which features the oxidation state

sequence PtII3PtIII. An impressive octanuclear string of this family of

complexes was isolated and characterized using acetamidate (amd) as a

bridging ligand; [Pt8(amd)8(NH3)16](NO3)10 (Figure 1).[14] In this complex,

the Pt atoms are formally in the oxidation state þ2.25.

The complex [Rh4(s-pqdi)2(pqdi)4(CO)4]2þ (Figure 2; pqdi¼ 9,10-

phenanthroquinonediimine; Scheme 1),[15] and its Ir analogue (both

metals in the oxidation state þ1, d8) exhibit so-called d8-d8 metal-metal

bonds, which are supported in a synergistic manner by p-p interactions

between the aromatic ligand. The stacking is in fact believed to occur

Figure 2. Representation of complex cation [Rh4(s-pqdi)2(pqdi)4(CO)4]2þ (pqdi¼ 9,10-

phenanthroquinonediimine). H atoms not shown. dRh���Rh (ext.)¼ 2.848 A, dRh���Rh (cent.)¼2.858 A. (Color figure available online.)

Scheme 1. 9,10-phenanthrosemiquinonediimine (s-pqdi) and 9,10-phenanthroquinonedii-

mine (pqdi).

166 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 7: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

between the p orbitals of the five-member chelate rings, which are

disposed in an eclipsed manner to facilitate this interaction.

Gold(I) to gold(II) electronic donation (d10-d9) is proposed to be the

driving force leading to the penta- and hexanuclear chains of this metal

with formulae [(RAu(CH2PPh2CH2)2Au)2AuR2]ClO4 and [(RAu(CH2

PPh2CH2)2Au)2(Au(CH2PPh2CH2)2Au)](ClO4)2 (R¼C6F5 and

C6F3H2), holding together as unsupported Au�Au bonds dinuclear and

mononuclear building blocks.[16] These compounds were obtained by

reacting Au(II) dinuclear precursors of the type [Au2(CH2PPh2CH2)2]2þ

with Au(I) mononuclear or dinuclear units, respectively.

3. METALLIC MOLECULAR CHAINS MADE WITH

DESIGNED POLYDENTATE LIGANDS

The distinctive feature of the various classes reviewed in this section will

be the type of ligands employed.

3.1. Polypyridylamine-Based Metal Chains

The group of discrete 1D metal arrays wrapped by four oligo�a�pyridy-

lamine type ligands (Scheme 2) certainly constitutes one of the most

extensive, varied, and studied classes of molecular wires. Indeed, a very

large number of molecules (around 260) of nuclearity ranging from three

to eleven have been prepared and studied for a large group of metals (Cr,

Co, Ni, Cu, Ru, Rh, Pd, Pt), of which trinuclear complexes represent

71%. The latter have been of paramount importance for understanding

the electronic structure within this family of metal chains and the metal-

metal bonds that exist within most of them.[17] The heptanuclear complex

[Ni7(teptra)4Cl2] (Figure 3, H3teptra¼ tetrapyridyltriamine) will serve to

describe the main structural features in this family of compounds.[18] This

complex features a central core of seven Ni(II) ions almost perfectly

aligned and wrapped helically by four teptra3� ligands in a syn-syn-syn-

syn-syn-syn configuration. Such arrangement ensures the alternative

Scheme 2. Oligo–pyridylamine type ligands.

METAL-BASED MOLECULAR CHAINS 167

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 8: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

presence of either four pyridyl or four amido N-donors around the met-

allic string, resulting on square planar environments, except for the two

most external ions, which exhibit apical Cl� ligands and thus square pyr-

amidal geometries. Such an arrangement locates adjacent Ni(II) ions at

an average distance of 2.302 A, including the third shortest distance ever

observed on any Ni complex, 2.214 A (only seen shorter on a Ni7 and a

Ni6 complex of the same family).[19,20] The electronic configuration of

the 3d orbitals in the [Nix] wires leads in general to only weak metal-metal

interactions, which is detrimental to the conductive properties of the mol-

ecular wire. Indeed, conductance studies using STM (scanning tunneling

microscopy) show that analogous [M(II)x] molecular chains exhibit grow-

ing conductivity along the sequences Ni(II), Co(II), and Cr(II), resulting

from increasing metal-metal bond orders of 0, 0.5, and 1, respectively.[4]

Modifications of the backbone in the oligo�a�pyridylamine ligand have

allowed the preparation of long-chain derivatives more resistant to oxi-

dation and versatile. For example, replacement of one pyridine group

by pyrazine within the ligand (H4mpz; Scheme 3, top) has led to the non-

achromium complexes, [Cr9(mpz)4(X)2] (X�¼Cl�, NCS�).[21] These two

Scheme 3. H4mpz and H2napany.

Figure 3. Two perpendicular views of complex [Ni7(teptra)4Cl2] (H3teptra¼ tetrapyridyl-

triamine). H atoms not shown. dNi���Ni (most ext.)¼ 2.383 and 2.374 A, dPt���Pt (sec. ext.)¼ 2.304

2.304 and 2.310 A, dPt���Pt (cent.)

¼ 2.214 and 2.226 A. (Color figure available online.)

168 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 9: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

chain compounds are 23 and 25 A long, respectively, and exhibit two

types of Cr�Cr distances; one corresponds to quadruple metal-metal

bonds (ranging 1.973 to 2.097 A, with four such contacts per molecule)

and another within the 2.397 to 2.497 A range, which might involve a weak

Cr�Cr bond of r-character. Another variation to the a�pyridylamine

sequence is, for example, the introduction of naphthyridine units within

the ligand backbone (see H2napany, Scheme 3, bottom). This has allowed

us to stabilize (Ni2)3þ fragments within the molecular chain, such as in

[Ni9(napany)4Cl](BF4)2.[5] Such metal pairs contain Ni(I) and Ni(II)

and feature a high level of delocalization, such that in the ground state

these fragments are carriers of an S¼ 3=2 spin magnetic moment. These

properties are likely to confer enhanced electron conductivity to the

chain, and have been observed for chains with other nuclearities, such

as [Ni5][22] and [Ni9].[23] The longest chains of this impressive family exhi-

bit eleven Ni ions and have the formula [Ni11(tentra)4X2](PF6)4 (X¼Cl�

or SCN�; H3tentra¼ tetranaphthyridyltriamine).[24]

3.2. Metal Chains Sandwiched within Conjugated

pp-extended Ligands

A fascinating family of molecular metallic chains is made by aggregates of

metal atoms sandwiched by the pp backbone of extended aromatic ligands.

In this class of linear organometallic complexes, the shape of the chain

may vary, depending on the structure of the organic ligand. Thus the first

complex of this kind was prepared from the simple sp2 carbon atom chain

all�trans�1,8�diphenyl�1,3,5,7�octatetraene (dpot, Scheme 4), leading

to the highly straight tetranuclear complex [Pd4(dpot)2]X2 (X¼various anions), which could also be made with two pyridine ligands,

Scheme 4. Various polyaryl-polyene ligands.

METAL-BASED MOLECULAR CHAINS 169

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 10: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

bound at end metals of the chain; [Pd4(dpot)2(py)2]X2 (Figure 4, top).[25]

Binding of pyridine causes a slippage of the dpot ligand along the direction

of the Pd4 chain, going from the g3:g2:g2:g3 coordination mode to

g2:g2:g2:g2. The electronic structure of a [Pdn]2þ metal chain is described

as resulting from 1 Pd(II)þ (n� 1) Pd(0), thus leaving 2(n� 1) electrons

for the establishment of n� 1 metal-metal bonds (six electrons for the case

of Pd4). In fact, a theoretical study demonstrates that the resulting 2þcharge is delocalized over the metallic chains as well as both polyenes of

the cationic assembly, providing the system with high stability.[26] The

linear pentanuclear chain [Pd5(dpdh)2]2þ was obtained from the longer

ligand all�trans�1,12�diphenyl�1,3,5,7,9,11�dodecahexaene (dpdh).

As a remarkable extension of this work, it was demonstrated that the poly-

cyclic aromatic compound perylene (pery) could also serve to sandwich

linear Pd4 chains in form of complex cations such as [Pd4(pery)2(py)2]2þ,

acting in the l�g2:g2:g2:g2 coordination mode.[27] The electronic structure

Figure 4. Representation of the complex cations (top) [Pd4(dpot)2(py)2]2þ (dpot¼ all�trans�1,8�diphenyl�1,3,5,7�octatetraene, dPd���Pd (ext.)¼ 2.746 A, dPd���Pd (cent.)¼2.721 A), and (bottom) [Pd5(p-bpbb)2(py)2]2þ (p-bpbb¼ 1,4�bis(4�phenyl�1,3�butadie-

nyl)benzene, dPd���Pd (ext.)¼ 2.733 and 2.722 A, dPd���Pd (cent.)¼ 2.721 and 2.713 A). (Color

figure available online.)

170 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 11: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

of the metal skeleton in this compound is the same as in the analogous

tetrametallic chain formed with the linear polyene dpot and, in fact, the

latter was shown to displace completely the perylene as ligand to form

quantitatively the corresponding tetranuclear chain of Figure 4. The bis-

perylene sandwich complex was then used as starting material in reactions

with other ligands to obtain new EMACs with unprecedented structures.

This was the case for ligands featuring a central ortho� or para�phenylene

inserted within a polyene chain, such as 1,2�bis(4�phenyl�1,3�butadie-

nyl)benzene (o�bpbb) or 1,4�bis(4�phenyl�1,3�butadienyl)benzene

(p�bpbb) (Figure 4, bottom), which exhibit a bent configuration

(Scheme 4). These reactions conduce to the formation of [Pd4] and

[Pd5] sandwiched metal chains with an arched and an angular topology,

respectively, demonstrating that the shape and topology of the polymetallic

chains may be controlled by the structure of the carbon backbone of the psystem, which plays the role of a template. A surprising recent addition to

this successful research program is the demonstration that the mixed-

valence chain [Pd4(dpot)2]2þ may be reversibly reduced by two electrons

with two equivalents of Cp2Co to the corresponding neutral [Pd�4] ana-

log.[28] The structure of this species could be obtained using the ligandtBu-dpot (a derivative of dpot with tert-butyl substituents on the aryl

groups), and revealed two different hapticities for both ligands in the

molecule; l�g3:g2:g2:g3 and l�g2:g2:g2:g2, respectively.

3.3. Trinuclear Linear Complexes with Pyrazole, Triazole,

or Tetrazole Ligands

All five-membered heterocycles derived from pyrazole, triazole, or tetra-

zole moieties (Scheme 5) share the presence of two bonded nitrogen

atoms [�N�N�], which constitutes a l:g2 bridging moiety very common

in coordination chemistry. These rings have originated a large family of

(more than 80 structurally characterized) trinuclear linear metal clusters

exhibiting either triple (class I; �66%) or double (class II; �14%) [M�N�N�M] bridges between metals, or two such bridges combined with

a third monoatomic bridge (class III; �20%). These compounds are

Scheme 5. Five membered heterocyclic azole-type ligands.

METAL-BASED MOLECULAR CHAINS 171

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 12: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

structurally related to their (extremely abundant) dinuclear counterparts.

Of these, the majority are made with 1,2,3� or 1,2,4�triazoles (around 60

examples), and only one case with a tetrazole[29] was found on the CSD

database (version 5.32, May 2011 update). The rest are made with

pyrazole derivatives. Complexes representative for classes I, II, and III

are, respectively, [Ni3(Htz)6(H2O)3](NO3)6 (Htz¼ 1,2,4�triazole),[30]

[Co3(AcO)2(dmpz)4(Hdmpz)2] (Hdmpz¼ 3,5�dimethylpyrazole)[31] and

[Co3F2(SCN)4(detz)6] (Hdetz¼ 3,5�diethyl�1,2,4�triazole),[32] shown

in Figure 5.

Class I complexes may be described as two fused [M2(azole)3] units

sharing the central metal of the complex. Each metal pair is bridged by

Figure 5. Representation of complexes (top) [Ni3(Htz)6(H2O)3](NO3)6 (Htz¼ 1,2,4�triazole, dNi���Ni¼ 3.737 A), (middle) [Co3(AcO)2(dmpz)4(Hdmpz)2] (Hdmpz¼ 3,5�dimethyl-

pyrazole, dCo���Co¼ 3.373 A) and (bottom) [Co3F2(SCN)4(detz)6] (Hdetz¼ 3,5�diethyl�1,2,4�triazole, dCo���Co¼ 3.609 and 3.601 A). H atoms not shown. (Color figure available

online.)

172 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 13: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

three heterocycles, their respective idealized planes being disposed 120�

from each other, in the form of a ‘‘paddlewheel.’’ The three ligands of

each [M2] unit are oriented in a mutually staggered manner. This causes

the central metal to adopt an octahedral coordination geometry. In

general, the external metal ions of the chain exhibit also octahedral

geometry. The three rings of the ‘‘paddlewheel’’ leave on each of the

outer metals three free positions for coordination that are usually

completed by solvent monodentate ligands (Figure 5; top),[30] anionic

groups,[33] the corresponding azole group acting as neutral terminal

ligand or additional groups attached to the azole five-membered ring.[34]

It is interesting to note that only the dinuclear and trinuclear analogues

of fused ‘‘paddlewheel’’ oligomers exist, and not longer discrete members

of the series. By contrast, several 1D [Cu(azole)3]n polymers have been

structurally characterized[35] and the corresponding [Fe]n derivative,

whose structure has been inferred by other methods, exhibits very inter-

esting spin crossover behavior.[36] Almost all complexes from this class

are homometallic (featuring mainly Ni, Fe, Co, Cu, Cd, and to a lesser

extent Mn and Zn). Most of them have been studied for their magnetic

properties.[37–39] The intramolecular exchange between paramagnetic

metal centers as mediated through the three [�N�N�] pathways was

always found to be antiferromagnetic. Many of the (numerous) Fe

analogs display interesting spin-crossover behavior.[40] A recent review

summarizes the properties of most trinuclear complexes of class I.[41]

A rare exception to the homometallic nature of these complexes is the

[ReIMnIIReI] compound (NHEt3)2[MnRe2(pz)6(CO)6].[42] A beautiful

molecule related to the latter is the tetranuclear EMAC with formula

(NHEt3)2[Mn2Re2O3(pz)6(CO)6] (Figure 6, Hpz¼ pyrazole), which

Figure 6. Representation of the complex anion [Mn2Re2O3(pz)6(CO)6]2� (Hpz¼pyrazole).

H atoms not shown. dRe���Mn¼ 3.897 A, dMn���Mn¼ 2.394 A. (Color figure available online.)

METAL-BASED MOLECULAR CHAINS 173

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 14: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

exhibits a ReIMnIVMnIVReI sequence, with the central Mn centers

bridged by two l�O2� ligands and the Re � � �Mn pairs linked by triple

l�pz� bridges. This complex is obtained from the heterometallic dinuc-

lear [MnIReI] analogue (NHEt3)[MnRe(CO)6], by reaction with O2.

Class II trinuclear complexes are made with pyrazole derivatives and

exhibit double l1,2�N,N�azole bridges between metals (Figure 5, mid-

dle). This group of complexes is less numerous than Class I, but it is much

more versatile. Metals involved span Ni, Co, Cu, Zn, Rh, Pd, and Ag and

some cases are heterometallic (such as complexes [PdIIPdIICoII][31] and

[RhIPdIIRhI]).[43] The central metal may be either in tetrahedral (CoII,

ZnII, AgII) or square planar (NiII, PdII) coordination. In the first case,

the cyclic ligands within each metal pair are approximately coplanar,

and approximately perpendicular to the rings from the adjacent M2 unit

(Figure 5, middle). When square planar, both pyrazole groups within each

M2 pair form a ‘‘butterfly’’ shape, with an angle (approximately between

60 and 70�). The ‘‘butterflies’’ from both M2 pairs point to opposite direc-

tions. The coordination geometry of the external metals in this class of

complexes comprises tetrahedral, square planar, and triangular. One

related complex made with pyrazole-based compartmental ligands

together with another pyrazole derivative is a [Cu4], which comprises

three fused [Cu2(azole)2] moieties yielding a tetranuclear chain.[44] It is

clear from the coordination geometries observed and the metals involved

in the compounds just reviewed that when the cations exhibit preferen-

tially coordination number six, Class I complexes may be expected. If

the metals favor square planar or tetrahedral geometry, the complexes

obtained will be of Class II. Only metals with very versatile coordination

geometries (such as ZnII or CoII) may exhibit either of both Class I and II

configurations, as well as metals with two strong preferences (e.g., NiII,

which exhibits stable complexes in an octahedral environment, but also

as square planar centers).

In Class III complexes (Figure 5, bottom), the bridging moiety

between metals is heteroleptic; it contains two l1,2�N,N�azole fragments

together with a monoatomic bridge (F�, Cl�, OH�, or NCS�). The struc-

ture and properties of this type of complexes have been recently summar-

ized.[41] The majority are made with CuII, while examples with Ni, Co,

Mn, and Cd also exist. The presence of a different ligand, other than

the azole groups, adds versatility to the magnetic exchange between

metals within the chain. For example, it was demonstrated that the mag-

netic coupling may be tuned by selectively changing the variable ligand

174 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 15: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

X�. Thus, in going from NCS� to F�, the coupling becomes more

antiferromagnetic (as seen for the Ni ant the Co cases).[32] Also, the effect

of the second (variable) ligand has allowed us to see ferromagnetic-type

couplings and therefore high-spin molecules.[45]

3.4. Use of Diazines to Link Chains of Metals

Cyclic diazines, such as pyridazine (pdz), could act in many ways as the

azole ligands discussed in the previous section. As it turns out, there is

not an analogous extensive family of complexes such as that discussed

above; however, there exists a small group of trinuclear complexes with

pyridazine that exhibit the same type of structure as complexes of Class

III of section 2.3, where essentially the five-membered ring is replaced

by pdz. The metals involved are NiII,[46] CuII,[47] and CoII,[48] and the

coordination geometry in all the metals is octahedral. Susceptibility mea-

surements on the complex [Ni3(pdz)6(NCS)6] unveil a quasi Curie law

behavior. This is interpreted as the result of the almost exact compen-

sation of the antiferromagnetic coupling mediated by the diazine bridges

by the ferromagnetic interaction induced by end-on l�NCS� groups.[46]

On the other hand, the comparison between complexes [Co3(pdz)6

(OH)2(NO3)2(H2O)2] and [Co3(pdz)6(NCS)6] is very interesting. The

difference of variable bridging ligand (OH� vs NCS�) causes the central

Co atom to be in a different spin state. Thus, for the OH� complex, the

central metal is in the low-spin state (S¼ 1=2) with the external Co

ions being high-spin (S¼ 3=2), whereas the complex with bridging NCS�

Scheme 6. Trinuclear metal complexes formed with diazine-type ligands.

METAL-BASED MOLECULAR CHAINS 175

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 16: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

exhibits high-spin states at the three metal centers. This is corroborated

crystallographically and by magnetic susceptibility measurements.

A beautiful series of triple stranded helicates hosting [M4] metal

chains is obtained from ligands combining a pyridazine ring with two

non-cyclic (flexible) azine groups, in addition to pyridyl groups at both

ends (L; Scheme 6). Thus complexes with CuII,[49] MnII,[50] and NiII[51]

have been prepared, exhibiting the formula [M4L3](ClO4)n (n¼ 2 or 8,

depending on whether the ligand is doubly deprotonated or not,

Figure 7). In general, the magnetic coupling between metals, as mediated

through triple l�NN diazine groups, is weakly antiferromagnetic and

modulated by the M�N�N�M torsion angles. In the MnII complex,

the open-chain diazine moieties seem to facilitate weakly ferromagnetic

interactions. The family of coordination chemistry metallohelicates is

relatively large;[52] however, unlike the above examples, the vast majority

of systems locate the metals too far away from each other for these centers

to exhibit any significant magnetic or electronic interaction.

Another important group of trinuclear linear complexes linked by

diazine groups is obtained with R�N�N�R0 ligands where R and R0 con-

tain donors capable of forming chelates together with the N-atom to

which they are attached. Thus, with divalent ions (mainly CuII[53] and

NiII[54] with some examples of CoII,[54] ZnII,[55] and PdII[56]), two such

ligands chelate a central metal with one N-atom of the azine and then each

ligand binds another metal through the other N-atom, forming trinuclear

structures, usually with an overall flat structure (Scheme 6).[57] If the

metals involved are trivalent (MIII), instead of linear structures, these

ligands usually form metalladiazamacrocycles of various nuclearities.[57]

Figure 7. Representation of the complex anion [Mn4L3]8þ. H atoms not shown. dMn���Mn

(ext)¼ 3.802 A, dMn���Mn (cent)¼ 3.850 A. (Color figure available online.)

176 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 17: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

3.5. Ferromagnetic Metal Chains with Oligo�m�phenyleneoxalamide Ligands

A series of ðCuIInÞ (n¼ 2–4) linear metal arrays has been synthesized with

a class of polytopic ligands containing two outer oxamato groups and x

(x¼ 0, 1, 2) inner oxamidato moieties, the various donor units being sepa-

rated by m�phenylene spacers (Scheme 8).[58] These wires exhibit the

well-established ferromagnetic interaction observed between CuII centers

separated by m�phenylene groups, as mediated by a spin polarization

mechanism, with coupling constants spanning þ15 to þ17 cm�1 (in the

H¼�JSiSj convention).[59] These linear species are described as double

stranded anionic complexes of 2, 3, or 4 metals (Figure 8), respectively

(and charges�4, �6 and �8), each metal being located within two chelat-

ing pockets of two ligands disposed in front of each other. The helical

chirality around the metals alternates between M and P thus providing

for an overall meso-helicate-type architecture. The configuration of the

ligands is all-syn, of the oxamidato or oxamato groups with respect to

the phenylene spacer, both types of planes forming angles with each other

that span 62 to 82 degrees. As a result of the ferromagnetic coupling

between adjacent metals, these [Cun] arrays exhibit maximum possible

molecular spin (S¼ n=2). In addition, the CuII centers of the chains

can be sequentially and reversibly oxidized to diamagnetic CuIII, thus

providing for a mechanism to externally interrupting the ferromagnetic

interaction along the chain and thus switching ON and OFF the electron-

exchange throughout the wire.

Scheme 8. Bis-oxamato-polyoxamidato ligands.

Scheme 7. A bis-diazine-pyridazine ligand.

METAL-BASED MOLECULAR CHAINS 177

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 18: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

3.6. Phosphine-Derived Ligands

The tri-phosphine ligand (diphenylphosphinomethyl)phenylphosphine

(dpmp, Scheme 9) exhibits quite versatile coordination chemistry. The

reaction of dinuclear [Pt(I)2] precursors [Pt2(dpmp)2(RNC)2]2þ (RNC¼isocyanide ligands) with mononuclear compounds of d10 M(0) centers

(M¼Pd, Pt) produces the corresponding PtPtM linear homo- and het-

erometallic chains with formulae [Pt2M(dpmp)2(RNC)2]2þ.[60] In these

compounds, the electronic structure was described as d9�d10�d9, follow-

ing an electron transfer from the external atom joining the cluster to the

central atom of the final assembly, leading to two single metal-metal r�bonds within the molecule. The reactivity of the [Pt(I)2] precursor was

also investigated with [MCl(cod)]2 species (M¼Rh(I), Ir(I); d8 metals),

which produced the corresponding linear heterometallic assemblies with

formula [Pt2MCl(dpmp)2(RNC)2].[61] The bonding within the backbone

of these chains occurs through one covalent (d9�d9) Pt�Pt bond and

one dative (d8!d9) M�Pt interaction. Reactivity studies of the Pt2Rh

Figure 8. Representation of the complex anion [Cu4L2]8þ (H8L¼ 2�methyl�1,3�phenylenebis�[N0�(2�methyl�3�phenylamine)oxamide]). H atoms not shown. dCu���Cu

(ext)¼ 7.294 A, dCu���Cu (cent)¼ 7.395 A. (Color figure available online.)

Scheme 9. Various posphine-type ligands.

178 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 19: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

system suggest that in this assembly, the Pt2 unit exhibits nucleophilic

character and the Rh center is electrophilic. In the presence of NaBH4

or NaOMe, the Pt2M species (M¼Pt, Pd) react to form hydride bridged

hexanuclear species [Pt4M2(l�H)(dpmp)4(RNC)2]3þ (Figure 9).[62] The

bridge between both trinuclear entities in these compounds features a

MHM three-center, two-electron unit. These species may be chemically

or electrochemically oxidized, leading to the linear chains with a Pt�Pt�M�M�Pt�Pt sequence, exhibiting Pt�Pt and Pt�M bonds, as a

result of a migration of bonding electrons from the Pt�M bonds to the

central M�M unit.

The use of a ligand analogous to dpmp, with a central As donor instead

of P (bis-(diphenylphosphinomethyl)phenylarsine (bdppa, Scheme 9),

allows the preparation of heterotrinuclear metal chains of the type MM0M,

by selectively locating at the center of the assembly a metal distinctly differ-

ent from the end metals. In this manner, a large variety of combinations

have been obtained and studied, with a great diversity of electronic config-

urations such as Ir2M (d8s2d8; M¼Pb(II), Sn(II), Ge(II), In(I), Tl(I),

Sb(III), Bi(III)),[63] Ir2M0 (d8d10d8; M0 ¼Cu(I), Ag(I), Au(I)),[64] Ir2M00

(d8d8d8; M00 ¼Pd(II), Rh(I), Ir(I)),[65] etc., the majority of which also

obtained with Rh(I) as external metals.

The phosphine ligand Hpyphos (6�diphenylphosphino�2�pyri-

done; Scheme 9) also contains donor atoms of different types (O,N,P),

aligned for the assembly of linear metal chains. These properties have

Figure 9. Representation of the complex cation [Pt4Pd2(m�H)(dpmp)4(RNC)2]3þ (dpmp¼(diphenylphosphinomethyl)phenylphosphine; RNC¼ isocyanide ligands. H atoms and H�

not shown. dPt���Pt¼ 2.723 and 2.714 A, dPt���Pd¼ 2.749 and 2.750 A, dPd���Pd¼ 3.355 A.

(Color figure available online.)

METAL-BASED MOLECULAR CHAINS 179

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 20: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

been employed for the formation of heterometallic tetranuclear arrays,

prepared in a sequential manner. This was done by obtaining first quad-

ruple bonded Mo2 complexes surrounded by four pyphos� ligands in a

‘‘paddlewheel’’ arrangement; [Mo2(pyphos)4]. The steric demands of

the ligand cause the P donors of different ligands to be oriented, pairwise,

to opposite sides of the Mo�Mo axis. This provides for two suitably dis-

posed axial sites on each molecular end, well tailored for coordination to

two additional, late-transition metal centers in the formation of [Mo2M2]

linear clusters (M¼PdII, PtII, IrI, RhI), of the type [Mo2M2(pyphos)4X4]

(X¼Cl, Br, I) for Pt or Pd,[66] and [Mo2M2(pyphos)4(RNC)4]Cl2 for Ir

and Rh.[67] Following the electronic configuration of the metallic back-

bone, none of these compounds exhibit M�Mo interactions. Interest-

ingly, reduction of the MII (Pd, Pt) centers leads to the formation of

Mo�M bonds with concomitant decrease of the Mo�Mo interaction

bond order (from 4 to 3) to produce chains exhibiting a fully bonded

[M�Mo�Mo�M] axis within the new complexes [Mo2M2(pyphos)4X4]

(M¼Pd, Pt; X¼Cl, Br, I). Likewise, oxidation of the MI metals (IR,

Rh) causes a very similar effect, by facilitating the interaction of the relectrons of the Mo2 core with the dz orbitals of the resulting MII (d7)

centers, producing the complexes, with a fully metal-metal bonded tetra-

metallic chain, [Mo2M2(pyphos)4X2(RNC)4]2þ (M¼ Ir, Rh; X¼Cl, Br, I;

Figure 10).

Figure 10. Representation of the complex cation [Mo2Rh2(pyphos)4Cl2(RNC)4]2þ

(Hpyphos¼ 6�diphenylphosphino�2�pyridone). H atoms not shown. dRh���Mo¼ 2.731 A,

dMo���Mo¼ 2.124 A. (Color figure available online.)

180 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 21: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

3.7. Homo and Heterometallic Chains with m�phenylene

Spaced Bis�b�diketones

A family of molecular linear chains of metallic cations has been synthe-

sized with ligands consisting of two b�diketone groups separated by a

1,3�phenylene�type moiety (derived from benzene, pyridine or phenol,

Scheme 10) in combination with other donor groups. 1,3�Diketonates

are very good chelating functions. The family of bis�b�diketones is

prone to form dinuclear complexes. If the metals are 3d trivalent (FeIII,

MnIII, etc.) these complexes are described as triple-stranded helicates

(or mesocates),[68,69] whereas with MII cations two bis�b�diketonate

ligands sandwich two metals and their (axial) coordination sites are com-

pleted by solvent molecules.[70,71] In this family of ligands, the diketone

groups are accompanied by a combination of phenol and=or pyridine-

type groups, disposed in a linear fashion, thus favoring the assembly of

coordination metal chains (Scheme 10). The preparation of these ligands

involves well-established reactions in organic chemistry synthesis and

always goes through one key step, the Claisen condensation between

one ester and one ketone, which yields a 1,3�diketone.[72,73] In this case,

Scheme 10. Bis-beta-diketone ligands.

METAL-BASED MOLECULAR CHAINS 181

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 22: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

two such condensations occur simultaneously in one molecule, producing

the corresponding bis�1,3�diketone.

Ligand H3L1 (Scheme 10) displays two b�diketones separated by a

phenol group, thus disposing five regularly spaced oxygen atoms in a

row, in a good arrangement to favor the assembly of coordination metal

chains. It was found that this ligand reacts with M(AcO)2 salts (MII¼CoII,

MnII), leading to trinuclear asymmetric linear chains, [M3(HL1)3],

described as triple stranded pseudo helicates (Figure 11).[74,75] The

chirality of the metals along the chain alternates between K and D(Scheme 11); thus both sequences KDK and DKD are found within the

crystal by virtue of the inversion center present within these systems’ sym-

metry space group. Therefore, the helices invert their sense of rotation

along the molecular axis. In these complexes, three metals are located

within three coordination pockets from each of the twice deprotonated

ligands, displaying a quasi trigonal prismatic coordination geometry. The

fourth potential coordination pocket of each ligand is, in fact, filled by

the proton remaining on each of them, which leads to the formation of

intramolecular hydrogen bonds. The bulk magnetization properties of

these complexes have both been modeled as a pair of coupled metals next

to a third isolated ion. For the Mn(II) system, the coupled pair exhibits a

coupling constant of �2.75 cm�1 (in the convention H¼�2JS1S2). The

complex of cobalt was modeled considering three isolated Co(II) ions sub-

ject to a considerable spin-orbit coupling effect, experiencing the effect of

magnetic coupling at low temperatures, which was treated as a pertur-

bation in form of interaction between effective S ¼ 12

centers, leading to

a coupling constant J ¼�4.9 cm�1 in the best fit.[75] One interesting

Figure 11. Representation of the complex [Co4(L3)2(py)6]. H atoms not shown. dCo���Co

(ext)¼ 3.160 A, dCo���Co (cent)¼ 6. 962 A. (Color figure available online.)

182 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 23: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

feature of the [M3(HL1)3] trinuclear chains is that they dissolve and react

in coordinating solvents, S, transforming into the corresponding solvate

dinuclear complex [M2(HL1)2(S)2] (Figure 11). This transformation is

reversible, thus isolated crystals of the dinuclear species dissolve into

non-coordinating solvents to generate the corresponding [M3(HL1)3]

compound.[75,76]

If in place of a central phenol spacer as in H3L1, there is a pyridyl

group, as in the H2RL4 type ligands (Scheme 11), the donor exhibits a

central larger coordination pocket instead of two smaller ones. These

unique structural properties have allowed the preparation of a remarkable

series of more than 30 complexes with a [M2Ln(MeL4)2]3þ backbone

(MII¼CuII, NiII; LnIII¼ all lanthanoids except Pm; Scheme 12 and

Figure 12).[77,78] In these 3d�4f heterometallic complexes, the axial sites

of square pyramidal Cu(II) or octahedral Ni(II) are occupied by solvent

molecules (DMF, MeOH, or H2O). The central Ln(III) ion, besides hav-

ing six coordination sites occupied by donors from two MeL42� ligands

(forming a hexagonal equatorial plane), is bound to three No �3 ligands,

completing coordination numbers of 10 or 12 (the latter only seen for

the [Cu2La] complex). This incredible synthetic and crystallographic

work allowed a systematic and quasi comprehensive study of the nature

of the magnetic interaction between lanthanides and Cu(II) or Ni(II).

Scheme 12. Schematic structure of [MLaM] complexes with ligand H2MeL4. (Color figure

available online.)

Scheme 11. Stereochemistry within [M3(HL1)3] complexes.

METAL-BASED MOLECULAR CHAINS 183

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 24: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

This was done by subtracting to the variable temperature molar suscepti-

bility of the [MLnM] complexes, the contribution from a [MLaM] (LaIII

being diamagnetic) and that from the corresponding [ZnLnZn] complex.

The resulting DvMT value (vM being the molar paramagnetic suscepti-

bility) is the contribution from the Ln � � �M coupling. If vMT< 0, the

coupling is antiferromagnetic; otherwise, it is ferromagnetic. These

studies have confirmed that for such systems the coupling with Cu(II)

is antiferromagnetic for Ce, Pr, Nd, and Sm and ferromagnetic for Gd,

Tb, Dy, Ho, and Er. For the case of Ni(II), the coupling was found to

be antiferromagnetic for Ce, Pr, and Nd, whereas it is ferromagnetic for

Gd, Tb, Dy, Ho, and Er. Interestingly, this extensive work allows us to

infer that larger Ln(III) ions tend to favor antiferromagnetic interactions

with 3d metals, while smaller cations seem to facilitate ferromagnetic

exchange. The equivalent Gd(III) and La(III) systems have also been

prepared with Co(II),[79] whereas with Fe(III), instead of two ligands

sandwiching the metal chain, a total of three ligands wrap the molecular

axis in a pseudo helical manner, while non-coordinated [FeCl4]� counter

ions compensate the three positive charges of the complex.[80]

Ligand H5L2 is related to H3L1, but includes two additional phenol

groups located at both ends of the ligand backbone, thus featuring an

array of six adjacent coordination pockets (or seven aligned oxygen

donors). This multidentate ligand was synthesized with the aim of

promoting the assembly of long metal chains. Initial reactions with

Mn(AcO)2 in DMF or pyridine facilitate the formation of the chain tetra-

nuclear aggregates [Mn4(H2L2)2(AcO)2(dmf)4] and [Mn4(H2L2)2

(AcO)2(py)5], respectively (Figure 13). These two related complexes

exhibit four Mn(II) ions sandwiched and chelated by two l4�(H2L2)3�

Figure 12. Representation of the complex cation [Ni2Gd(NO3)2(MeL4)2]þ. H atoms not

shown. dNi���Pr¼ 3.688 A. (Color figure available online.)

184 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 25: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

donors and exhibit two additional syn,syn�AcO� bridges, with coordi-

nation sites completed by terminal solvent ligands. The topology of the

metals in these compounds, however, is removed from straight linear;

instead, the metals describe a zigzag (Figure 13). It is very likely that

the reason for this is related to structural constraints imposed by the

multidentate ligand. Indeed, the backbone’s rigidity of H5L2 may prevent

restraining the metals within a straight line and instead the ligand rotates

around some of its �C�C�bonds, thus allowing the metals to externally

coordinate other ligands. This provides more flexibility and enables the

formation of M�O bonds with the bis�b�diketonate of the appropriate

length. If the reaction is performed in conditions favoring the oxidation to

Mn(III), another linear chain is formed which exhibits the unprecedented

sequence [MnIIIMnIIIMnII]; [Mn3(HL2)2(py)6] (Figure 14). In this

unique complex, the metals are within a quasi straight line (Mn�Mn�Mn angle of 172.92�), perhaps because they are more spaced and also

because MnIII favors shorter Mn�O equatorial bonds. Magnetic mea-

surements show that the coupling between the metals within the molecule

is antiferromagnetic, leading to a ground state of S¼ 5=2. Fitting of

Figure 13. Two views of the complex [Mn4(H2L2)2(AcO)2(dmf)4]. H atoms not shown.

dMn���Mn (ext)¼ 3.318 A, dMn���Mn (cent)¼ 3.388 A. (Color figure available online.)

METAL-BASED MOLECULAR CHAINS 185

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 26: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

the susceptibility data reveals that the coupling between Mn(III) ions is

stronger by one order of magnitude than the MnIII � � �MnII exchange.

Therefore, the unusual asymmetry of this linear disposition of metals does

not only relate to the sequence in oxidation states, but also in the magni-

tude of the magnetic exchange.

Another interesting ligand is H4L3, which is similar to H5L2 without

the central OH group. The structure of H4L3 (Scheme 10) suggests that

this donor may favor the assembly of linear complexes in the form of

two well-defined, separate groups. Indeed, in the presence of a strong

base (such as Bu4NOH), pyridine solutions of H4L3 and MX2 salts

(MII¼NiII, CoII, CuII; X¼ counter anions) produce tetranuclear chain

assemblies with formula [M4(L3)2(py)x] (x¼ 4 or 6).[81,82] These mole-

cules exhibit a linear disposition of four metals of the type MM � � �MM

(Figure 11), i.e., divided in two dimers. The structure of the assembly

causes both distinct metal sites to embody different stereochemical

properties. This has been exploited for the formation of heterometallic

metal chains with the same structure. Thus, the complex [Cu2Ni2(L3)2

(py)6] has been isolated and well characterized,[82] whereas the structure

of [Cu2Co2(L3)2(py)6] and [Co2Ni2(L3)2(py)6] has also been deter-

mined.[83] Within each dimer of the chain, the metals are coupled antifer-

romagnetically. If the cluster is homometallic, this leads to a diamagnetic

ground state; however, if both metals of the dimer are different, the mag-

netic exchange causes a quasi independent non-zero ground state on each

side of the cluster. This has allowed us to propose this kind of system as a

possible carrier of two spin-based qubits and thus constituting a proto-

type of 2qubit quantum gate for quantum computing.[84] Represented in

Figure 16 are DvMT vs T plots for a series of tetranuclear clusters of this

kind, reflecting these properties at the thermodynamic level.

Figure 14. Representation of the complex [Mn3(HL2)2(py)6]. H atoms not shown.

dMn(III)���Mn(III)¼ 5.246 A, dMn(II)���Mn(III)¼ 5.155 A. (Color figure available online.)

186 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 27: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

Ligand H2L5 displays one 1,3�pyridyl group separating two b�dike-

tones and two more lateral pyridyl fragments. The structure of this donor

suggests that it can accommodate rows of various metals, exhibiting dif-

ferent coordination geometry. This has been observed from reactions of

H2L5 with Co(NO3)2 in MeOH, which led to the tetranuclear chain

[Co4(L5)2(MeOH)8](NO3)4 (Figure 17).[85] In this complex, the metals

are accommodated within two unconventional coordination environ-

ments: pentagonal bipyramidal and very distorted octahedral

(Figure 17). This system was found to exhibit significant intramolecular

magnetic exchange. The [Co4(L5)2] backbone structure of this EMAC

constitutes a very flat platform with axial MeOH ligands perpendicular

to it, which proved to be very labile. This affords the opportunity to

replace these axial ligands by bridging species with the aim of linking

[Co4] units to each other.[86]

Figure 15. Representation of the complexes [Mn3(HL1)3] (dMn���Mn¼ 3.032 and 5.044 A)

and [Mn2(HL1)2(py)2] (dMn���Mn¼ 9.302 A), showing their solvent-driven interconversion.

H atoms not shown. (Color figure available online.)

Figure 16. 4vMT vs T plots for complexes [M2M02(L3)2(py)x] (x¼ 4 or 6; M, M0 ¼Ni, Co or

Cu). vM is the molar paramagnetic susceptibility.

METAL-BASED MOLECULAR CHAINS 187

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 28: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

The family of 1,3�diketones may be readily converted into a new

family of interesting ligands by converting the dicarbonyl groups into pyr-

azoles through cyclization with hydrazine (N2H4)[87] (Scheme 13). The

resulting species also exhibits a series of donor atoms in a row, potentially

capable of assembling metals chains. However, because of the bent struc-

ture of such ligands, these are not suited to obtain straight EMACS. In

fact, multinuclear complex structures instead of linear molecules are

often obtained.[88,89] Nevertheless, some unique metallic chains with a

bent structure have been prepared with these polypyrazolyl derivatives.

Examples are the following: i) The trinuclear complex [Mn3(HL10)2

(OAc)3(MeOH)3], made with H3L10 (the bis-pyrazol derived from

H3L1)[90]; ii) two mixed CoII=CoIII tetranuclear chains with formula

[Co4(OR)(OAc)(L30)2(py)4] (R¼H, CH3; H4L30 ¼ bis-pyrazolyl from

H4L3; Figure 18)[91]; iii) a unique tetranuclear manganese complex with

the unprecedented sequence of oxidation states (MnIIIMnIIIMnIIMnIII) and

formula [Mn4(L20)2(OAc)(MeOH)5] (H5L20 ¼ bispyrazolyl of H5L2).[92]

Figure 17. Representation of the complex cation [Co4(L5)2(MeOH)8]4þ, and the coordi-

nation environment of the CoII centers. H atoms not shown. dCo���Co (ext)¼ 3.550 A, dCo���Co

Co (cent)¼ 3.692 A. (Color figure available online.)

Scheme 13.

188 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 29: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

4. CONCLUSIONS AND OUTLOOK

The main conclusion extracted from this overview on molecular EMACS

is that the methods and strategies employed for the preparation of this

type of species have reached a high degree of diversity and versatility.

From this work, it follows that the limit to the length that can be attained

for discrete molecular chains will likely be imposed, either by the struc-

ture of the supporting ligands or by a fine balance between thermodyn-

amic stability and solubility (for the case of many ‘‘unsupported’’

chains). The progress made has not only allowed accessing linear clusters

of a large variety of metals with many types of chemical and electronic

properties; the various methods of molecular assembly should also open

ways for introducing specific properties to these architectures such as

precise conductivity, redox-behavior, magnetic properties, etc. Examples

of this are the capacity to induce spin cross-over properties to [Co3]

chains[93] or influence the conductance of [Ni5] and [Cr5] chains[94]

through chemical or electrochemical oxidation. The capacity of linear

chains to respond in desired manners to external stimuli or the ability

of integrating these into more complex molecular devices remain future

challenges for their potential applications in nanotechnology. One excit-

ing proposal awaiting such control on addressability and processing

suggests that spin chains could be used for the realization of one- or

two-qubit quantum gates.[95] For this and other potential applications,

the ability to fix linear molecules on surfaces of various types is of

paramount importance. This is being achieved to a large degree with

single-chain, 1D, molecule-based systems.[96] As great progress is being

made on these areas, a growing tendency for interdisciplinary activity

Figure 18. Representation of the complex [Co4(OMe)(OAc)(L30)2(py)4] (H4L30 ¼bis-

pyrazolyl from H4L3). H atoms not shown. dCo(III)���Co(II) (ext.)¼ 3.691 A, dCo(II)���Co(II)

(cent.)¼ 3.193 A. (Color figure available online.)

METAL-BASED MOLECULAR CHAINS 189

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 30: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

within molecular materials science will lead to major breakthroughs in the

use of molecular chains as part of nanoscopic functional assemblies.

REFERENCES

1. Bera, J. K.; Dunbar, K. R. Angew. Chem.-Int. Edit. 2002, 41, 4453–4457.

2. Mas-Balleste, R.; Castillo, O.; Sanz Miguel, P. J.; Olea, D.; Gomez-Herrero,

J.; Zamora, F. Eur. J. Inorg. Chem. 2009, 2885–2896.

3. Tsai, T. W.; Huang, Q. R.; Peng, S. M.; Jin, B. Y. J. Phys. Chem. C 2010, 114,

3641–3644.

4. Chen, I. W. P.; Fu, M. D.; Tseng, W. H.; Yu, J. Y.; Wu, S. H.; Ku, C. J.; Chen,

C. H.; Peng, S. M. Angew. Chem.-Int. Edit. 2006, 45, 5814–5818.

5. Liu, I. P. C.; Chen, C. F.; Hua, S. A.; Chen, C. H.; Wang, H. T.; Lee, G. H.;

Peng, S. M. Dalton Trans. 2009, 3571–3573.

6. Mann, K. R.; Dipierro, M. J.; Gill, T. P. J. Am. Chem. Soc. 1980, 102,

3965–3967.

7. Toth, A.; Floriani, C.; Chiesi-Villa, A.; Guastini, C. Inorg. Chem. 1987, 26,

236–241.

8. Sevryugina, Y.; Olenev, A. V.; Petrukhina, M. A. J. Cluster Sci. 2005, 16, 217–229.

9. Pruchnik, F. P.; Jutarska, A.; Ciunik, Z.; Pruchnik, M. Inorg. Chim. Acta

2003, 350, 609–616.

10. Sakai, K.; Ishigami, E. Acta Cryst. E 2004, 60, m65–m68.

11. Tejel, C.; Ciriano, M. A.; Villarroya, B. E.; Gelpi, R.; Lopez, J. A.; Lahoz, F.

J.; Oro, L. A. Angew. Chem. Int. Ed. 2001, 40, 4084–4086.

12. Tejel, C.; Ciriano, M. A.; Villarroya, B. E.; Lopez, J. A.; Lahoz, F. J.; Oro, L.

A. Angew. Chem. Int. Ed. 2003, 42, 529–532.

13. Barton, J. K.; Caravana, C.; Lippard, S. J. J. Am. Chem. Soc. 1979, 101,

7269–7277.

14. Matsumoto, K.; Sakai, K.; Nishio, K.; Tokisue, Y.; Ito, R.; Nishide, T.;

Shichi, Y. J. Am. Chem. Soc. 1992, 114, 8110–8118.

15. Chern, S. S.; Lee, G. H.; Peng, S. M. J. Chem. Soc., Chem. Commun. 1994,

1645–1646.

16. Laguna, A.; Laguna, M.; Jimenez, J.; Lahoz, F. J.; Olmos, E. Organometallics

1994, 13, 253–257.

17. Berry, J. F. Struct. Bonding 2010, 136, 1–28.

18. Lai, S.-Y.; Lin, T.-W.; Chen, Y.-H.; Wang, C.-C.; Lee, G.-H.; Yang, M.-H.;

Leung, M.-K.; Peng, S.-M. J. Am. Chem. Soc. 1998, 121, 250–251.

19. Chien, C. H.; Chang, J. C.; Yeh, C. Y.; Lee, G. H.; Fang, J. M.; Song, Y.;

Peng, S. M. Dalton Trans. 2006, 3249–3256.

20. Lai, S. Y.; Wang, C. C.; Chen, Y. H.; Lee, C. C.; Liu, Y. H.; Peng, S. M. J.

Chin. Chem. Soc. 1999, 46, 477–485.

190 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 31: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

21. Ismayilov, R. H.; Wang, W. Z.; Wang, R. R.; Yeh, C. Y.; Lee, G. H.; Peng, S.

M. Chem. Commun. 2007, 1121–1123.

22. Liu, I. P.-C.; Benard, M.; Hasanov, H.; Chen, I. W. P.; Tseng, W.-H.; Fu,

M.-D.; Rohmer, M.-M.; Chen, C.-H.; Lee, G.-H.; Peng, S.-M. Chem. Eur.

J. 2007, 13, 8667–8677.

23. Ismayilov, R. H.; Wang, W.-Z.; Wang, R.-R.; Huang, Y.-L.; Yeh, C.-Y.; Lee,

G.-H.; Peng, S.-M. Eur. J. Inorg. Chem. 2008, 4290–4295.

24. Ismayilov, R. H.; Wang, W. Z.; Lee, G. H.; Yeh, C. Y.; Hua, S. A.; Song, Y.;

Rohmer, M. M.; Benard, M.; Peng, S. M. Angew. Chem.-Int. Edit. 2011, 50,

2045–2048.

25. Murahashi, T.; Mochizuki, E.; Kai, Y.; Kurosawa, H. J. Am. Chem. Soc. 1999,

121, 10660–10661.

26. Labeguerie, P.; Benard, M.; Rohmer, M.-M. Inorg. Chem. 2007, 46,

5283–5291.

27. Murahashi, T.; Uemura, T.; Kurosawa, H. J. Am. Chem. Soc. 2003, 125,

8436–8437.

28. Tatsumi, Y.; Murahashi, T.; Okada, M.; Ogoshi, S.; Kurosawa, H. Chem.

Commun. 2008, 477–479.

29. Li, J.-M.; Yong, G.-P.; Yu, Z.-P.; Wang, Z.-Y. Acta Crystallographica Section E

2006, 62, m327–m328.

30. Reimann, C. W.; Zocchi, M. Acta Crystallographica Section B 1971, 27,

682–691.

31. Miras, H. N.; Zhao, H.; Herchel, R.; Rinaldi, C.; Perez, S.; Raptis, R. G. Eur.

J. Inorg. Chem. 2008, 4745–4755.

32. Rietmeijer, F. J.; Vanalbada, G. A.; Degraaff, R. A. G.; Haasnoot, J. G.;

Reedijk, J. Inorg. Chem. 1985, 24, 3597–3601.

33. Antolini, L.; Fabretti, A. C.; Gatteschi, D.; Giusti, A.; Sessoli, R. Inorg.

Chem. 1990, 29, 143–145.

34. Mochizuki, T.; Nogami, T.; Ishida, T. Inorg. Chem. 2009, 48, 2254–2259.

35. Drabent, K.; Ciunik, Z. Chem. Commun. 2001, 1254–1255.

36. Roubeau, O.; Castro, M.; Burriel, R.; Haasnoot, J. G.; Reedijk, J. J. Phys.

Chem. B 115, 3003–3012.

37. Beckmann, U.; Brooker, S. Coord. Chem. Rev. 2003, 245, 17–29.

38. Liu, B.; Xu, L.; Guo, G. C.; Huang, J. S. J. Mol. Struct. 2006, 825, 79–86.

39. Shakirova, O. G.; Virovets, A. V.; Naumov, D. Y.; Shvedenkov, Y. G.;

Elokhina, V. N.; Lavrenova, L. G. Inorg. Chem. Commun. 2002, 5, 690–693.

40. Garcia, Y.; Guionneau, P.; Bravic, G.; Chasseau, D.; Howard, J. A. K.; Kahn, O.;

Ksenofontov, V.; Reiman, S.; Gutlich, P. Eur. J. Inorg. Chem. 2000, 1531–1538.

41. Aromi, G.; Barrios, L. A.; Roubeau, O.; Gamez, P. Coord. Chem. Rev. 2011,

255, 485–546.

42. Ardizzoia, G. A.; LaMonica, G.; Maspero, A.; Moret, M.; Masciocchi, N.

Eur. J. Inorg. Chem. 2000, 181–187.

METAL-BASED MOLECULAR CHAINS 191

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 32: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

43. Pinillos, M. T.; Tejel, C.; Oro, L. A.; Apreda, M. C.; Foces-Foces, C.; Cano,

F. H. J. Chem. Soc., Dalton Trans. 1989, 1133–1138.

44. Pal, S.; Barik, A. K.; Gupta, S.; Hazra, A.; Kar, S. K.; Peng, S.-M.; Lee,

G.-H.; Butcher, R. J.; El Fallah, M. S.; Ribas, J. Inorg. Chem. 2005, 44,

3880–3889.

45. Zhao, Q. H.; Li, H. F.; Chen, Z. D.; Fang, R. B. Inorg. Chim. Acta 2002, 336,

142–146.

46. Cano, J.; De Munno, G.; Lloret, F.; Julve, M. Inorg. Chem. 2000, 39,

1611–1614.

47. Otieno, T.; Rettig, S. J.; Thompson, R. C.; Trotter, J. Inorg. Chem. 1995, 34,

1718–1725.

48. Yi, T.; Ho-Chol, C.; Gao, S.; Kitagawa, S Eur. J. Inorg. Chem. 2006,

1381–1387.

49. Matthews, C. J.; Onions, S. T.; Morata, G.; Davis, L. J.; Heath, S. L.; Price,

D. J. Angew. Chem. Int. Ed. 2003, 42, 3166–3169.

50. Dey, S. K.; Abedin, T. S. M.; Dawe, L. N.; Tandon, S. S.; Collins, J. L.;

Thompson, L. K.; Postnikov, A. V.; Alam, M. S.; Muller, P. Inorg. Chem.

2007, 46, 7767–7781.

51. Shuvaev, K. V.; Tandon, S. S.; Dawe, L. N.; Thompson, L. K. Chem. Commun.

2010, 46, 4755–4757.

52. Piguet, C.; Bernardinelli, G.; Hopfgartner, G. Chem. Rev. 1997, 97,

2005–2062.

53. Wu, W.-S.; Cheng, W.-D.; Wu, D.-S.; Zhang, H.; Gong, Y.-J.; Lu, Y. Inorg.

Chem. Commun. 2006, 9, 559–562.

54. Li, D.; Wang, S.; Xu, H.; Yang, Y.; Zeng, S.; Zhao, J.; Wang, D.; Dou, J.

Inorg. Chim. Acta 2011, 365, 85–95.

55. Yu, Z.-X.; Sun, Y.-X. Acta Crystallogr., E 2006, 62, m3521–m3523.

56. Hoskins, B. F.; McDonald, I. A. S. Aust. J. Chem. 1989, 42, 1057–1065.

57. Shi, X.; Li, D.; Wang, S.; Zeng, S.; Wang, D.; Dou, J. J. Solid State Chem.

2010, 183, 2144–2153.

58. Pardo, E.; Ferrando-Soria, J.; Dul, M. C.; Lescouezec, R.; Journaux, Y.;

Ruiz-Garcia, R.; Cano, J.; Julve, M.; Lloret, F.; Canadillas-Delgado, L.;

Pasan, J.; Ruiz-Perez, C. Chem. Eur. J. 2010, 16, 12838–12851.

59. Paital, A. R.; Mitra, T.; Ray, D.; Wong, W. T.; Ribas-Arino, J.; Novoa, J. J.;

Ribas, J.; Aromi, G. Chem. Commun. 2005, 5172–5174.

60. Kourkine, I. V.; Sargent, M. D.; Glueck, D. S.Organometallics 1998, 17, 125–127.

61. Tanase, T.; Begum, R. A.; Toda, H.; Yamamoto, Y. Organometallics 2001, 20,

968–979.

62. Goto, E.; Begum, R. A.; Zhan, S.; Tanase, T.; Tanigaki, K.; Sakai, K. Angew.

Chem. Int. Ed. 2004, 43, 5029–5032.

63. Balch, A. L.; Catalano, V. J.; Chatfield, M. A.; Nagle, J. K.; Olmstead, M.

M.; Reedy, P. E. J. Am. Chem. Soc. 1991, 113, 1252–1258.

192 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 33: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

64. Balch, A. L.; Olmstead, M. M.; Neve, F.; Ghedini, M. New. J. Chem. 1988, 12,

529–537.

65. Balch, A. L.; Fossett, L. A.; Olmstead, M. M.; Oram, D. E.; Reedy, P. E.

J. Am. Chem. Soc. 1985, 107, 5272–5274.

66. Mashima, K.; Nakano, H.; Nakamura, A. J. Am. Chem. Soc. 1996, 118,

9083–9095.

67. Ruffer, T.; Ohashi, M.; Shima, A.; Mizomoto, H.; Kaneda, Y.; Mashima, K.

J. Am. Chem. Soc. 2004, 126, 12244–12245.

68. Grillo, V. A.; Seddon, E. J.; Grant, C. M.; Aromi, G.; Bollinger, J. C.; Folting,

K.; Christou, G. Chem. Commun. 1997, 1561–1562.

69. Saalfrank, R. W.; Seitz, V.; Caulder, D. L.; Raymond, K. N.; Teichert, M.;

Stalke, D. Eur. J. Inorg. Chem. 1998, 1313–1317.

70. Aromi, G.; Gamez, P.; Roubeau, O.; Carrero-Berzal, P.; Kooijman, H.; Spek,

A. L.; Driessen, W. L.; Reedijk, J. Eur. J. Inorg. Chem. 2002, 1046–1048.

71. Clegg, J. K.; Lindoy, L. F.; McMurtrie, J. C.; Schilter, D. Dalton Trans. 2005,

857–864.

72. Aromi, G.; Boldron, C.; Gamez, P.; Roubeau, O.; Kooijman, H.; Spek, A. L.;

Stoeckli-Evans, H.; Ribas, J.; Reedijk, J. Dalton Trans. 2004, 3586–3592.

73. Aromi, G.; Gamez, P.; Berzal, P. C.; Driessen, W. L.; Reedijk, J. Synth.

Commun. 2003, 33, 11–18.

74. Aromi, G.; Berzal, P. C.; Gamez, P.; Roubeau, O.; Kooijman, H.; Spek, A. L.;

Driessen, W. L.; Reedijk, J. Angew. Chem., Int. Edit. 2001, 40, 3444–3446.

75. Aromi, G.; Stoeckli-Evans, H.; Teat, S. J.; Cano, J.; Ribas, J. J. Mater. Chem.

2006, 16, 2635–2644.

76. Aromi, G.; Gamez, P.; Roubeau, O.; Berzal, P. C.; Kooijman, H.; Spek, A. L.;

Driessen, W. L.; Reedijk, J. Inorg. Chem. 2002, 41, 3673–3683.

77. Shiga, T.; Ito, N.; Hidaka, A.; AŒkawa, H.; Kitagawa, S.; Ohba, M. Inorg.

Chem. 2007, 46, 3492–3501.

78. Shiga, T.; Ohba, M.; Okawa, H. Inorg. Chem. 2004, 43, 4435–4446.

79. Shiga, T.; AŒkawa, H.; Kitagawa, S.; Ohba, M. J. Am. Chem. Soc. 2006, 128,

16426–16427.

80. Saalfrank, R. W.; Seitz, V.; Caulder, D. L.; Raymond, K. N.; Teichert, M.;

Stalke, D. Eur. J. Inorg. Chem. 1998, 1313–1317.

81. Barrios, L. A.; Aguila, D.; Mellat, S.; Roubeau, O.; Teat, S. J.; Gamez, P.;

Aromi, G. C. R. Chim. 2008, 11, 1117–1120.

82. Barrios, L. A.; Aguila, D.; Roubeau, O.; Gamez, P.; Ribas-Arino, J.; Teat, S.

J.; Aromi, G. Chem.-Eur. J. 2009, 15, 11235–11243.

83. Aromi, G.; Roubeau, O.; Barrios, L. A.; Aguila, D. Unpublished results.

84. Aromi, G.; Aguila, D.; Gamez, P.; Roubeau, O.; Luis, F. Chem. Soc. Rev.

2012, 41, 537–546.

85. Barrios, L. A.; Aguila, D.; Roubeau, O.; Murray, K. S.; Aromi, G. Aust. J.

Chem. 2009, 62, 1130–1136.

METAL-BASED MOLECULAR CHAINS 193

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3

Page 34: METAL-BASED MOLECULAR CHAINS: DESIGN BY COORDINATION CHEMISTRY

86. Aguila, D.; Barrios, L. A.; Roubeau, O.; Teat, S. J.; Aromi, G. Chem.

Commun. 2011, 47, 707–709.

87. Gosselin, F.; O’Shea, P. D.; Webster, R. A.; Reamer, R. A.; Tillyer, R. D.;

Grabowski, E. J. J. Synlett 2006, 3267–3270.

88. Newton, G. N.; Onuki, T.; Shiga, T.; Noguchi, M.; Matsumoto, T.;

Mathieson, J. S.; Nihei, M.; Nakano, M.; Cronin, L.; Oshio, H. Angew.

Chem., Int. Ed. 2011, 50, 4844–4848.

89. Costa, J. S.; Craig, G. A.; Barrios, L. A.; Roubeau, O.; Ruiz, E.;

Gomez-Coca, S.; Teat, S. J.; Aromi, G. Chem.-Eur. J. 2011, 17, 4960–4963.

90. Barrios, L. A.; Aromi, G.; Ribas, J.; Uber, J. S.; Roubeau, O.; Sakai, K.;

Masaoka, S.; Gamez, P.; Reedijk, J. Eur. J. Inorg. Chem. 2008, 3871–3876.

91. Craig, G. A.; Costa, J. S.; Aguila, D.; Barrios, L. A.; Roubeau, O.; Teat, S. J.;

Aromi, G. New J. Chem. 2011, 35, 1202–1204.

92. Aromi, G.; Roubeau, O.; Craig, G. A. Unpublished results.

93. Clerac, R.; Cotton, F. A.; Dunbar, K. R.; Lu, T. B.; Murillo, C. A.; Wang, X.

P. J. Am. Chem. Soc. 2000, 122, 2272–2278.

94. Lin, S. Y.; Chen, I. W. P.; Chen, C. H.; Hsieh, M. H.; Yeh, C. Y.; Lin, T. W.;

Chen, Y. H.; Peng, S. M. J. Phys. Chem. B 2004, 108, 959–964.

95. Meier, F.; Levy, J.; Loss, D. Phys. Rev. Lett. 2003, 90, 047901.

96. Mas-Balleste, R.; Gomez-Herrero, J.; Zamora, F. Chem. Soc. Rev. 2010, 39,

4220–4233.

194 G. AROMI

Dow

nloa

ded

by [

Uni

vers

ity o

f W

yom

ing

Lib

rari

es]

at 1

0:22

05

Oct

ober

201

3


Recommended