11.01.2013 Views

Photochemistry and Photophysics of Coordination Compounds

Photochemistry and Photophysics of Coordination Compounds

Photochemistry and Photophysics of Coordination Compounds

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

280<br />

Topics in Current Chemistry<br />

Editorial Board:<br />

V. Balzani · A. de Meijere · K. N. Houk · H. Kessler · J.-M. Lehn<br />

S. V. Ley · S. L. Schreiber · J. Thiem · B. M. Trost · F. Vögtle<br />

H. Yamamoto


Topics in Current Chemistry<br />

Recently Published <strong>and</strong> Forthcoming Volumes<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong><br />

<strong>Coordination</strong> <strong>Compounds</strong> II<br />

Volume Editors: Balzani, C., Campagna, S.<br />

Vol. 281, 2007<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong><br />

<strong>Coordination</strong> <strong>Compounds</strong> I<br />

Volume Editors: Balzani, C., Campagna, S.<br />

Vol. 280, 2007<br />

Metal Catalyzed Reductive C–C Bond Formation<br />

A Departure from Preformed Organometallic<br />

Reagents<br />

VolumeEditor:Krische,M.J.<br />

Vol. 279, 2007<br />

Combinatorial Chemistry on Solid Supports<br />

Volume Editor: Bräse, S.<br />

Vol. 278, 2007<br />

Creative Chemical Sensor Systems<br />

Volume Editor: Schrader, T.<br />

Vol. 277, 2007<br />

In situ NMR Methods in Catalysis<br />

Volume Editors: Bargon, J., Kuhn, L. T.<br />

Vol. 276, 2007<br />

Sulfur-Mediated Rearrangements II<br />

Volume Editor: Schaumann, E.<br />

Vol. 275, 2007<br />

Sulfur-Mediated Rearrangements I<br />

Volume Editor: Schaumann, E.<br />

Vol. 274, 2007<br />

Bioactive Conformation II<br />

Volume Editor: Peters, T.<br />

Vol. 273, 2007<br />

Bioactive Conformation I<br />

Volume Editor: Peters, T.<br />

Vol. 272, 2007<br />

Biomineralization II<br />

Mineralization Using Synthetic Polymers <strong>and</strong><br />

Templates<br />

Volume Editor: Naka, K.<br />

Vol. 271, 2007<br />

Biomineralization I<br />

Crystallization <strong>and</strong> Self-Organization Process<br />

Volume Editor: Naka, K.<br />

Vol. 270, 2007<br />

Novel Optical Resolution Technologies<br />

Volume Editors:<br />

Sakai, K., Hirayama, N., Tamura, R.<br />

Vol. 269, 2007<br />

Atomistic Approaches in Modern Biology<br />

From Quantum Chemistry<br />

to Molecular Simulations<br />

Volume Editor: Reiher, M.<br />

Vol. 268, 2006<br />

Glycopeptides <strong>and</strong> Glycoproteins<br />

Synthesis, Structure, <strong>and</strong> Application<br />

Volume Editor: Wittmann, V.<br />

Vol. 267, 2006<br />

Microwave Methods in Organic Synthesis<br />

Volume Editors: Larhed, M., Ol<strong>of</strong>sson, K.<br />

Vol. 266, 2006<br />

Supramolecular Chirality<br />

Volume Editors: Crego-Calama, M.,<br />

Reinhoudt, D. N.<br />

Vol. 265, 2006


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong><br />

<strong>Coordination</strong> <strong>Compounds</strong> I<br />

Volume Editors: Vincenzo Balzani · Sebastiano Campagna<br />

With contributions by<br />

G. Accorsi · N. Armaroli · V. Balzani · G. Bergamini · S. Campagna<br />

F. Cardinali · C. Chiorboli · M. T. Indelli · N. A. P. Kane-Maguire<br />

A. Listorti · F. Nastasi · F. Puntoriero · F. Sc<strong>and</strong>ola<br />

123


The series Topics in Current Chemistry presents critical reviews <strong>of</strong> the present <strong>and</strong> future trends in<br />

modern chemical research. The scope <strong>of</strong> coverage includes all areas <strong>of</strong> chemical science including<br />

the interfaces with related disciplines such as biology, medicine <strong>and</strong> materials science. The goal <strong>of</strong><br />

each thematic volume is to give the nonspecialist reader, whether at the university or in industry,<br />

a comprehensive overview <strong>of</strong> an area where new insights are emerging that are <strong>of</strong> interest to a larger<br />

scientific audience.<br />

As a rule, contributions are specially commissioned. The editors <strong>and</strong> publishers will, however, always<br />

be pleased to receive suggestions <strong>and</strong> supplementary information. Papers are accepted for Topics in<br />

Current Chemistry in English.<br />

In references Topics in Current Chemistry is abbreviated Top Curr Chem <strong>and</strong> is cited as a journal.<br />

Visit the TCC content at springerlink.com<br />

ISSN 0340-1022<br />

ISBN 978-3-540-73346-1 Springer Berlin Heidelberg New York<br />

DOI 10.1007/978-3-540-73347-8<br />

This work is subject to copyright. All rights are reserved, whether the whole or part <strong>of</strong> the material<br />

is concerned, specifically the rights <strong>of</strong> translation, reprinting, reuse <strong>of</strong> illustrations, recitation, broadcasting,<br />

reproduction on micr<strong>of</strong>ilm or in any other way, <strong>and</strong> storage in data banks. Duplication <strong>of</strong><br />

this publication or parts there<strong>of</strong> is permitted only under the provisions <strong>of</strong> the German Copyright Law<br />

<strong>of</strong> September 9, 1965, in its current version, <strong>and</strong> permission for use must always be obtained from<br />

Springer. Violations are liable for prosecution under the German Copyright Law.<br />

Springer is a part <strong>of</strong> Springer Science+Business Media<br />

springer.com<br />

c○ Springer-Verlag Berlin Heidelberg 2007<br />

The use <strong>of</strong> registered names, trademarks, etc. in this publication does not imply, even in the absence<br />

<strong>of</strong> a specific statement, that such names are exempt from the relevant protective laws <strong>and</strong> regulations<br />

<strong>and</strong> therefore free for general use.<br />

Cover design: WMXDesign GmbH, Heidelberg<br />

Typesetting <strong>and</strong> Production: LE-TEXJelonek,Schmidt&VöcklerGbR,Leipzig<br />

Printed on acid-free paper 02/3180 YL – 5 4 3 2 1 0


Volume Editors<br />

Pr<strong>of</strong>. Vincenzo Balzani<br />

Dipartimento di Chimica “G. Ciamician”<br />

Università di Bologna<br />

Via Selmi 2<br />

40126 Bologna<br />

Italy<br />

vincenzo.balzani@unibo.it<br />

Editorial Board<br />

Pr<strong>of</strong>. Vincenzo Balzani<br />

Dipartimento di Chimica „G. Ciamician“<br />

University <strong>of</strong> Bologna<br />

via Selmi 2<br />

40126 Bologna, Italy<br />

vincenzo.balzani@unibo.it<br />

Pr<strong>of</strong>. Dr. Armin de Meijere<br />

Institut für Organische Chemie<br />

der Georg-August-Universität<br />

Tammanstr. 2<br />

37077 Göttingen, Germany<br />

ameijer1@uni-goettingen.de<br />

Pr<strong>of</strong>. Dr. Kendall N. Houk<br />

University <strong>of</strong> California<br />

Department <strong>of</strong> Chemistry <strong>and</strong><br />

Biochemistry<br />

405 Hilgard Avenue<br />

Los Angeles, CA 90024-1589<br />

USA<br />

houk@chem.ucla.edu<br />

Pr<strong>of</strong>. Dr. Horst Kessler<br />

Institut für Organische Chemie<br />

TU München<br />

Lichtenbergstraße 4<br />

86747 Garching, Germany<br />

kessler@ch.tum.de<br />

Pr<strong>of</strong>. Sebastiano Campagna<br />

Dipartimento di Chimica Inorganica<br />

Chimica Analitica e Chimica Fisica<br />

University <strong>of</strong> Messina<br />

Via Sperone 31<br />

98166 Vill. S.Agata, Messina<br />

Italy<br />

campagna@unime.it<br />

Pr<strong>of</strong>. Jean-Marie Lehn<br />

ISIS<br />

8, allée Gaspard Monge<br />

BP 70028<br />

67083 Strasbourg Cedex, France<br />

lehn@isis.u-strasbg.fr<br />

Pr<strong>of</strong>. Steven V. Ley<br />

University Chemical Laboratory<br />

Lensfield Road<br />

Cambridge CB2 1EW<br />

Great Britain<br />

Svl1000@cus.cam.ac.uk<br />

Pr<strong>of</strong>. Stuart L. Schreiber<br />

Chemical Laboratories<br />

Harvard University<br />

12 Oxford Street<br />

Cambridge, MA 02138-2902<br />

USA<br />

sls@slsiris.harvard.edu<br />

Pr<strong>of</strong>. Dr. Joachim Thiem<br />

Institut für Organische Chemie<br />

Universität Hamburg<br />

Martin-Luther-King-Platz 6<br />

20146 Hamburg, Germany<br />

thiem@chemie.uni-hamburg.de


VI Editorial Board<br />

Pr<strong>of</strong>. Barry M. Trost<br />

Department <strong>of</strong> Chemistry<br />

Stanford University<br />

Stanford, CA 94305-5080<br />

USA<br />

bmtrost@lel<strong>and</strong>.stanford.edu<br />

Pr<strong>of</strong>. Dr. F. Vögtle<br />

Kekulé-Institut für Organische Chemie<br />

und Biochemie<br />

der Universität Bonn<br />

Gerhard-Domagk-Str. 1<br />

53121 Bonn, Germany<br />

voegtle@uni-bonn.de<br />

Pr<strong>of</strong>. Dr. Hisashi Yamamoto<br />

Department <strong>of</strong> Chemistry<br />

The University <strong>of</strong> Chicago<br />

5735 South Ellis Avenue<br />

Chicago, IL 60637<br />

USA<br />

yamamoto@uchicago.edu


Topics in Current Chemistry<br />

Also Available Electronically<br />

For all customers who have a st<strong>and</strong>ing order to Topics in Current Chemistry,<br />

we <strong>of</strong>fer the electronic version via SpringerLink free <strong>of</strong> charge. Please contact<br />

your librarian who can receive a password or free access to the full articles by<br />

registering at:<br />

springerlink.com<br />

If you do not have a subscription, you can still view the tables <strong>of</strong> contents <strong>of</strong> the<br />

volumes <strong>and</strong> the abstract <strong>of</strong> each article by going to the SpringerLink Homepage,<br />

clicking on “Browse by Online Libraries”, then “Chemical Sciences”, <strong>and</strong><br />

finally choose Topics in Current Chemistry.<br />

You will find information about the<br />

– Editorial Board<br />

–Aims<strong>and</strong>Scope<br />

– Instructions for Authors<br />

–SampleContribution<br />

at springer.com using the search function.


Preface<br />

<strong>Photochemistry</strong> (a term that broadly speaking includes photophysics) is<br />

a branch <strong>of</strong> modern science that deals with the interaction <strong>of</strong> light with matter<br />

<strong>and</strong> lies at the crossroads <strong>of</strong> chemistry, physics, <strong>and</strong> biology. However, before<br />

being a branch <strong>of</strong> modern science, photochemistry was (<strong>and</strong> still is today),<br />

an extremely important natural phenomenon. When God said: “Let there be<br />

light”, photochemistry began to operate, helping God to create the world as<br />

we now know it. It is likely that photochemistry was the spark for the origin <strong>of</strong><br />

life on Earth <strong>and</strong> played a fundamental role in the evolution <strong>of</strong> life. Through<br />

the photosynthetic process that takes place in green plants, photochemistry<br />

is responsible for the maintenance <strong>of</strong> all living organisms. In the geological<br />

past photochemistry caused the accumulation <strong>of</strong> the deposits <strong>of</strong> coal, oil, <strong>and</strong><br />

natural gas that we now use as fuels. <strong>Photochemistry</strong> is involved in the control<br />

<strong>of</strong> ozone in the stratosphere <strong>and</strong> in a great number <strong>of</strong> environmental processes<br />

that occur in the atmosphere, in the sea, <strong>and</strong> on the soil. <strong>Photochemistry</strong> is the<br />

essence <strong>of</strong> the process <strong>of</strong> vision <strong>and</strong> causes a variety <strong>of</strong> behavioral responses in<br />

living organisms.<br />

<strong>Photochemistry</strong> as a science is quite young; we only need to go back less<br />

than one century to find its early pioneer [1]. The concept <strong>of</strong> coordination<br />

compounds is also relatively young; it was established in 1892, when Alfred<br />

Werner conceived his theory <strong>of</strong> metal complexes [2]. Since then, the terms<br />

coordination compound <strong>and</strong> metal complex have been used as synonyms,<br />

even if in the last 30 years, coordination chemistry has extended its scope to<br />

the binding <strong>of</strong> all kinds <strong>of</strong> substrates [3, 4].<br />

The photosensitivity <strong>of</strong> metal complexes has been recognized for a long<br />

time, but the photochemistry <strong>and</strong> photophysics <strong>of</strong> coordination compounds<br />

as a science only emerged in the second half <strong>of</strong> the last century. The first attempt<br />

to systematize the photochemical reactions <strong>of</strong> coordination compounds<br />

was carried out in an exhaustive monograph published in 1970 [5], followed by<br />

an authoritative multi-authored volume in 1975 [6]. These two books gained<br />

the attention <strong>of</strong> the scientific community <strong>and</strong> certainly helped several inorganic<br />

<strong>and</strong> physical chemists to enter the field <strong>and</strong> to enrich <strong>and</strong> diversify their<br />

research activities. Interestingly, 1974 marked the beginning <strong>of</strong> the series <strong>of</strong><br />

International Symposia on the <strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> Coordi-


X Preface<br />

nation <strong>Compounds</strong>. The venue <strong>of</strong> the 17th symposium <strong>of</strong> this series is Dublin<br />

<strong>and</strong> will be held in June 2007.<br />

Up until about 1975, most activity was focused on intramolecular photochemical<br />

reactions. Subsequently, due partly also to the more diffuse availability<br />

<strong>of</strong> flash techniques, the interest <strong>of</strong> several groups moved to investigations<br />

<strong>of</strong> luminescence <strong>and</strong> bimolecular energy <strong>and</strong> electron transfer processes. In<br />

the last decade <strong>of</strong> the past century, with the development <strong>of</strong> supramolecular<br />

chemistry, it was clear that photochemistry would play a very important role in<br />

the achievement <strong>of</strong> valuable functions, such as charge separation, energy migration<br />

<strong>and</strong> conformational changes [7], related to applications spanning from<br />

solar energy conversion to signal processing <strong>and</strong> molecular machines [8, 9]. In<br />

the last few years, an increasing number <strong>of</strong> scientists have become involved in<br />

these fields. Because <strong>of</strong> their unique ground <strong>and</strong> excited state properties, metal<br />

complexeshavebecomeinvaluablecomponents<strong>of</strong>moleculardevices<strong>and</strong>machines<br />

exploiting light (<strong>of</strong>ten sunlight) to perform useful functions [8, 9, 10].<br />

The photochemistry <strong>of</strong> coordination compounds can also contribute to solving<br />

the energy crisis by converting sunlight into electricity or fuel [11]. In the<br />

meantime, the basic knowledge <strong>of</strong> the excited state properties <strong>of</strong> coordination<br />

compounds <strong>of</strong> several metal ions has increased considerably. However, this<br />

has resulted in an unavoidable loss <strong>of</strong> general knowledge <strong>and</strong> an increase in<br />

specialization. Currently, all scientists working in the field <strong>of</strong> the photochemistry<br />

<strong>and</strong> photophysics <strong>of</strong> coordination compounds have their own preferred<br />

metal. There is, therefore, an urgent need to spread the most recent developments<br />

in the field among the photochemical community. To write an exhaustive<br />

monograph like [5], however, would now be an impossible enterprise. For this<br />

reason, we decided to ask experts to write separate chapters, each one dealing<br />

with a specific metal whose complexes are currently at the frontier <strong>of</strong> research.<br />

It has been a delight as well as a privilege to work with an outst<strong>and</strong>ing group<br />

<strong>of</strong> contributing authors <strong>and</strong> we thank them for all their efforts. We would also<br />

like to thank all the members <strong>of</strong> our research groups for their support.<br />

Bologna <strong>and</strong> Messina, March 2007 Vincenzo Balzani<br />

Sebastiano Campagna<br />

References<br />

1. Ciamician G (1912) Science 36:385<br />

2. Werner A (1893) Zeit Anorg Chem 3:267<br />

3. LehnJ-M(1992)Fromcoordinationchemistrytosupramolecular chemistry.In:Williams<br />

AF, Floriani C, Merbach AE (eds) Perspectives in coordination chemistry. VCH, Weinheim,p.447<br />

4. Balzani V, Credi A, Venturi M (1998) Coord Chem Rev 171:3<br />

5. Balzani V, Carassiti V (1970) <strong>Photochemistry</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>. Academic<br />

Press, London


Preface XI<br />

6. Adamson AW, Fleischauer PD (eds) (1975) Concepts in inorganic photochemistry.<br />

Wiley-Interscience, New York<br />

7. Balzani V, Sc<strong>and</strong>ola F (1991) Supramolecular photochemistry. Horwood, Chichester<br />

8. Balzani V, Credi A, Venturi M (2003) Molecular devices <strong>and</strong> machines. Wiley-VCH,<br />

Weinheim<br />

9. Kelly TR (ed) (2005) Molecular machines, Topics Current Chem 262<br />

10. Kay EU, Leigh DA, Zerbetto F (2007) Angew Chem Int Ed 46:72<br />

11. Armaroli N, Balzani V (2007) Angew Chem Int Ed 46:52


Contents<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Overview <strong>and</strong> General Concepts<br />

V.Balzani·G.Bergamini·S.Campagna·F.Puntoriero ......... 1<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Chromium<br />

N.A.P.Kane-Maguire ........................... 37<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Copper<br />

N.Armaroli·G.Accorsi·F.Cardinali·A.Listorti ............ 69<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Ruthenium<br />

S. Campagna · F. Puntoriero · F. Nastasi · G. Bergamini · V. Balzani . . . 117<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Rhodium<br />

M.T.Indelli·C.Chiorboli·F.Sc<strong>and</strong>ola.................. 215<br />

Author Index Volumes 251–280 ...................... 257<br />

Subject Index ................................ 271


Contents <strong>of</strong> Volume 281<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong><br />

<strong>Compounds</strong> II<br />

Volume Editors: Balzani, S., Campagna, V.<br />

ISBN: 978-3-540-73348-5<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Lanthanides<br />

J. P. Leonard · C. B. Nolan · F. Stomeo · T. Gunnlaugsson<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Rhenium<br />

R. A. Kirgan · B. P. Sullivan · D. P. Rillema<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Osmium<br />

D. Kumaresan · K. Shankar · S. Vaidya · R. H. Schmehl<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Iridium<br />

L.Flamigni·A.Barbieri·C.Sabatini·B.Ventura·F.Barigelletti<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Platinum<br />

J. A. G. Williams<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Gold<br />

V.W.-W.Yam·E.C.-C.Cheng


Top Curr Chem (2007) 280: 1–36<br />

DOI 10.1007/128_2007_132<br />

© Springer-Verlag Berlin Heidelberg<br />

Published online: 23 June 2007<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong><br />

<strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Overview <strong>and</strong> General Concepts<br />

Vincenzo Balzani 1 (✉)·GiacomoBergamini 1 · Sebastiano Campagna 2 ·<br />

Fausto Puntoriero 1<br />

1 Dipartimento di Chimica “G. Ciamician”, Università di Bologna, 40100 Bologna, Italy<br />

vincenzo.balzani@unibo.it<br />

2 Dipartimento di Chimica Inorganica, Chimica Analitica, e Chimica Fisica,<br />

Università di Messina, 98166 Messina, Italy<br />

1 Early History<br />

And God said: “Let there be light”;<br />

And there was light.<br />

And God saw that the light was good.<br />

(Genesis, 1, 3–4)<br />

.................................. 2<br />

2 Molecular <strong>Photochemistry</strong> ........................... 3<br />

2.1OrganicMolecules................................ 3<br />

2.2MetalComplexes ................................ 5<br />

2.3LightAbsorption<strong>and</strong>IntramolecularExcited-StateDecay.......... 8<br />

3 Bimolecular Processes ............................. 10<br />

3.1GeneralFeatures................................. 10<br />

3.2BimolecularProcessesInvolvingMetalComplexes.............. 11<br />

4 Supramolecular <strong>Photochemistry</strong> ........................ 12<br />

4.1OperationalDefinition<strong>of</strong>SupramolecularSpecies .............. 12<br />

4.2PhotoinducedProcessesinSupramolecularSystems ............. 15<br />

4.3ElectronTransfer ................................ 16<br />

4.3.1MarcusTheory.................................. 16<br />

4.3.2QuantumMechanicalTheory.......................... 19<br />

4.3.3OpticalElectronTransfer............................ 20<br />

4.4EnergyTransfer ................................. 21<br />

4.4.1CoulombicMechanism ............................. 22<br />

4.4.2ExchangeMechanism.............................. 23<br />

5 <strong>Coordination</strong> <strong>Compounds</strong> as Components<br />

<strong>of</strong> Photochemical Molecular Devices <strong>and</strong> Machines ............. 24<br />

5.1AMolecularWire ................................ 24<br />

5.2AnAntennaSystem............................... 26<br />

5.3AnExtensionCable............................... 28<br />

5.4AnXORLogicGatewithanIntrinsicThresholdMechanism ........ 29<br />

5.5ASunlight-PoweredNanomotor ........................ 30<br />

6 Conclusions ................................... 33<br />

References ....................................... 33


2 V. Balzani et al.<br />

Abstract Investigations in the field <strong>of</strong> the photochemistry <strong>and</strong> photophysics <strong>of</strong> coordination<br />

compounds have proceeded along several steps <strong>of</strong> increasing complexity in the last<br />

50 years. Early studies on lig<strong>and</strong> photosubstitution <strong>and</strong> photoredox decomposition reactions<br />

<strong>of</strong> metal complexes <strong>of</strong> simple inorganic lig<strong>and</strong>s (e.g., NH3, CN – )werefollowed<br />

by accurate investigations on the photophysical behavior (luminescence quantum yields<br />

<strong>and</strong> lifetimes) <strong>and</strong> use <strong>of</strong> metal complexes in bimolecular processes (energy <strong>and</strong> electron<br />

transfer). The most significant differences between Jablonski diagrams for organic<br />

molecules <strong>and</strong> coordination compounds are illustrated. A large number <strong>of</strong> complexes<br />

stable toward photodecomposition, but capable <strong>of</strong> undergoing excited-state redox processes,<br />

have been used for interconverting light <strong>and</strong> chemical energy. The rate constants<br />

<strong>of</strong> a great number <strong>of</strong> photoinduced energy- <strong>and</strong> electron-transfer processes involving coordination<br />

compounds have been measured in order to prove the validity <strong>and</strong>/or extend<br />

the scope <strong>of</strong> modern kinetic theories. More recently, the combination <strong>of</strong> supramolecular<br />

chemistry <strong>and</strong> photochemistry has led to the design <strong>and</strong> construction <strong>of</strong> supramolecular<br />

systems capable <strong>of</strong> performing light- induced functions. In this field, luminescent<br />

<strong>and</strong>/or photoredox reactive metal complexes are presently used as essential components<br />

for a bottom-up approach to the construction <strong>of</strong> molecular devices <strong>and</strong> machines. A few<br />

examples <strong>of</strong> molecular devices for processing light signals <strong>and</strong> <strong>of</strong> molecular machines<br />

powered by light energy, based on coordination compounds, are briefly illustrated.<br />

Keywords <strong>Coordination</strong> compounds · Electron transfer · Energy transfer ·<br />

Excited-state properties · <strong>Photochemistry</strong> · Supramolecular photochemistry<br />

1<br />

Early History<br />

The photosensitivity <strong>of</strong> metal complexes has been known for a long time.<br />

The first paper exhibiting some scientific character was that <strong>of</strong> Scheele (1772)<br />

on the effect <strong>of</strong> light on AgCl, <strong>and</strong> photography was becoming established in<br />

several countries in the 1830s [1]. The light sensitivity <strong>of</strong> other metal complexes<br />

(particularly Na4[Fe(CN)6]) was also observed very early [2]. At the<br />

beginning <strong>of</strong> the last century the importance <strong>of</strong> photochemistry became more<br />

widely recognized, mainly due to the work <strong>and</strong> the ideas <strong>of</strong> Giacomo Ciamician<br />

[3], Pr<strong>of</strong>essor <strong>of</strong> Chemistry at the University <strong>of</strong> Bologna. In the same<br />

period (1912–1913), modern physics introduced the concept that light absorption<br />

corresponds to the capture <strong>of</strong> a photon by a molecule. This concept,<br />

<strong>and</strong> the distinction (sometimes difficult) between primary <strong>and</strong> secondary<br />

photoprocesses, led to the definition <strong>of</strong> quantum yield. In the following years,<br />

investigations on Fe 3+ <strong>and</strong> UO2 2+ complexes were performed in looking for<br />

useful chemical actinometers (see, e.g., [4]). Several quantitative works also<br />

appeared on the photochemical behavior <strong>of</strong> [Fe(CN)6] 4– <strong>and</strong> Co(III)–amine<br />

complexes in aqueous solution [2]. The lack <strong>of</strong> a theory on the absorption<br />

spectra <strong>and</strong> on the nature <strong>of</strong> the excited states, however, prevented any mechanistic<br />

interpretation <strong>of</strong> the observed photoreactions as well as <strong>of</strong> the few<br />

scattered reports on luminescent complexes.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 3<br />

After the Second World War, the interpretation <strong>of</strong> the absorption spectra<br />

started thanks to the development <strong>of</strong> the lig<strong>and</strong> field theory [5, 6] <strong>and</strong> the first<br />

attempts to rationalize the charge-transfer b<strong>and</strong>s [7, 8]. Following these developments,<br />

the photochemistry <strong>of</strong> coordination compounds could take its first<br />

steps as a modern science <strong>and</strong> in a time span <strong>of</strong> 2 years four important laboratories<br />

published their first photochemical paper [9–12]. Much <strong>of</strong> the attention<br />

was focused on Cr(III) complexes, whose luminescence was also investigated<br />

in some detail [13]. Later, Co(III) complexes attracted a great deal <strong>of</strong> attention<br />

since their photochemical behavior was found to change drastically with<br />

excitation wavelength [14, 15]. A few, isolated flash photolysis investigations<br />

began to appear, but this technique remained unavailable to most inorganic<br />

photochemists for several years.<br />

Since the late 1960s, the great development <strong>of</strong> photochemical <strong>and</strong> luminescence<br />

investigations on organic compounds led to the publication <strong>of</strong><br />

books [16–19] illustrating fundamental photochemical concepts that were<br />

also quickly exploited for coordination compounds [2]. From that period, it<br />

became common to discuss the photochemical <strong>and</strong> photophysical behavior <strong>of</strong><br />

a species (be it an organic molecule or a charged metal complex) on the basis<br />

<strong>of</strong> electronic configurations, selection rules, <strong>and</strong> energy level diagram, as we<br />

do today.<br />

2<br />

Molecular <strong>Photochemistry</strong><br />

Molecules are multielectron systems. Approximate electronic wavefunctions<br />

<strong>of</strong> a molecule can be written as products <strong>of</strong> one-electron wavefunctions, each<br />

consisting <strong>of</strong> an orbital <strong>and</strong> a spin part:<br />

Ψ = ΦS = Πiϕisi . (1)<br />

The ϕis are appropriate molecular orbitals (MOs) <strong>and</strong> si is one <strong>of</strong> the two<br />

possible spin eigenfunctions, α or β. The orbital part <strong>of</strong> this multielectron<br />

wavefunction defines the electronic configuration.<br />

We illustrate now the procedure to construct energy level diagrams, using<br />

as examples an organic molecule <strong>and</strong> a few coordination compounds.<br />

2.1<br />

Organic Molecules<br />

The MO diagram for formaldehyde is shown in Fig. 1 [20]. It consists <strong>of</strong> three<br />

low-lying σ-bonding orbitals, a π-bonding orbital <strong>of</strong> the CO group, a nonbonding<br />

orbital n <strong>of</strong> the oxygen atom (highest occupied molecular orbital,<br />

HOMO), a π-antibonding orbital <strong>of</strong> the CO group (lowest unoccupied molecular<br />

orbital, LUMO), <strong>and</strong> three high-energy σ-antibonding orbitals. The


4 V. Balzani et al.<br />

Fig. 1 Molecular orbital diagram for formaldehyde. The arrows indicate the n → π ∗ <strong>and</strong><br />

π → π ∗ transitions<br />

lowest-energy electronic configuration is (neglecting the filled low-energy orbitals)<br />

π 2 n 2 . Excited configurations can be obtained from the ground configuration<br />

by promoting one electron from occupied to vacant MOs. At relatively<br />

low energies, one expects to find n → π ∗ <strong>and</strong> π → π ∗ electronic transitions<br />

(Fig. 1), leading to π 2 nπ ∗ <strong>and</strong> πn 2 π ∗ excited configurations (Fig. 2a).<br />

In a very crude zero-order description, the energy associated with a particular<br />

electronic configuration would be given by the sum <strong>of</strong> the energies<br />

<strong>of</strong> the occupied MOs. In order to obtain a more realistic description <strong>of</strong> the<br />

energy states <strong>of</strong> the molecule, two features should be added to the simple<br />

configuration picture: (1) spin functions must be attached to the orbital<br />

functions describing the electronic configurations, <strong>and</strong> (2) interelectronic repulsion<br />

must be taken into account. These two closely interlocked points have<br />

important consequences, since they may lead to the splitting <strong>of</strong> an electronic<br />

configuration into several states.<br />

In the case <strong>of</strong> formaldehyde, the inclusion <strong>of</strong> spin <strong>and</strong> electronic repulsion<br />

leads to the schematic energy level diagram shown in Fig. 2b: each excited<br />

electronic configuration is split into a pair <strong>of</strong> triplet <strong>and</strong> singlet states, with<br />

the latter at higher energy because electronic repulsion is higher for spinpaired<br />

electrons. It can be noticed that the singlet–triplet splitting for the<br />

states arising from the ππ ∗ configuration is larger than that <strong>of</strong> the states cor-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 5<br />

Fig. 2 Configurations (a) <strong>and</strong>states(b) diagrams for formaldehyde<br />

responding to the nπ ∗ configuration. This result arises from the dependence<br />

<strong>of</strong> the interelectronic repulsions on the amount <strong>of</strong> spatial overlap between the<br />

MOs containing the two electrons, <strong>and</strong> this overlap is greater in the first than<br />

in the second case (see the MO shapes in Fig. 1). The electronic states can<br />

be designated by symbols that specify the symmetry <strong>of</strong> the wavefunction in<br />

thesymmetrygroup<strong>of</strong>themolecule(e.g.,A1, A2, etc.intheC2v group <strong>of</strong><br />

formaldehyde) <strong>and</strong> the spin multiplicity (number <strong>of</strong> unpaired electrons + 1)<br />

as a left superscript. In organic photochemistry, it is customary to label the<br />

singlet <strong>and</strong> triplet states as Sn <strong>and</strong> Tn, respectively, with n =0forthesinglet<br />

ground state <strong>and</strong> n =1,2,etc.forstatesarisingfromthevariousexcited<br />

configurations (<strong>of</strong>ten indicated in parentheses). Both notations are shown for<br />

formaldehyde in Fig. 2b. The situation sketched above (i.e., singlet ground<br />

state, pairs <strong>of</strong> singlet <strong>and</strong> triplet excited states arising from each excited configuration,<br />

lowest excited state <strong>of</strong> multiplicity higher than the ground state) is<br />

quite general for organic molecules that usually exhibit a closed-shell groundstate<br />

configuration.<br />

State energy diagrams <strong>of</strong> this type, usually called “Jablonski diagrams”,<br />

are used for the description <strong>of</strong> light absorption <strong>and</strong> <strong>of</strong> the photophysical processes<br />

that follow light excitation (vide infra).<br />

2.2<br />

Metal Complexes<br />

For metal complexes, the construction <strong>of</strong> Jablonski diagrams via electronic<br />

configurations from the MO description follows the same general lines described<br />

above for organic molecules [2]. A schematic MO diagram for an<br />

octahedral transition metal complex is shown in Fig. 3. The various MOs<br />

can be conveniently classified according to their predominant atomic orbital


6 V. Balzani et al.<br />

Fig. 3 Molecular orbital diagram for an octahedral complex <strong>of</strong> a transition metal. The<br />

arrows indicate the four types <strong>of</strong> transitions based on localized MO configurations. For<br />

more details, see text<br />

contributions as: (1) strongly bonding, predominantly lig<strong>and</strong> centered σL orbitals;<br />

(2) bonding, predominantly lig<strong>and</strong>-centered πL orbitals; (3) essentially<br />

nonbonding, metal-centered πM orbitals <strong>of</strong> t2g symmetry; (4) antibonding,<br />

predominantly metal-centered σ ∗ M orbitals <strong>of</strong> eg symmetry; (5) antibonding,<br />

predominantly lig<strong>and</strong>-centered π∗ L orbitals; <strong>and</strong> (6) strongly antibonding,<br />

predominantly metal-centered σ ∗ M<br />

orbitals. In the ground electronic configu-<br />

ration <strong>of</strong> an octahedral complex <strong>of</strong> a d n metalion,orbitals<strong>of</strong>types1<strong>and</strong>2are<br />

completely filled, while n electrons reside in the orbitals <strong>of</strong> types 3 <strong>and</strong> 4.<br />

As for organic molecules, excited configurations can be obtained from the<br />

ground configuration by promoting one electron from occupied to vacant<br />

MOs. At relatively low energies, one expects to find electronic transitions <strong>of</strong><br />

the following types (Fig. 3): metal-centered (MC) transitions from orbitals <strong>of</strong><br />

type 3 to orbitals <strong>of</strong> type 4; lig<strong>and</strong>-centered (LC) transitions <strong>of</strong> type 2 →5;<br />

lig<strong>and</strong>-to-metal charge-transfer (LMCT) transitions, e.g., <strong>of</strong> type 2 →4; <strong>and</strong><br />

metal-to-lig<strong>and</strong> charge-transfer (MLCT) transitions, e.g., <strong>of</strong> type 3 →5. The<br />

relative energy ordering <strong>of</strong> the resulting excited electronic configurations depends<br />

on the nature <strong>of</strong> metal <strong>and</strong> lig<strong>and</strong>s in more or less predictable ways.<br />

Low-energy metal-centered transitions are expected for metals <strong>of</strong> the first<br />

transition row, low-energy lig<strong>and</strong>-to-metal charge-transfer transitions are expected<br />

when at least one <strong>of</strong> the lig<strong>and</strong>s is easy to oxidize <strong>and</strong> the metal is<br />

easy to reduce, low-energy metal-to-lig<strong>and</strong> charge-transfer transitions are expected<br />

when the metal is easy to oxidize <strong>and</strong> a lig<strong>and</strong> is easy to reduce,


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 7<br />

<strong>and</strong> low-energy lig<strong>and</strong>-centered transitions are expected for aromatic lig<strong>and</strong>s<br />

with extended π <strong>and</strong> π ∗ orbitals (vide infra).<br />

The step from configurations to states is conceptually less simple than for<br />

organic molecules because coordination compounds may have high symmetry<br />

(i.e., degenerate MOs) <strong>and</strong> open-shell ground configurations (i.e., partially<br />

occupied HOMOs).<br />

For octahedral complexes <strong>of</strong> Co(III), Ru(II), <strong>and</strong> the other d 6 metal ions,<br />

the σL <strong>and</strong> πL orbitals are fully occupied <strong>and</strong> the ground-state configuration<br />

is closed-shell since the HOMO, πM(t2g) 6 , is also completely occupied.<br />

The ground state is therefore a singlet, <strong>and</strong> the excited states are either singlets<br />

or triplets, as in the case <strong>of</strong> formaldehyde. In octahedral symmetry,<br />

the ground-state configuration gives rise to the state 1 A1g. Inthecase<strong>of</strong><br />

[M(NH3)6] n+ complexes (e.g., M = Co or Ru), whose lig<strong>and</strong>s do not possess<br />

orbitals, the lowest-energy transition is metal centered <strong>and</strong> the re-<br />

πL <strong>and</strong> π∗ L<br />

sulting πM(t2g) 5σ ∗ M (eg) configuration gives rise to the singlet states 1S1g <strong>and</strong><br />

1S2g <strong>and</strong> the corresponding triplets 3T1g <strong>and</strong> 3T2g. The energy level diagram<br />

(at low energies) for [Ru(NH3)6] 2+ is shown in Fig. 4 (the triplet-state energy<br />

has been obtained by comparison with the analogous Ir(III) complex [21]).<br />

In the case <strong>of</strong> [M(bpy)3] 2+ (M = Ru or Os), however, since the M(II) metal<br />

is easy to oxidize <strong>and</strong> the 2,2 ′ -bipyridine lig<strong>and</strong>s are easy to reduce, the lowest<br />

triplet <strong>and</strong> singlet excited states are metal-to-lig<strong>and</strong> charge-transfer in<br />

character (Fig. 4). For the corresponding [M(bpy)3] 3+ complexes, the lowest<br />

triplet <strong>and</strong> singlet excited states are lig<strong>and</strong>-to-metal charge-transfer in character<br />

[22], since the M(III) metal can be easily reduced <strong>and</strong> the 2,2 ′ -bipyridine<br />

lig<strong>and</strong>s are not too difficult to oxidize (Fig. 4).<br />

In Cr(III) complexes (d3 metal ion), there are three electrons in the HOMO<br />

πM(t2g) orbitals. Therefore, these complexes exhibit an open-shell groundstate<br />

configuration, πM(t2g) 3 , that splits into quartet <strong>and</strong> doublet states<br />

Fig. 4 Schematic energy level diagrams for [Ru(NH3)6] 2+ , [Ru(bpy)3] 2+ ,<strong>and</strong>[Ru(bpy)3] 3+


8 V. Balzani et al.<br />

(Fig. 5). For most Cr(III) complexes, e.g., for [Cr(NH3)6] 3+ ,thelowest-energy<br />

transition is metal centered <strong>and</strong> the resulting πM(t2g) 2 σ ∗ M (eg) configuration<br />

gives rise to 4 T2g <strong>and</strong> 4 T1g excited states (Fig. 5). Several other coordination<br />

compounds, including the complexes <strong>of</strong> the lanthanide ions, have an openshell<br />

ground-state configuration <strong>and</strong>, as a consequence, a ground state with<br />

high-multiplicity <strong>and</strong> low-energy intraconfigurational metal-centered excited<br />

states.<br />

Fig. 5 Configurations (a) <strong>and</strong>state(b) diagrams for an octahedral Cr(III) complex. Only<br />

the lower-lying excited states <strong>of</strong> each configuration are shown [20]<br />

In conclusion, metal complexes tend to have more complex <strong>and</strong> specific<br />

Jablonski diagrams than organic molecules. Points to be noticed are: (1) spin<br />

multiplicity other than singlet <strong>and</strong> triplet can occur, but for each electronic<br />

configuration the state with highest multiplicity remains the lowest one;<br />

(2) excited states can exist that belong to the same configuration <strong>of</strong> the ground<br />

state (this implies that the ground state has the highest multiplicity); <strong>and</strong><br />

(3) more than one pair <strong>of</strong> states <strong>of</strong> different multiplicity can arise from a single<br />

electron configuration. In the following, in order to discuss some general<br />

concept <strong>of</strong> molecular photochemistry we will make use <strong>of</strong> a generic Jablonski<br />

diagram based on singlet <strong>and</strong> triplet states.<br />

2.3<br />

Light Absorption <strong>and</strong> Intramolecular Excited-State Decay<br />

Figure 6 shows a schematic energy level diagram for a generic molecule [23].<br />

In principle, transitions between states having the same multiplicity are allowed,<br />

whereas those between states <strong>of</strong> different multiplicity are forbidden.<br />

Therefore, the electronic absorption b<strong>and</strong>s observed in the UV–visible spec-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 9<br />

Fig. 6 Schematic energy level diagram for a generic molecule<br />

trum <strong>of</strong> such a generic molecule would display b<strong>and</strong>s corresponding to the<br />

S0 → Sn transitions <strong>of</strong> the diagram. For metal complexes, which usually are<br />

highly symmetric species, symmetry selection rules can also play a role in determining<br />

the intensity <strong>of</strong> the absorption b<strong>and</strong>s. Furthermore, the presence <strong>of</strong><br />

a heavy atom (namely, the metal) relaxes the spin-conservation rule.<br />

The excited states are unstable species that decay not only by intramolecular<br />

chemical reactions (e.g., dissociation, isomerization) but also (actually,<br />

more <strong>of</strong>ten) by intramolecular radiative <strong>and</strong> nonradiative deactivations.<br />

When a species is excited to upper spin-allowed excited states, it usually<br />

undergoes a fast <strong>and</strong> 100% efficient radiationless deactivation (internal conversion,<br />

ic) to the lowest spin-allowed excited (S1 in Fig. 6). Setting aside the<br />

intramolecular photochemical processes, such an excited state undergoes deactivation<br />

via three competing first-order processes: nonradiative decay to the<br />

ground state (internal conversion, rate constant kic); radiative decay to the<br />

ground state (fluorescence, kfl); <strong>and</strong> intersystem crossing (isc) to the lowest<br />

triplet state T1 (kisc). In its turn, T1 can undergo deactivation via nonradiative<br />

(intersystem crossing, k ′ isc ) or radiative (phosphorescence, kph) decayto<br />

the ground state S0. When the species contains heavy atoms, as in the case<br />

<strong>of</strong> metal complexes, the formally forbidden intersystem crossing <strong>and</strong> phosphorescence<br />

processes become faster. The lifetime (τ) <strong>of</strong> an excited state, i.e.,<br />

the time needed to reduce the excited-state concentration by 2.718, is given


10 V. Balzani et al.<br />

by the reciprocal <strong>of</strong> the summation <strong>of</strong> the deactivation rate constants. For the<br />

molecule <strong>of</strong> Fig. 6,<br />

τ(S1)=<br />

1<br />

(kic + kfl + kisc)<br />

1<br />

τ(T1)=<br />

(k ′ . (3)<br />

isc + kph)<br />

The lifetimes <strong>of</strong> the lowest spin-allowed <strong>and</strong> spin-forbidden excited state<br />

(τ(S1) <strong>and</strong>τ(T1) in the example <strong>of</strong> Fig. 6) are approximately 10 –9 –10 –7 s<strong>and</strong><br />

10 –3 –100 s, respectively, for organic molecules, but they become shorter by<br />

several orders <strong>of</strong> magnitude for metal complexes. For example, the lifetime <strong>of</strong><br />

the lowest spin-forbidden excited state <strong>of</strong> naphthalene is around 2 s, whereas<br />

that <strong>of</strong> [Ru(bpy)3] 2+ , because <strong>of</strong> the presence <strong>of</strong> the heavy Ru ion, is about<br />

1 µs [24].<br />

The quantum yields <strong>of</strong> fluorescence (ratio between the number <strong>of</strong> photons<br />

emitted by the lowest spin-allowed excited state, S1 in Fig. 6, <strong>and</strong> the number<br />

<strong>of</strong> absorbed photons) <strong>and</strong> phosphorescence (ratio between the number <strong>of</strong><br />

photons emitted by the lowest spin-forbidden excited state, T1 in Fig. 6, <strong>and</strong><br />

the number <strong>of</strong> absorbed photons) can range between 0 <strong>and</strong> 1 <strong>and</strong> are given<br />

by the following expressions:<br />

Φ fl =<br />

k fl<br />

(kic + k fl + kisc)<br />

kph × kisc<br />

Φph =<br />

(k ′ isc + kph)<br />

. (5)<br />

× (kic + kfl + kisc)<br />

The excited-state lifetimes <strong>and</strong> fluorescence <strong>and</strong> phosphorescence quantum<br />

yields <strong>of</strong> a great number <strong>of</strong> organic molecules <strong>and</strong> metal complexes are<br />

known [24].<br />

3<br />

Bimolecular Processes<br />

3.1<br />

General Features<br />

In fluid solution, when the intramolecular deactivation processes are not too<br />

fast, i.e., when the lifetime <strong>of</strong> the excited state is sufficiently long, an excited<br />

molecule ∗ A may have a chance to encounter a molecule <strong>of</strong> another solute B.<br />

In such a case, some specific interaction can occur leading to the deactivation<br />

<strong>of</strong> the excited state by second-order kinetic processes. The two most important<br />

types <strong>of</strong> interactions in an encounter are those leading to electron or<br />

(2)<br />

(4)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 11<br />

energy transfer:<br />

∗ A+B→ A + +B –<br />

∗ A+B→ A – +B +<br />

oxidative electron transfer (6)<br />

reductive electron transfer (7)<br />

∗ ∗<br />

A+B→ A+ B energy transfer . (8)<br />

Bimolecular electron- <strong>and</strong> energy-transfer processes are important because<br />

they can be used (1) to quench an electronically excited state, i.e., to prevent<br />

its luminescence <strong>and</strong>/or intramolecular reactivity, <strong>and</strong> (2) to sensitize other<br />

species, for example, to cause chemical changes <strong>of</strong>, or luminescence from,<br />

species that do not absorb light.<br />

Simple kinetic arguments show that only the excited states that live longer<br />

than ca. 10 –9 s may have a chance to be involved in encounters with other solute<br />

molecules. Usually, in the case <strong>of</strong> metal complexes only the lowest excited<br />

state satisfies this requirement. The kinetic aspects <strong>of</strong> energy- <strong>and</strong> electrontransfer<br />

processes are discussed in detail elsewhere [17, 20, 23]. A point that<br />

must be stressed is that an electronically excited state is a species with quite<br />

different properties compared with those <strong>of</strong> the ground-state molecule. In<br />

particular, because <strong>of</strong> its higher energy content, an excited state is both<br />

a stronger reductant <strong>and</strong> a stronger oxidant than the corresponding ground<br />

state. To a first approximation, the redox potentials <strong>of</strong> the excited-state couples<br />

may be calculated from the potentials <strong>of</strong> the ground-state couples <strong>and</strong> the<br />

one-electron potential corresponding to the zero–zero excited-state energy,<br />

E0–0 , as shown by Eqs. 9 <strong>and</strong> 10 [25]:<br />

E � A + / ∗ A � ≈ E � A + /A � – E 0–0<br />

E � ∗ A/A – � ≈ E � A/A –� + E 0–0 . (10)<br />

3.2<br />

Bimolecular Processes Involving Metal Complexes<br />

From an exhaustive monograph that appeared in 1970 [2] <strong>and</strong> a multiauthored<br />

volume <strong>of</strong> 1975 [26], it clearly appears that most <strong>of</strong> the interest was<br />

then focused on lig<strong>and</strong> photosubstitution reactions, photoredox decomposition,<br />

<strong>and</strong> photoisomerization reactions, while bimolecular processes were<br />

barely investigated. This picture, however, changed pr<strong>of</strong>oundly in a few years<br />

following the extensive work carried out by several research groups on the luminescence<br />

<strong>of</strong> coordination compounds [27–29] <strong>and</strong> the discovery that the<br />

lowest excited state <strong>of</strong> a number <strong>of</strong> Cr(III), Ru(II), <strong>and</strong> Os(II) complexes exhibits<br />

a sufficiently long lifetime in fluid solution to be able to participate<br />

as a reactant in bimolecular reactions [25, 30]. A further advantage <strong>of</strong>fered<br />

by Ru(II) <strong>and</strong> Os(II) bipyridine-type complexes is that they can undergo reversible<br />

redox reactions both in the ground <strong>and</strong> excited state, so they were<br />

soon used as reactants <strong>and</strong>, even more interesting, as mediators, in light-<br />

(9)


12 V. Balzani et al.<br />

induced [30–34] <strong>and</strong> light-generating [35, 36] electron-transfer processes.<br />

Such studies were further boosted by the fact that, after the energy crisis <strong>of</strong><br />

the early 1970s, several photochemists became involved in the problem <strong>of</strong> solar<br />

energy conversion. Particular interest arose around photosensitized water<br />

splitting [37–41] <strong>and</strong> it was soon realized [42] that [Ru(bpy)3] 2+ <strong>and</strong> related<br />

complexes, because <strong>of</strong> their excited-state redox properties, might function as<br />

photocatalysts for such a process.<br />

As a matter <strong>of</strong> fact, in the period 1975–1985 a real revolution occurred in<br />

the field <strong>of</strong> the photochemistry <strong>of</strong> coordination compounds. The study <strong>of</strong> intramolecular<br />

lig<strong>and</strong> photosubstitution, photoredox decomposition, <strong>and</strong> photoisomerization<br />

reactions was almost completely set apart, about 300 Ru(II)<br />

bipyridine-type complexes were synthesized <strong>and</strong> investigated in an attempt<br />

(mostly vain) to improve the already outst<strong>and</strong>ing excited-state properties<br />

<strong>of</strong> [Ru(bpy)3] 2+ [43], <strong>and</strong>, thanks to an extensive use <strong>of</strong> pulsed techniques,<br />

huge amounts <strong>of</strong> data were collected on the rate constants <strong>of</strong> bimolecular<br />

processes [44]. The high exergonicity <strong>of</strong> the excited-state electron-transfer<br />

reactions (<strong>and</strong>/or <strong>of</strong> their back reactions) <strong>of</strong>fered the opportunity for the<br />

first time to investigate some fundamental aspects <strong>of</strong> electron-transfer theories<br />

[45], with particular attention to the so-called Marcus inverted region.<br />

4<br />

Supramolecular <strong>Photochemistry</strong><br />

4.1<br />

Operational Definition <strong>of</strong> Supramolecular Species<br />

In the late 1980s, following the award <strong>of</strong> the 1987 Nobel prize to Pedersen,<br />

Cram, <strong>and</strong> Lehn, there was a sudden increase <strong>of</strong> interest in supramolecular<br />

chemistry, a highly interdisciplinary field based on concepts such as molecular<br />

recognition, preorganization, <strong>and</strong> self-assembling.<br />

The classical definition <strong>of</strong> supramolecular chemistry is that given by<br />

J.-M. Lehn, namely “the chemistry beyond the molecule, bearing on organized<br />

entities <strong>of</strong> higher complexity that result from the association <strong>of</strong> two<br />

or more chemical species held together by intermolecular forces” [46]. There<br />

is, however, a problem with this definition. With supramolecular chemistry<br />

there has been a change in focus from molecules to molecular assemblies<br />

or multicomponent systems. According to the original definition, however,<br />

when the components <strong>of</strong> a chemical system are linked by covalent bonds, the<br />

system should not be considered a supramolecular species, but a molecule.<br />

This point is particularly important in dealing with light-powered molecular<br />

devices <strong>and</strong> machines (vide infra), which are usually multicomponent systems<br />

in which the components can be linked by chemical bonds <strong>of</strong> various<br />

natures.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 13<br />

To better underst<strong>and</strong> this point, consider, for example, the three systems<br />

[47] shown in Fig. 7, which play the role <strong>of</strong> photoinduced chargeseparation<br />

molecular devices [48]. In each one <strong>of</strong> them, two components,<br />

a Zn(II) porphyrin <strong>and</strong> an Fe(III) porphyrin, can be immediately singled<br />

out. In 1, these two components are linked by a hydrogen-bonded bridge,<br />

i.e., by intermolecular forces, whereas in 2 <strong>and</strong> 3 they are linked by covalent<br />

bonds. According to the above-reported classical definition <strong>of</strong> supramolecular<br />

chemistry, 1 is a supramolecular species, whereas 2 <strong>and</strong> 3 are (large)<br />

molecules. In each one <strong>of</strong> the three systems, the two components substantially<br />

maintain their intrinsic properties <strong>and</strong>, upon light excitation, electron<br />

transfer takes place from the Zn(II) porphyrin unit to the Fe(III) porphyrin<br />

one. The values <strong>of</strong> the rate constants for photoinduced electron transfer<br />

(k el = 8.1 × 10 9 , 8.8 × 10 9 ,<strong>and</strong>4.3 × 10 9 s –1 for 1, 2, <strong>and</strong>3, respectively) show<br />

that the electronic interaction between the two components in 1 is comparable<br />

to that in 2, <strong>and</strong> is slightly stronger than that in 3. Clearly, as far as<br />

photoinduced electron transfer is concerned, it would sound strange to say<br />

that 1 is a supramolecular species, <strong>and</strong> 2 <strong>and</strong> 3 are molecules.<br />

Fig. 7 Three dyads possessing Zn(II) porphyrin <strong>and</strong> Fe(III) porphyrin units linked by an<br />

H-bonded bridge (1), a partially unsaturated bridge (2), <strong>and</strong> a saturated bridge (3) [47]


14 V. Balzani et al.<br />

The example discussed above shows that, although the classical definition<br />

<strong>of</strong> supramolecular chemistry as “the chemistry beyond the molecule” [46] is<br />

quite useful in general, from a functional viewpoint the distinction between<br />

what is molecular <strong>and</strong> what is supramolecular can be better based on the degree<br />

<strong>of</strong> intercomponent electronic interactions [20, 49–52]. This concept is<br />

illustrated, for example, in Fig. 8 [53]. In the case <strong>of</strong> photon stimulation, a system<br />

A∼B, consisting <strong>of</strong> two units (∼ indicates any type <strong>of</strong> “bond” that keeps<br />

the units together), can be defined as a supramolecular species if light absorption<br />

leads to excited states that are substantially localized on either A or B, or<br />

causes an electron transfer from A to B (or vice versa). By contrast, when the<br />

excited states are substantially delocalized on the entire system, the species<br />

can be better considered as a large molecule. Similarly (Fig. 8), oxidation <strong>and</strong><br />

reduction <strong>of</strong> a supramolecular species can substantially be described as oxidation<br />

<strong>and</strong> reduction <strong>of</strong> specific units, whereas oxidation <strong>and</strong> reduction <strong>of</strong><br />

a large molecule leads to species where the hole or the electron are delocalized<br />

on the entire system. In more general terms, when the interaction energy<br />

between units is small compared to the other relevant energy parameters,<br />

a system can be considered a supramolecular species, regardless <strong>of</strong> the nature<br />

<strong>of</strong> the bonds that link the units. Species made <strong>of</strong> covalently linked (but<br />

weakly interacting) components, e.g., 2 <strong>and</strong> 3 showninFig.7,cantherefore<br />

be regarded as belonging to the supramolecular domain when they are stimulated<br />

by photons or electrons. It should be noted that the properties <strong>of</strong> each<br />

component <strong>of</strong> a supramolecular species, i.e., <strong>of</strong> an assembly <strong>of</strong> weakly interacting<br />

molecular components, can be known from the study <strong>of</strong> the isolated<br />

components or <strong>of</strong> suitable model molecules.<br />

Fig. 8 Schematic representation <strong>of</strong> the difference between a supramolecular system <strong>and</strong><br />

alargemoleculebasedontheeffectscausedbyaphotonoranelectroninput.Formore<br />

details, see text


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 15<br />

4.2<br />

Photoinduced Processes in Supramolecular Systems<br />

Clearly, preorganization is a most valuable property from a photochemical<br />

viewpoint since a supramolecular system, for example, can be preorganized<br />

so as to favor the occurrence <strong>of</strong> energy- <strong>and</strong> electron-transfer processes [20].<br />

Consider, for example, an A–L–B supramolecular system, where A is the<br />

light-absorbing molecular unit (Eq. 11), B is the other molecular unit to be involved<br />

in the light-induced processes, <strong>and</strong> L is a connecting unit (<strong>of</strong>ten called<br />

bridge). In such a system, after light excitation <strong>of</strong> A there is no need to wait<br />

for a diffusion-controlled encounter between ∗ A <strong>and</strong> B, as in molecular photochemistry<br />

(vide supra), since the two reaction partners can already be at an<br />

interaction distance suitable for electron <strong>and</strong> energy transfer:<br />

A–L–B + hν → ∗ A–L–B photoexcitation (11)<br />

∗ A–L–B → A + –L–B –<br />

∗ A–L–B → A – –L–B +<br />

oxidative electron transfer (12)<br />

reductive electron transfer (13)<br />

∗ A–L–B → A–L– ∗ B electronic energy transfer . (14)<br />

In the absence <strong>of</strong> chemical complications (e.g., fast decomposition <strong>of</strong> the<br />

oxidized <strong>and</strong>/or reduced species), photoinduced electron-transfer processes<br />

are followed by spontaneous back electron-transfer reactions that regenerate<br />

the starting ground-state system (Eqs. 15 <strong>and</strong> 16), <strong>and</strong> photoinduced energy<br />

transfer is followed by radiative <strong>and</strong>/or nonradiative deactivation <strong>of</strong> the excited<br />

acceptor (Eq. 17):<br />

A + –L–B – → A–L–B back oxidative electron transfer (15)<br />

A – –L–B + → A–L–B back reductive electron transfer (16)<br />

A–L– ∗ B → A–L–B excited-state decay . (17)<br />

In supramolecular systems, electron- <strong>and</strong> energy-transfer processes take<br />

place by first-order kinetics. As a consequence, in suitably designed supramolecular<br />

systems these processes can involve even very short lived excited<br />

states.<br />

In most cases, the interaction between excited <strong>and</strong> ground-state components<br />

in a supramolecular system is weak. When the interaction is strong,<br />

new chemical species are formed, which are called excimers (from excited<br />

dimers) or exciplexes (from excited complexes), depending on whether the<br />

two interacting units have the same or different chemical nature (Fig. 9). It<br />

is important to notice that excimer <strong>and</strong> exciplex formation is a reversible<br />

process <strong>and</strong> that both excimers <strong>and</strong> exciplexes are sometimes luminescent.


16 V. Balzani et al.<br />

Fig. 9 Schematic representation <strong>of</strong> excimer <strong>and</strong> exciplex formation in a supramolecular<br />

system<br />

Compared with the “monomer” emission, the emission <strong>of</strong> an excimer or exciplex<br />

is always displaced to lower energy (longer wavelengths) <strong>and</strong> usually<br />

corresponds to a broad <strong>and</strong> rather weak b<strong>and</strong>.<br />

Excimers are usually obtained when an excited state <strong>of</strong> an aromatic<br />

molecule interacts with the ground state <strong>of</strong> a molecule <strong>of</strong> the same type. For<br />

example, between the excited <strong>and</strong> ground states <strong>of</strong> anthracene units. Exciplexes<br />

are obtained when an electron donor (acceptor) excited state interacts<br />

with an electron acceptor (donor) ground-state molecule; for example, between<br />

excited states <strong>of</strong> aromatic molecules (electron acceptors) <strong>and</strong> amines<br />

(electron donors). Excited states <strong>of</strong> coordination compounds are seldom involved<br />

in excimers or exciplexes, since their components (metal <strong>and</strong> lig<strong>and</strong>s)<br />

have already used their electron donor or acceptor properties in forming<br />

the complex. Furthermore, the three-dimensional structure <strong>of</strong> coordination<br />

compounds usually prevents strong electronic interaction with other species.<br />

However, for some square planar complexes excimer emission has long been<br />

reported [54] <strong>and</strong> can indeed be found for some families <strong>of</strong> Au <strong>and</strong> Pt complexes,<br />

as discussed in other chapters <strong>of</strong> this volume.<br />

The working mechanisms <strong>of</strong> a number <strong>of</strong> biological <strong>and</strong> artificial molecular<br />

devices <strong>and</strong> machines are based on photoinduced electron- <strong>and</strong> energytransfer<br />

processes [20, 48, 55]. Since these processes have to compete with<br />

the intrinsic decays <strong>of</strong> the relevant excited states, a key problem is that <strong>of</strong><br />

maximizing their rates. It is therefore appropriate to summarize some basic<br />

principles <strong>of</strong> electron- <strong>and</strong> energy-transfer kinetics. [56].<br />

4.3<br />

Electron Transfer<br />

4.3.1<br />

Marcus Theory<br />

Electron-transfer processes involving excited-state <strong>and</strong>/or ground-state molecules<br />

can be dealt with in the frame <strong>of</strong> the Marcus theory [57] <strong>and</strong> <strong>of</strong> the


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 17<br />

successive, more sophisticated theoretical models [58, 59]. Of course, when<br />

excited states are involved, the redox potential <strong>of</strong> the excited-state couple has<br />

to be used (Eqs. 9 <strong>and</strong> 10).<br />

According to the Marcus theory [57], the rate constant for an electrontransfer<br />

process can be expressed as<br />

κel = νNκel exp<br />

�<br />

– ∆G‡<br />

RT<br />

�<br />

, (18)<br />

where νN is the average nuclear frequency factor, κel is the electronic transmission<br />

coefficient, <strong>and</strong> ∆G ‡ is the free energy <strong>of</strong> activation. This last term<br />

can be expressed by the Marcus quadratic relationship<br />

∆G ‡ = λ<br />

�<br />

1+<br />

4<br />

∆G0<br />

�2<br />

, (19)<br />

λ<br />

where ∆G 0 is the st<strong>and</strong>ard free energy change <strong>of</strong> the reaction <strong>and</strong> λ is the<br />

nuclear reorganizational energy (Fig. 10). This equation predicts that for<br />

a homogeneous series <strong>of</strong> reactions (i.e., for reactions having the same λ <strong>and</strong><br />

κ el values), a ln k el vs ∆G 0 plot is a bell-shaped curve (Fig. 11) involving<br />

(1) a “normal” region for endoergonic <strong>and</strong> slightly exoergonic reactions, in<br />

which ln k el increases with increasing driving force; (2) an activationless maximum<br />

for λ ≈ – ∆G 0 ; <strong>and</strong> (3) an “inverted” region for strongly exoergonic<br />

reactions, in which ln kel decreases with increasing driving force.<br />

Fig. 10 Pr<strong>of</strong>ile <strong>of</strong> the potential energy curves <strong>of</strong> an electron-transfer reaction: i <strong>and</strong> f indicate<br />

the initial <strong>and</strong> final states <strong>of</strong> the system. The dashed curve indicates the final state<br />

for a self-exchange (isoergonic) process. For more details, see text


18 V. Balzani et al.<br />

Fig. 11 Free energy dependence <strong>of</strong> electron-transfer rate (i, initialstate;f ,finalstate)according<br />

to the Marcus (a) <strong>and</strong> quantum mechanical (b) treatments.Thethreekinetic<br />

regimes (normal, activationless, <strong>and</strong> “inverted”) are shown schematically in terms <strong>of</strong><br />

Marcus parabolas<br />

The reorganizational energy λ can be expressed as the sum <strong>of</strong> two independent<br />

contributions corresponding to the reorganization <strong>of</strong> the “inner” (bond<br />

lengths <strong>and</strong> angles within the two reaction partners) <strong>and</strong> “outer” (solvent<br />

reorientation around the reacting pair) nuclear modes:<br />

λ = λi + λo . (20)<br />

The electronic transmission coefficient κel is related to the probability <strong>of</strong><br />

crossing at the intersection region (Fig. 10). It can be expressed by the equation<br />

κel = 2 � 1–exp � ��<br />

– νel/2νN 2–exp � � , (21)<br />

– νel/2νN<br />

where<br />

νel =<br />

�<br />

2 Hel� 2<br />

h<br />

� π 3<br />

λRT<br />

�1/2<br />

, (22)<br />

<strong>and</strong> Hel is the matrix element for electronic interaction (Fig. 10).<br />

If Hel is large, νel ≫ νN, κel =1<strong>and</strong><br />

�<br />

– ∆G ‡ �<br />

kel = νN exp<br />

(adiabatic limit) . (23)<br />

RT


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 19<br />

If Hel is small, νel ≪ νN, κel = νel/νN <strong>and</strong><br />

�<br />

– ∆G ‡ �<br />

kel = νel exp<br />

(nonadiabatic limit) . (24)<br />

RT<br />

Under the latter condition, kel is proportional to (H el ) 2 .Thevalue<strong>of</strong>H el depends<br />

on the overlap between the electronic wavefunctions <strong>of</strong> the donor <strong>and</strong><br />

acceptor groups, which should decrease exponentially with increasing donor–<br />

acceptor distance.<br />

It should be noticed that the amount <strong>of</strong> electronic interaction required to<br />

promote photoinduced electron transfer is very small. By substituting reasonable<br />

numbers for the parameters in Eq. 24, for an activationless reaction H el<br />

values <strong>of</strong> a few wavenumbers are sufficient to give rates in the sub-nanosecond<br />

timescale, while a few hundred wavenumbers may be sufficient to reach the<br />

limiting adiabatic regime (Eq. 23).<br />

4.3.2<br />

Quantum Mechanical Theory<br />

From a quantum mechanical viewpoint, both the photoinduced <strong>and</strong> back<br />

electron-transfer processes can be viewed as radiationless transitions between<br />

different, weakly interacting electronic states <strong>of</strong> the A–L–B supermolecule<br />

(Fig. 12). The rate constant <strong>of</strong> such processes is given by an appropriate<br />

Fermi “golden rule” expression:<br />

k el = 4π2<br />

h<br />

�<br />

H el� 2<br />

FC el , (25)<br />

Fig. 12 Electron-transfer processes in a supramolecular system: 1 photoexcitation; 2 photoinduced<br />

electron transfer; 3 thermal back electron transfer; 4 optical electron transfer


20 V. Balzani et al.<br />

where H el <strong>and</strong> FC el are the electronic coupling <strong>and</strong> the Franck–Condon density<br />

<strong>of</strong> states, respectively.<br />

In the absence <strong>of</strong> any intervening medium (through-space mechanism),<br />

the electronic factor decreases exponentially with increasing distance:<br />

H el = H el �<br />

(0) exp – βel<br />

2<br />

� �<br />

rAB – r0<br />

�<br />

, (26)<br />

where rAB is the donor–acceptor distance, H el (0) is the interaction at the “contact”<br />

distance r0,<strong>and</strong>β el is an appropriate attenuation parameter.<br />

For donor–acceptor components separated by vacuum, β el is estimated to<br />

be in the range 2–5 ˚A –1 . When donor <strong>and</strong> acceptor are separated by “matter”<br />

(e.g., a bridge L), the electronic coupling can be mediated by mixing <strong>of</strong><br />

the initial <strong>and</strong> final states <strong>of</strong> the system with virtual, high-energy electrontransfer<br />

states involving the intervening medium (superexchange mechanism)<br />

[60, 61].<br />

The FC el term <strong>of</strong> Eq. 25 is a thermally averaged Franck–Condon factor connecting<br />

the initial <strong>and</strong> final states. In the high temperature limit (hν


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 21<br />

The Hush theory [62] correlates the parameters that are involved in the<br />

corresponding thermal electron-transfer process by means <strong>of</strong> Eqs. 28–30:<br />

(28)<br />

∆ν1/2 = 48.06 � Eop – ∆G 0�1/2 (29)<br />

�<br />

εmax∆ν1/2 = H el� 2 r2 4.20 × 10 –4 ,<br />

Eop<br />

(30)<br />

Eop = λ + ∆G 0<br />

where Eop, ∆ν1/2 (both in cm –1 ), <strong>and</strong> εmax are the energy, halfwidth, <strong>and</strong><br />

maximum intensity <strong>of</strong> the electron-transfer b<strong>and</strong>, respectively, <strong>and</strong> r is the<br />

center-to-center distance. As shown by Eqs. 28–30, the energy depends on<br />

both reorganizational energy <strong>and</strong> thermodynamics, the halfwidth reflects the<br />

reorganizational energy, <strong>and</strong> the intensity <strong>of</strong> the transition is mainly related<br />

to the magnitude <strong>of</strong> the electronic coupling between the two redox centers.<br />

In principle, therefore, important kinetic information on a thermal<br />

electron-transfer process could be obtained from the study <strong>of</strong> the corresponding<br />

optical transition. In practice, it can be shown that weakly coupled<br />

systems may undergo relatively fast electron-transfer processes without exhibiting<br />

appreciably intense optical electron-transfer b<strong>and</strong>s. More details on<br />

optical electron transfer <strong>and</strong> related topics (i.e., mixed valence metal complexes)<br />

can be found in the literature [63–65].<br />

4.4<br />

Energy Transfer<br />

The thermodynamic ability <strong>of</strong> an excited state to intervene in energy-transfer<br />

processes is related to its zero–zero spectroscopic energy, E0–0 .Bimolecular<br />

energy-transfer processes involving encounters can formally be treated using<br />

a Marcus-type approach with ∆G0 = E0–0 A – E0–0<br />

B <strong>and</strong> λ ∼ λi [66].<br />

Energy transfer in a supramolecular system can be viewed as a radiationless<br />

transition between two “localized” electronically excited states. Therefore,<br />

the rate constant can again be obtained by an appropriate “golden rule”<br />

expression, similar to that seen above for electron transfer:<br />

ken = 4π2 � en<br />

H<br />

h<br />

�2 en<br />

FC , (31)<br />

where Hen is the electronic coupling between the two excited states interconverted<br />

by the energy-transfer process <strong>and</strong> FCen is an appropriate Franck–<br />

Condon factor. As for electron transfer, the Franck–Condon factor can be<br />

cast either in classical [67] or quantum mechanical [68–70] terms. Classically,<br />

it accounts for the combined effects <strong>of</strong> energy gradient <strong>and</strong> nuclear<br />

reorganization on the rate constant. In quantum mechanics terms, the FC<br />

factor is a thermally averaged sum <strong>of</strong> vibrational overlap integrals. Experimental<br />

information on this term can be obtained from the overlap integral


22 V. Balzani et al.<br />

between the emission spectrum <strong>of</strong> the donor <strong>and</strong> the absorption spectrum <strong>of</strong><br />

the acceptor.<br />

The electronic factor H en is a two-electron matrix element involving the<br />

HOMOs <strong>and</strong> LUMOs <strong>of</strong> the energy–donor <strong>and</strong> energy–acceptor components.<br />

By following st<strong>and</strong>ard arguments [20, 23], this factor can be split into two<br />

additive terms, a coulombic term <strong>and</strong> an exchange term. The two terms depend<br />

differently on the parameters <strong>of</strong> the system (spin <strong>of</strong> ground <strong>and</strong> excited<br />

states, donor–acceptor distance, etc.) <strong>and</strong> each <strong>of</strong> them can become predominant<br />

depending on the specific system <strong>and</strong> experimental conditions.<br />

The orbital aspects <strong>of</strong> the two mechanisms are schematically represented in<br />

Fig. 13.<br />

Fig. 13 Pictorial representation <strong>of</strong> the coulombic <strong>and</strong> exchange energy-transfer mechanisms<br />

4.4.1<br />

Coulombic Mechanism<br />

The coulombic (also called resonance, Förster-type [71], or through-space)<br />

mechanism is a long-range mechanism that does not require physical contact<br />

between donor <strong>and</strong> acceptor. It can be shown that the most important<br />

term within the coulombic interaction is the dipole–dipole term, which obeys<br />

the same selection rules as the corresponding electric dipole transitions <strong>of</strong><br />

the two partners ( ∗ A → A<strong>and</strong>B→ ∗ B, Fig. 13). Therefore, coulombic energy<br />

transfer is expected to be efficient in systems in which the radiative transitions<br />

connecting the ground <strong>and</strong> the excited states <strong>of</strong> each partner have high<br />

oscillator strength. The rate constant for the dipole–dipole coulombic energy


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 23<br />

transfer can be expressed as a function <strong>of</strong> the spectroscopic <strong>and</strong> photophysical<br />

properties <strong>of</strong> the two molecular components:<br />

k F en = 8.8 × 10–25 K2 Φ<br />

JF =<br />

n4r6 JF<br />

(32)<br />

ABτ � F(ν)ε(ν)/ν 4 dν<br />

� F(ν)dν<br />

, (33)<br />

where K is an orientation factor which accounts for the directional nature<br />

<strong>of</strong> the dipole–dipole interaction (K 2 =2/3 for r<strong>and</strong>om orientation), Φ <strong>and</strong> τ<br />

are the luminescence quantum yield <strong>and</strong> lifetime <strong>of</strong> the donor, respectively,<br />

n is the solvent refractive index, rAB is the distance (in ˚A) between donor<br />

<strong>and</strong> acceptor, <strong>and</strong> JF is the Förster overlap integral between the luminescence<br />

spectrum <strong>of</strong> the donor, F � ν � , <strong>and</strong> the absorption spectrum <strong>of</strong> the acceptor,<br />

ε � ν � ,onanenergyscale(cm –1 ). With a good spectral overlap integral <strong>and</strong><br />

appropriate photophysical properties, the 1/r6 AB distance dependence allows<br />

energy transfer to occur efficiently over distances largely exceeding the molecular<br />

diameters. The typical example <strong>of</strong> an efficient coulombic mechanism<br />

is that <strong>of</strong> singlet–singlet energy transfer between large aromatic molecules,<br />

a process used by nature in the “antenna” systems <strong>of</strong> the photosynthetic apparatus<br />

[72]:<br />

∗<br />

A(S1)–L–B(S0) → A(S0)–L– ∗ B(S1). (34)<br />

4.4.2<br />

Exchange Mechanism<br />

The exchange (also called Dexter-type [73]) mechanism requires orbital overlap<br />

between donor <strong>and</strong> acceptor, either directly or mediated by the bridge<br />

(through-bond), <strong>and</strong> its rate constant, therefore, decreases with increasing<br />

distance:<br />

k D 4π2 � en<br />

en = H<br />

h<br />

�2 JD , (35)<br />

where<br />

H en = H en �<br />

(0) exp – βen � �<br />

rAB – r0<br />

2<br />

�<br />

(36)<br />

� � � � �<br />

F ν ε ν dν<br />

JD = � � � � � � . (37)<br />

F ν dν ε ν dν<br />

The exchange interaction can be regarded (Fig. 13) as a double electrontransfer<br />

process, one electron moving from the LUMO <strong>of</strong> the excited donor<br />

to the LUMO <strong>of</strong> the acceptor, <strong>and</strong> the other from the acceptor HOMO to<br />

the donor HOMO. Therefore, the attenuation factor β en for exchange energy


24 V. Balzani et al.<br />

transfer should be approximately equal to the sum <strong>of</strong> the attenuation factors<br />

for two separated electron-transfer processes, i.e., βel for electron transfer between<br />

the LUMOs <strong>of</strong> the donor <strong>and</strong> acceptor, <strong>and</strong> βht for the electron transfer<br />

between the HOMOs (superscript ht is for hole transfer from the donor to the<br />

acceptor). This prediction has been confirmed by experiments [74].<br />

The spin selection rules for this type <strong>of</strong> mechanism arise from the need<br />

to obey spin conservation in the reacting pair as a whole. This allows the<br />

exchange mechanism to be operative in many cases in which the excited<br />

states involved are spin-forbidden in the usual spectroscopic sense. Thus, the<br />

typical example <strong>of</strong> an efficient exchange mechanism is that <strong>of</strong> triplet–triplet<br />

energy transfer:<br />

∗<br />

A(T1)–L – B(S0) → A(S0)–L – ∗ B(T1) . (38)<br />

Exchange energy transfer from the lowest spin-forbidden excited state is expected<br />

to be the rule for metal complexes [61, 75].<br />

Although the exchange mechanism was originally formulated in terms <strong>of</strong><br />

direct overlap between donor <strong>and</strong> acceptor orbitals, it is clear that it can be<br />

extended to cover the case in which coupling is mediated by the intervening<br />

medium (i.e., the connecting bridge), as discussed above for electron-transfer<br />

processes (superexchange mechanism) [61].<br />

5<br />

<strong>Coordination</strong> <strong>Compounds</strong> as Components<br />

<strong>of</strong> Photochemical Molecular Devices <strong>and</strong> Machines<br />

In the last few years, a combination <strong>of</strong> supramolecular chemistry <strong>and</strong> photochemistry<br />

has led to the design <strong>and</strong> construction <strong>of</strong> supramolecular systems<br />

capable <strong>of</strong> performing interesting light-induced functions. Photoinduced energy<br />

<strong>and</strong> electron transfer are indeed basic processes for connecting light<br />

energy inputs with a variety <strong>of</strong> optical, electrical, <strong>and</strong> mechanical functions,<br />

i.e., to obtain molecular-level devices <strong>and</strong> machines [48, 55]. We will now describe<br />

a few classical examples <strong>of</strong> molecular devices <strong>and</strong> machines in which<br />

coordination compounds are used to process light signals or to exploit light<br />

energy. Other examples are, <strong>of</strong> course, described in the chapters dealing with<br />

the complexes <strong>of</strong> the various metals.<br />

5.1<br />

A Molecular Wire<br />

An important function at the molecular level is photoinduced energy <strong>and</strong><br />

electron transfer over long distances <strong>and</strong>/or along predetermined directions.<br />

This function can be obtained by linking donor <strong>and</strong> acceptor components by<br />

a rigid spacer, as illustrated in Fig. 14.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 25<br />

Fig. 14 Photoinduced energy (a) <strong>and</strong> electron (b) transfer processes in a molecular wire<br />

based on coordination compounds [76]<br />

An example [76] is given by the [Ru(bpy)3] 2+ -(ph)n-[Os(bpy)3] 2+ compounds<br />

(bpy=2,2 ′ -bipyridine; ph = 1,4-phenylene; n =3,5,7),inwhichexcitation<br />

<strong>of</strong> the [Ru(bpy)3] 2+ unit is followed by electronic energy transfer<br />

to the ground state [Os(bpy)3] 2+ unit, as shown by the sensitized emission<br />

<strong>of</strong> the latter. For the compound with n = 7 (Fig. 14a), the rate constant for<br />

energy transfer over the 4.2-nm metal-to-metal distance is 1.3 × 10 6 s –1 .In<br />

the [Ru(bpy)3] 2+ -(ph)n-[Os(bpy)3] 3+ compounds, obtained by chemical oxidation<br />

<strong>of</strong> the Os-based moiety, photoexcitation <strong>of</strong> the [Ru(bpy)3] 2+ unit<br />

causesthetransfer<strong>of</strong>anelectrontotheOs-basedonewitharateconstant<br />

<strong>of</strong> 3.4 × 10 7 s –1 for n = 7 (Fig. 14b). Unless the electron added to the<br />

[Os(bpy)3] 3+ unit is rapidly removed, a back electron-transfer reaction (rate<br />

constant 2.7 × 10 5 s –1 for n = 7) takes place from the [Os(bpy)3] 2+ unit to the<br />

[Ru(bpy)3] 3+ one.<br />

Spacers with energy levels or redox states in between those <strong>of</strong> the donor<br />

<strong>and</strong> acceptor may help energy or electron transfer (hopping mechanism).<br />

Spacers whose energy or redox levels can be manipulated by an external<br />

stimulus can play the role <strong>of</strong> switches for the energy- or electron-transfer processes<br />

[48]. For a more thorough discussion <strong>of</strong> photoinduced energy- <strong>and</strong><br />

electron-transfer processes in systems involving metal complexes, see [61].


26 V. Balzani et al.<br />

5.2<br />

An Antenna System<br />

In suitably designed dendrimers, electronic energy transfer can be channeled<br />

toward a specific position <strong>of</strong> the array. <strong>Compounds</strong> <strong>of</strong> this kind play the role<br />

<strong>of</strong> antennas for light harvesting. We briefly illustrate an example involving<br />

luminescent lanthanide ions. For a more extended discussion <strong>of</strong> dendritic antenna<br />

systems, see [77].<br />

Lanthanide ions are known to show a very long-lived luminescence which<br />

is a potentially useful property. Because <strong>of</strong> the forbidden nature <strong>of</strong> their electronic<br />

transitions, however, lanthanide ions exhibit very weak absorption<br />

b<strong>and</strong>s, which is a severe drawback for applications based on luminescence.<br />

In order to overcome this difficulty, lanthanide ions are usually coordinated<br />

to lig<strong>and</strong>s containing organic chromophores whose excitation, followed by<br />

energy transfer, causes the sensitized luminescence <strong>of</strong> the metal ion (antenna<br />

effect). Such a process can involve either direct energy transfer from<br />

the singlet excited state <strong>of</strong> the chromophoric group with quenching <strong>of</strong> the<br />

chromophore fluorescence [78], or, most frequently, via S1 → T1 intersystem<br />

crossing followed by energy transfer from the T1 excited state <strong>of</strong> the chromophoric<br />

unit to the lanthanide ion [79, 80].<br />

Amide groups are known to be good lig<strong>and</strong>s for lanthanide ions. The<br />

dendrimer shown in Fig. 15 is based on a benzene core branched in the<br />

1, 3, <strong>and</strong> 5 positions, <strong>and</strong> it contains 18 amide groups in its branches <strong>and</strong><br />

24 chromophoric dansyl units in the periphery [81]. The dansyl units show<br />

strong absorption b<strong>and</strong>s in the near-UV spectral region <strong>and</strong> an intense fluorescence<br />

b<strong>and</strong> in the visible region. In acetonitrile/dichloromethane (5 : 1<br />

v/v) solutions, the absorption spectrum <strong>and</strong> the fluorescence properties <strong>of</strong><br />

the dendrimer are those expected for a species containing 24 noninteracting<br />

dansyl units. Upon addition <strong>of</strong> lanthanide ions to dendrimer solutions<br />

the following effects were observed [81]: (a) the fluorescence <strong>of</strong> the dansyl<br />

units is quenched; (b) the quenching effect is very large for Nd 3+ <strong>and</strong><br />

Eu 3+ ,moderateforYb 3+ ,smallforTb 3+ ,<strong>and</strong>verysmallforGd 3+ ;<strong>and</strong><br />

(c) in the case <strong>of</strong> Nd 3+ ,Er 3+ ,<strong>and</strong>Yb 3+ the quenching <strong>of</strong> the dansyl fluorescence<br />

is accompanied by the sensitized near-infrared emission <strong>of</strong> the<br />

lanthanide ion. Interpretation <strong>of</strong> the results obtained on the basis <strong>of</strong> the energy<br />

levels <strong>and</strong> redox potentials <strong>of</strong> the dendrimer <strong>and</strong> <strong>of</strong> the metal ions<br />

has led to the following conclusions: (1) at low metal ion concentrations,<br />

each dendrimer hosts only one metal ion; (2) when the hosted metal ion<br />

is Nd 3+ or Eu 3+ , all 24 dansyl units <strong>of</strong> the dendrimer are quenched with<br />

unitary efficiency; (3) quenching by Nd 3+ takes place by direct energy transfer<br />

from the fluorescent (S1) excitedstate<strong>of</strong>dansyltoamanifold<strong>of</strong>Nd 3+<br />

energy levels, followed by sensitized near-infrared emission from the metal<br />

ion (λmax = 1064 nm for Nd 3+ ); (4) quenching by Eu 3+ does not lead to<br />

any sensitized emission since a lig<strong>and</strong>-to-metal charge-transfer level lies be-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 27<br />

Fig. 15 Dendrimer based on a benzene core branched in the 1, 3, <strong>and</strong> 5 positions, which<br />

contains 18 amide groups in its branches <strong>and</strong> 24 chromophoric dansyl units in the periphery<br />

[81]<br />

low the luminescent Eu 3+ excited state; (5) in the case <strong>of</strong> Yb 3+ ,thesensitization<br />

<strong>of</strong> the near-infrared metal-centered emission occurs via the intermediate<br />

formation <strong>of</strong> an upper lying charge-transfer excited state; (6) the<br />

small quenching effect observed for Tb 3+ is partly caused by a direct energy<br />

transfer from the fluorescent (S1) excitedstate<strong>of</strong>dansyl;<strong>and</strong>(7)thevery<br />

small quenching effect observed for Gd 3+ is assigned to either induced intersystem<br />

crossing or, more likely, to charge perturbation <strong>of</strong> the S1 dansyl<br />

excited state.


28 V. Balzani et al.<br />

5.3<br />

An Extension Cable<br />

In the attempt <strong>of</strong> constructing a molecular-level extension cable, the [3]pseudorotaxane<br />

shown in Fig. 16, made <strong>of</strong> the three components A 2+ ,[BH] 3+ ,<strong>and</strong><br />

C, has been synthesized <strong>and</strong> studied [82]. Component A 2+ consists <strong>of</strong> two<br />

moieties: a [Ru(bpy)3] 2+ unit, which plays the role <strong>of</strong> electron donor under<br />

light excitation, <strong>and</strong> a crown ether, which plays the role <strong>of</strong> a first socket.<br />

The ammonium center <strong>of</strong> [BH] 3+ , driven by hydrogen-bonding interactions,<br />

threads as a plug into the first socket, whereas the bipyridinium unit, owing<br />

to charge-transfer (CT) interactions, threads as a plug into the third component,<br />

C, which plays the role <strong>of</strong> a second socket. In CH2Cl2/CH3CN (98 : 2 v/v)<br />

solution, reversible connection/disconnection <strong>of</strong> the two plug/socket functions<br />

can be controlled independently by acid–base <strong>and</strong> redox stimulation,<br />

respectively. In the fully connected triad, light excitation <strong>of</strong> the [Ru(bpy)3] 2+<br />

unit <strong>of</strong> component A 2+ is followed by electron transfer to the bipyridinium<br />

unit <strong>of</strong> component [BH] 3+ , which is plugged into component C. Although<br />

the transferred electron does not reach the final component <strong>of</strong> the assembly,<br />

the intercomponent connections employed fulfill an important requirement,<br />

namely, they can be controlled reversibly <strong>and</strong> independently. An improved<br />

example <strong>of</strong> a molecular extension cable based on [Ru(bpy)3] 2+ has been reported<br />

more recently [83].<br />

Fig. 16 Schematic representation <strong>of</strong> a supramolecular system that behaves as a molecularlevel<br />

extension cable [82]


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 29<br />

5.4<br />

An XOR Logic Gate with an Intrinsic Threshold Mechanism<br />

A system based on the photochemistry <strong>of</strong> a metal complex has been reported<br />

to mimic some elementary properties <strong>of</strong> neurons [84]. Such a system consists<br />

<strong>of</strong> an aqueous solution containing the trans-chalcone form (Ct, Fig. 17a)<br />

<strong>of</strong> the 4 ′ -methoxyflavylium ion (AH + ), <strong>and</strong> the [Co(CN)6] 3– complex ion (as<br />

a potassium salt). Excitation by 365-nm light <strong>of</strong> Ct, which is the thermodynamically<br />

stable form <strong>of</strong> the flavylium species in the pH range 3–7, causes<br />

a trans → cis photoisomerization reaction (Φ = 0.04). If the solution is sufficiently<br />

acid (pH


30 V. Balzani et al.<br />

in acid or neutral aqueous solution causes the dissociation <strong>of</strong> a CN – lig<strong>and</strong><br />

from the metal coordination sphere (Φ = 0.31), with a consequent increase in<br />

pH (Fig. 17b).<br />

When an acid solution (pH=3.6) containing 2.5 × 10 –5 mol L –1 Ct <strong>and</strong><br />

2.0 × 10 –2 mol L –1 [Co(CN)6] 3– is irradiated at 365 nm most <strong>of</strong> the incident<br />

light is absorbed by Ct, which undergoes photoisomerization to Cc.Sincethe<br />

pH <strong>of</strong> the solution is sufficiently acid, Cc is rapidly protonated (Fig. 17a), with<br />

the consequent appearance <strong>of</strong> the absorption b<strong>and</strong> with maximum at 434 nm<br />

<strong>and</strong> <strong>of</strong> the emission b<strong>and</strong> with maximum at 530 nm characteristic <strong>of</strong> the AH +<br />

species. On continuing irradiation, it can be observed that such absorption<br />

<strong>and</strong> emission b<strong>and</strong>s increase in intensity, reach a maximum value, <strong>and</strong> then<br />

decrease up to complete disappearance. In other words, AH + first forms <strong>and</strong><br />

then disappears with increasing irradiation time. The reason for the <strong>of</strong>f–on–<br />

<strong>of</strong>fbehavior<strong>of</strong>AH + under continuous light excitation is related to the effect <strong>of</strong><br />

the [Co(CN)6] 3– photoreaction (Fig. 17b) on the Ct photoreaction (Fig. 17a).<br />

As Ct is consumed with formation <strong>of</strong> AH + , an increasing fraction <strong>of</strong> the incident<br />

light is absorbed by [Co(CN)6] 3– , whose photoreaction causes an increase<br />

in the pH <strong>of</strong> the solution. This change in pH not only prevents further formation<br />

<strong>of</strong> AH + , which would imply protonation <strong>of</strong> the Cc molecules that continue<br />

to be formed by light excitation <strong>of</strong> Ct, but also causes the back reaction to Cc<br />

(<strong>and</strong>, then, to Ct) <strong>of</strong> the previously formed AH + molecules. Clearly, the examined<br />

solution performs like a threshold device as far as the input (light)/output<br />

(spectroscopic properties <strong>of</strong> AH + ) relationship is concerned.<br />

Instead <strong>of</strong> a continuous light source, pulsed (flash) irradiation can be<br />

used [83]. Under the input <strong>of</strong> only one flash, a strong change in absorbance at<br />

434 nm is observed, due to the formation <strong>of</strong> AH + . After two flashes, however,<br />

the change in absorbance practically disappears. In other words, an output<br />

(434-nm absorption) can be obtained only when either input 1 (flash 1) or input<br />

2 (flash 2) are used, whereas there is no output under the action <strong>of</strong> none<br />

or both inputs. This finding shows that the above-described system behaves<br />

according to XOR logic, under control <strong>of</strong> an intrinsic threshold mechanism<br />

(Fig. 17c).<br />

Two important features <strong>of</strong> the above system should be emphasized:<br />

(1) intermolecular communication takes place in the form <strong>of</strong> H + ions, <strong>and</strong><br />

(2) the input <strong>and</strong> output signals have the same nature (light) <strong>and</strong> the fluorescence<br />

output can be fed, in principle, into another device [85].<br />

5.5<br />

A Sunlight-Powered Nanomotor<br />

In the last few years, a great number <strong>of</strong> light-driven molecular machines have<br />

been developed <strong>and</strong> the field has been extensively reviewed [48, 55, 86–92]. In<br />

several cases, the working principle <strong>of</strong> such machines exploits photoinduced<br />

electron transfer by [Ru(bpy)3] 2+ or related coordination compounds.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 31<br />

Fig. 18 Chemical formula <strong>and</strong> cartoon representation <strong>of</strong> the rotaxane D 6+ ,showingits<br />

modular structure [94]<br />

In order to achieve photoinduced ring switching in rotaxanes containing<br />

two different recognition sites in the dumbbell-shaped component, the thoroughly<br />

designed rotaxane (D6+ ) shown in Fig. 18 was synthesized [93–95].<br />

This compound is made <strong>of</strong> the electron-donor macrocycle R,<strong>and</strong>adumbbellshaped<br />

component which contains (1) [Ru(bpy)3] 2+ (P)asone<strong>of</strong>itsstoppers,<br />

(2) a 4,4 ′ -bipyridinium unit (A1) <strong>and</strong>a3,3 ′ -dimethyl-4,4 ′ -bipyridinium unit<br />

(A2) aselectron-acceptingstations,(3)ap-terphenyl-type ring system as<br />

arigidspacer(S), <strong>and</strong> (4) a tetraarylmethane group as the second stopper<br />

(T). The stable translational isomer is the one in which the R component<br />

encircles the A1 unit, in keeping with the fact that this station is a better<br />

electron acceptor than the other one. The strategy devised in order to obtain<br />

the photoinduced abacus-like movement <strong>of</strong> the R macrocycle, between the<br />

two stations A1 <strong>and</strong> A2 illustrated in Fig. 19, is based on the following four<br />

operations:<br />

(a) Destabilization <strong>of</strong> the stable translational isomer: light excitation <strong>of</strong> the<br />

photoactive unit P (step 1) is followed by the transfer <strong>of</strong> an electron from<br />

the excited state to the A1 station, which is encircled by the ring R (step 2),<br />

with the consequent “deactivation” <strong>of</strong> this station; such a photoinduced<br />

electron-transfer process has to compete with the intrinsic excited-state<br />

decay (step 3).<br />

(b) Ring displacement: the ring moves by Brownian motion from the reduced<br />

station A – 1 to A2 (step 4), a step that has to compete with the back electron-<br />

transfer process from A – 1 (still encircled by R) to the oxidized photoactive<br />

unit P + (step 5).


32 V. Balzani et al.<br />

Fig. 19 Schematic representation <strong>of</strong> the operation <strong>of</strong> rotaxane D 6+ as a four-stroke linear<br />

nanomotor powered by sunlight [94]<br />

(c) Electronic reset: a back electron-transfer process from the “free” reduced<br />

station A – 1 to P+ (step 6) restores the electron-acceptor power to the A1<br />

station.<br />

(d) Nuclear reset: as a consequence <strong>of</strong> the electronic reset, back movement <strong>of</strong><br />

the ring from A2 to A1 takes place (step 7).<br />

The crucial point is the competion between ring displacement (step 4) <strong>and</strong><br />

back electron transfer (step 5). The results revealed that in acetonitrile solution<br />

at room temperature the ring shuttling rate is one order <strong>of</strong> magnitude<br />

slower than the back electron transfer. Hence, the absorption <strong>of</strong><br />

light can cause the occurrence <strong>of</strong> a forward <strong>and</strong> back ring movement (i.e.,<br />

afullcycle)witha2% quantum yield. The low efficiency is compensated<br />

by the following features: (1) the operation <strong>of</strong> the system relies exclusively<br />

on intramolecular processes, without generation <strong>of</strong> any waste product; <strong>and</strong><br />

(2) since [Ru(bpy)3] 2+ shows an intense absorption b<strong>and</strong> in the visible region,<br />

the system works simply upon exposure to sunlight. A much higher efficiency


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 33<br />

can be obtained by using an electron relay, again consuming only sunlight<br />

without generation <strong>of</strong> waste products [95].<br />

6<br />

Conclusions<br />

Research on the photochemistry <strong>and</strong> photophysics <strong>of</strong> coordination compounds<br />

has shown an extraordinary quantitative development as well as pr<strong>of</strong>ound<br />

qualitative changes over the years. Studies on intramolecular photoreactions<br />

<strong>and</strong> luminescence properties <strong>of</strong> coordination compounds <strong>of</strong> simple<br />

lig<strong>and</strong>s have been followed by investigations on compounds containing complex<br />

synthetic lig<strong>and</strong>s. Characterization <strong>of</strong> excited-state properties has been<br />

followed by extensive use <strong>of</strong> metal complexes in bimolecular processes. With<br />

the advent <strong>of</strong> supramolecular chemistry, luminescent <strong>and</strong>/or photoredox reactive<br />

metal complexes have been used as essential components in a bottomup<br />

approach to the construction <strong>of</strong> molecular devices <strong>and</strong> machines. In the<br />

next few years research on the photochemistry <strong>and</strong> photophysics <strong>of</strong> coordination<br />

compounds will largely be concentrated on the development <strong>of</strong> supramolecular<br />

systems for solar energy conversion <strong>and</strong> information processes. In this<br />

regard, it should be noted that the photoactive components presently used are<br />

very limited in number. Therefore, there is a need to extend basic research<br />

in order to discover novel mononuclear coordination compounds capable <strong>of</strong><br />

exhibiting long excited-state lifetimes, reversible redox behavior, <strong>and</strong> stability<br />

toward photodecomposition. The large number <strong>of</strong> metals that can be used<br />

<strong>and</strong> the endless number <strong>of</strong> lig<strong>and</strong>s that can be designed <strong>and</strong> synthesized open<br />

an ample horizon to these studies, as illustrated in the other chapters in this<br />

volume.<br />

Acknowledgements We acknowledge MIUR (PRIN projects no. 2006034123 & 2006030320),<br />

the University <strong>of</strong> Bologna <strong>and</strong> the University <strong>of</strong> Messina for financial support.<br />

References<br />

1. John DHD, Field GTJ (1963) A textbook <strong>of</strong> photographic chemistry. Chapman <strong>and</strong><br />

Hall, London<br />

2. Balzani V, Carassiti V (1970) <strong>Photochemistry</strong> <strong>of</strong> coordination compounds. Academic,<br />

London<br />

3. Ciamician G (1912) Science 36:385<br />

4. Leighton WG, Forbes GS (1930) J Am Chem Soc 52:3139<br />

5. Basolo F, Pearson RG (1958) Mechanisms <strong>of</strong> inorganic reactions. Wiley, New York<br />

6. Ballhausen CJ (1962) Introduction to lig<strong>and</strong> field theory. McGraw-Hill, New York<br />

7. Orgel LE (1954) Q Rev London 8:422<br />

8. Jorgensen CK (1962) Absorption spectra <strong>and</strong> chemical bonding in complexes. Pergamon,<br />

Oxford


34 V. Balzani et al.<br />

9. Schlaefer HL (1957) Z Phys Chem 11:65<br />

10. Plane RA, Hunt JP (1957) J Am Chem Soc 79:3843<br />

11. Adamson AW, Sporer AH (1958) J Am Chem Soc 80:3865<br />

12. Carassiti V, Claudi M (1958) Ann Chim (Rome) 49:1697<br />

13. Porter GB, Schlaefer HL (1964) Ber Bunsenges Phys Chem 68:316<br />

14. Endicott JF, H<strong>of</strong>fman MZ (1965) J Am Chem Soc 87:3348<br />

15. Balzani V, Carassiti V, Moggi L, Sabbatini N (1965) Inorg Chem 4:1247<br />

16. Calvert JG, Pitts JN Jr (1966) <strong>Photochemistry</strong>. Wiley, New York<br />

17. Turro NJ (1967) Molecular photochemistry. Benjamin, New York<br />

18. Wayne RP (1970) <strong>Photochemistry</strong>. Butterworths, London<br />

19. Simons JP (1971) <strong>Photochemistry</strong> <strong>and</strong> spectroscopy. Wiley-Interscience, London<br />

20. Balzani V, Sc<strong>and</strong>ola F (1991) Supramolecular photochemistry. Horwood, Chichester<br />

21. Schmidtke HH (1966) Inorg Chem 5:1683<br />

22. Nazeeruddin MK, Zakeeruddin SM, Kalyanasundaram K (1993) J Phys Chem 97:9607<br />

23. Gilbert A, Baggot J (1991) Essentials <strong>of</strong> molecular photochemistry. Blackwell Science,<br />

London<br />

24. Montalti M, Credi A, Prodi L, G<strong>and</strong>olfi MT (eds) (2006) H<strong>and</strong>book <strong>of</strong> photochemistry,<br />

3rdedn.CCR,Taylor<strong>and</strong>Francis,NewYork<br />

25. Balzani V, Bolletta F, G<strong>and</strong>olfi MT, Maestri M (1978) Top Curr Chem 75:1<br />

26. Adamson AW, Fleishauer PD (1975) (eds) Concepts <strong>of</strong> inorganic photochemistry.<br />

Wiley-Interscience, New York<br />

27. DeArmond MK (1974) Acc Chem Res 7:309<br />

28. Crosby GA (1975) Acc Chem Res 8:231<br />

29. Turro NJ (1978) Modern molecular photochemistry. Benjamin, Menlo Park<br />

30. Gafney HD, Adamson AW (1972) J Am Chem Soc 94:8238<br />

31. Balzani V, Moggi L, Manfrin MF, Bolletta F, Laurence GS (1975) Coord Chem Rev<br />

15:321<br />

32. Meyer TJ (1978) Acc Chem Res 11:94<br />

33. Whitten DG (1980) Acc Chem Res 13:83<br />

34. Sutin N, Creutz C (1980) Pure Appl Chem 52:2717<br />

35. Balzani V, Bolletta F, Ciano M, Maestri M (1983) J Chem Educ 60:447<br />

36. Velasco JG, Rubistein I, Crutchley RJ, Lever ABP, Bard AJ (1983) Inorg Chem 22:822<br />

37. Balzani V, Moggi L, Manfrin MF, Bolletta F, Gleria M (1975) Science 189:852<br />

38. Lehn JM, Sauvage JP (1977) Nouv J Chim 1:449<br />

39. Kalyanasundaram K, Graetzel M (1979) Angew Chem Int Ed Engl 18:701<br />

40. Gray HB, Maverick AW (1981) Science 214:1201<br />

41. Harriman A (1984) J Photochem 25:33<br />

42. Creutz C, Sutin N (1975) Proc Natl Acad Sci USA 72:2858<br />

43. Juris A, Balzani V, Barigelletti F, Campagna S, Belser P, Von Zelewsky A (1988) Coord<br />

Chem Rev 84:277<br />

44. H<strong>of</strong>fman MZ, Bolletta F, Moggi L, Hug GL (1989) J Phys Chem Ref Data 18:219<br />

45. Sutin N (1983) Prog Inorg Chem 30:441<br />

46. Lehn JM (1995) Supramolecular chemistry: concepts <strong>and</strong> perspectives. VCH, Weinheim<br />

47. de Rege PJF, Williams SA, Therien MJ (1995) Science 269:1409<br />

48. Balzani V, Venturi M, Credi A (2003) Molecular devices <strong>and</strong> machines. Wiley-VCH,<br />

Weinheim<br />

49. Schneider HJ, Yatsimirsky A (2000) Principles <strong>and</strong> methods in supramolecular chemistry.<br />

Wiley, Chichester<br />

50. Steed JW, Atwood JL (2000) Supramolecular chemistry. Wiley, Chichester


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong> 35<br />

51. Lehn JM (1993) In: Kisakürek MV (ed) Organic chemistry: its language <strong>and</strong> its state<br />

<strong>of</strong>theart.VCH,Weinheim,p77<br />

52. Kaifer AM, Gómez-Kaifer M (1999) Supramolecular electrochemistry. Wiley-VCH,<br />

Weinheim<br />

53. Balzani V, Sc<strong>and</strong>ola F (1996) In: Atwood JL, Davies JED, Macnicol DD, Vögtle F (eds)<br />

Comprehensive supramolecular chemistry, vol 10. Pergamon, Oxford, p 687<br />

54. Kunkely H, Vogler A (1990) J Am Chem Soc 112:5625<br />

55. Balzani V (2003) Photochem Photobiol Sci 2:459<br />

56. Balzani V (ed) (2001) Electron transfer in chemistry. Wiley-VCH, Weinheim<br />

57. Marcus RA, Sutin N (1985) Biochim Biophys Acta 811:265<br />

58. Newton MD (2001) In: Balzani V (ed) Electron transfer in chemistry, vol 1. Wiley-<br />

VCH, Weinheim, p 3<br />

59. Sc<strong>and</strong>ola F, Chiorboli C, Indelli MT, Rampi MA (2001) In: Balzani V (ed) Electron<br />

transfer in chemistry, vol 3. Wiley-VCH, Weinheim, p 337<br />

60. McConnell HM (1961) J Chem Phys 35:508<br />

61. Chiorboli C, Indelli MT, Sc<strong>and</strong>ola F (2005) Top Curr Chem 257:102<br />

62. Hush NS (1967) Prog Inorg Chem 8:391<br />

63. Nelsen SF (2001) In: Balzani V (ed) Electron transfer in chemistry, vol 1. Wiley-VCH,<br />

Weinheim, p 342<br />

64. Launay JP, Coudret C (2001) In: Balzani V (ed) Electron transfer in chemistry, vol 5.<br />

Wiley-VCH, Weinheim, p 3<br />

65. Brunschwig BS, Creutz C, Sutin N (2002) Chem Soc Rev 31:168<br />

66. Balzani V, Indelli MT, Maestri M, S<strong>and</strong>rini D, Sc<strong>and</strong>ola F (1980) J Phys Chem 84:852<br />

67. Balzani V, Bolletta F, Sc<strong>and</strong>ola F (1980) J Am Chem Soc 102:2152<br />

68. Orl<strong>and</strong>i G, Monti S, Barigelletti F, Balzani V (1980) Chem Phys 52:313<br />

69. Murtaza Z, Zipp AP, World LA, Graff D, Jones WE Jr, Bates WD, Meyer TJ (1991) J Am<br />

Chem Soc 113:5113<br />

70. Naqvi KR, Steel C (1993) Spectrosc Lett 26:1761<br />

71. Förster T (1959) Discuss Faraday Soc 27:7<br />

72. Pullerits T, Sundström V (1996) Acc Chem Res 29:381<br />

73. Dexter DL (1953) J Chem Phys 21:836<br />

74. Closs GL, Johnson DM, Miller JR, Piotrowiak P (1989) J Am Chem Soc 111:3751<br />

75. Sc<strong>and</strong>ola F, Balzani V (1983) J Chem Educ 60:814<br />

76. Schlicke B, Belser P, De Cola L, Sabbioni E, Balzani V (1999) J Am Chem Soc 121:4207<br />

77. Ceroni P, Bergamini G, Marchioni F, Balzani V (2005) Prog Polym Sci 30:453<br />

78. Hebbink GA, Klink SI, Grave L, Oude Alink PGB, van Veggel FCJM (2002) Chem Phys<br />

Chem 3:1014<br />

79. Parker D, Dickins RS, Puschmann H, Crossl<strong>and</strong> C, Howard JAK (2002) Chem Rev<br />

102:1977<br />

80. Sabbatini N, Guardigli M, Lehn J-M (1993) Coord Chem Rev 123:201<br />

81. Vicinelli V, Ceroni P, Maestri M, Balzani V, Gorka M, Vögtle F (2002) J Am Chem Soc<br />

124:6461<br />

82. Ballardini R, Balzani V, Clemente-León M, Credi A, G<strong>and</strong>olfi MT, Ishow E, Perkins J,<br />

Stoddart JF, Tseng HR, Wenger S (2002) J Am Chem Soc 124:12786<br />

83. Ferrer B, Rogez G, Credi A, Ballardini R, G<strong>and</strong>olfi MT, Balzani V (2006) Proc Natl<br />

Acad Sci USA 103:18411<br />

84. Pina F, Melo MJ, Maestri M, Passaniti P, Balzani V (2000) J Am Chem Soc 122:4496<br />

85. Balzani V, Venturi M, Credi A (2003) Logic gates. In: Molecular devices <strong>and</strong> machines.<br />

Wiley-VCH, Weinheim, p 235<br />

86. Balzani V, Credi A, Raymo FM, Stoddart JF (2000) Angew Chem Int Ed 39:3348


36 V. Balzani et al.<br />

87. Stoddart JF (ed) (2001) Special issue on molecular machines. Acc Chem Res 34:409<br />

88. Venturi M, Credi A, Balzani V (2001) Electron-transfer processes in pseudorotaxanes.<br />

In: Balzani V (ed) Electron transfer in chemistry, vol 3. Wiley-VCH, Weinheim, p 501<br />

89. Ballardini R, G<strong>and</strong>olfi MT, Balzani V (2001) Electron-transfer processes in rotaxanes<br />

<strong>and</strong> catenanes In: Balzani V (ed) Electron transfer in chemistry, vol 3. Wiley-VCH,<br />

Weinheim, p 539<br />

90. Ballardini R, Balzani V, Credi A, G<strong>and</strong>olfi MT, Venturi M (2001) Acc Chem Res<br />

34:445–455<br />

91. Sauvage JP (ed) (2001) Molecular machines <strong>and</strong> motors. Springer, Berlin<br />

92. Kelly TR (ed) (2005) Molecular machines. Top Curr Chem, vol 262. Springer, Berlin<br />

93. AshtonPR, BallardiniR, BalzaniV, Credi A, DressR, Ishow E, KleverlaanCJ, Kocian<br />

O, Preece JA, Spencer N, Stoddart JF, Venturi M, Wenger S (2000) Chem Eur J<br />

6:3558<br />

94. Balzani V, Credi A, Ferrer B, Silvi S, Venturi M (2005) In: Kelly TR (ed) Molecular<br />

machines. Top Curr Chem 262:1<br />

95. Balzani V, Clemente-León M, Credi A, Ferrer B, Venturi M, Flood AH, Stoddart JF<br />

(2006) Proc Natl Acad Sci USA 103:1178


Top Curr Chem (2007) 280: 37–67<br />

DOI 10.1007/128_2007_141<br />

© Springer-Verlag Berlin Heidelberg<br />

Published online: 11 July 2007<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong><br />

<strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium<br />

Noel A. P. Kane-Maguire<br />

Department <strong>of</strong> Chemistry, Furman University, 3300 Poinsett Highway,<br />

Greenville, SC 29613, USA<br />

noel.kane-maguire@furman.edu<br />

1 Introduction ................................... 38<br />

2 State Energy Levels <strong>and</strong> General Excited State Behavior ........... 38<br />

3 Ultrafast Dynamics <strong>of</strong> Cr(III) Lig<strong>and</strong> Field Excited States ......... 41<br />

3.1 Ultrafast Dynamics <strong>of</strong> 2 Eg State Formation in Cr(acac)3 ........... 42<br />

4 Photosubstitution Studies ........................... 43<br />

4.1 [Cr(phen)3] 3+ Photoracemization/Hydrolysis................. 43<br />

4.2 Axial Lig<strong>and</strong> Photodissociation in Cr(III) Porphyrins ............ 45<br />

5 Thermally Activated 2 Eg Excited State Relaxation Studies<br />

<strong>of</strong> Sterically Constrained Systems ....................... 47<br />

5.1 [Cr(sen)] 3+ <strong>and</strong> [Cr[18]aneN6] 3+ ....................... 49<br />

5.1.1 [Cr(sen)3] 3+ ................................... 49<br />

5.1.2 [Cr[18]aneN6] 3+ ................................. 50<br />

5.2 trans-[Cr(N4)(CN)2] +<br />

(where N4 = cyclam, 1,11-C3-cyclam, <strong>and</strong> 1,4-C2-cyclam).......... 51<br />

6 Energy Transfer Studies ............................ 52<br />

6.1 Self-Exchange Energy Transfer Between Identical Chromophores . . . . . . 53<br />

7 Photoredox Behavior <strong>of</strong> [Cr(diimine)3] 3+ Systems .............. 54<br />

7.1 DNAInteractions ................................ 56<br />

8 Photoredox Involving Coordinated Lig<strong>and</strong>s ................. 59<br />

8.1 Photolabilization <strong>of</strong> NO from Cr(III)-Coordinated Nitrite . . . ....... 59<br />

8.2 Photogeneration <strong>of</strong> Nitrido Complexes from Cr(III) Coordinated Azide . . 61<br />

9 Final Comments ................................. 62<br />

References ....................................... 63<br />

Abstract The study <strong>of</strong> the photochemistry <strong>and</strong> photophysics <strong>of</strong> octahedral <strong>and</strong> pseudooctahedral<br />

Cr(III) complexes has a rich history. An initial discussion is devoted to<br />

a general appraisal <strong>of</strong> the state <strong>of</strong> these two subjects up to December 1998, after providing<br />

a framework <strong>of</strong> state energy levels <strong>and</strong> radiative <strong>and</strong> non-radiative relaxation processes<br />

relevant to Cr(III) systems. The remaining sections cover some <strong>of</strong> the more active areas<br />

in the Cr(III) field, such as ultrafast dynamics, photosubstitution, thermally activated


38 N.A.P. Kane-Maguire<br />

excited state relaxation, energy transfer, <strong>and</strong> photoactivated redox processes (both intermolecular<br />

<strong>and</strong> intramolecular). Each <strong>of</strong> these sections begins with an overview <strong>of</strong> the<br />

subject area, <strong>and</strong> then one or more representative papers from the recent literature are<br />

selected for more detailed discussion.<br />

Keywords Energy transfer · Photoredox · Photosubstitution ·<br />

Thermal excited state relaxation · Ultrafast dynamics<br />

1<br />

Introduction<br />

Investigations <strong>of</strong> the photobehavior <strong>of</strong> octahedral (O h) or pseudo-octahedral<br />

chromium(III) complexes have played a pivotal role in the development <strong>of</strong><br />

transition metal photochemistry as a vital scientific discipline. Except for<br />

the case <strong>of</strong> ruthenium(II), the photochemistry <strong>and</strong> photophysics <strong>of</strong> Cr(III)<br />

systems have been explored more fully than those <strong>of</strong> any other transition<br />

metal ion. Their photoactivity has been extensively reviewed previously, <strong>and</strong><br />

readers are directed to the coverage in three texts [1–3], <strong>and</strong> the excellent discussion<br />

in the most recent comprehensive review <strong>of</strong> the topic by Kirk (which<br />

covered the literature up to December 1998) [4]. Several shorter Cr(III) reviews<br />

have since appeared, which have focused on a diverse range <strong>of</strong> more<br />

specific topics such as the excited state chemistry <strong>of</strong> pentacyanochromate(III)<br />

anions [5], intermediates in Cr(III) photochemistry [6], emission properties<br />

<strong>of</strong> hexam(m)ine Cr(III) systems [7], the interaction <strong>of</strong> [M(diimine)3] n+<br />

complexes <strong>of</strong> Ru(II) <strong>and</strong> Cr(III) with DNA [8], <strong>and</strong> thermal excited state relaxation<br />

[9].<br />

The Cr(III) field remains an active one, <strong>and</strong> the objective <strong>of</strong> the present<br />

chapter is to provide an overview <strong>of</strong> some <strong>of</strong> the interesting developments in<br />

the area from 1999 to December 2006.<br />

2<br />

State Energy Levels <strong>and</strong> General Excited State Behavior<br />

The electronic configuration <strong>of</strong> the Cr 3+ ion is [Ar]3d 3 . In its octahedral (Oh)<br />

complexes the degeneracy <strong>of</strong> the Cr d orbitals is lifted, resulting in two orbital<br />

subsets <strong>of</strong> t2g <strong>and</strong> eg symmetry. The 4 A2g (quartet) ground state has the<br />

electronic configuration (t2g) 3 ,withthed orbitals filled according to Hund’s<br />

Rule. Two excited states with (t2g) 2 (eg) 1 electronic configuration result from<br />

promotion <strong>of</strong> a t2g electron to an eg orbital while preserving electronic spin.<br />

Six such spin-allowed promotions are possible, which in O h symmetry are<br />

divided into two sets differing in the magnitude <strong>of</strong> the interelectronic repulsion<br />

terms. The associated quartet excited states generated are labeled 4 T2g


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 39<br />

<strong>and</strong> 4 T1g , respectively, with the former level being the more stable. For almost<br />

all octahedral Cr(III) complexes studied, the two lowest-lying excited states<br />

are the 2 T1g <strong>and</strong> 2 Eg (doublet) levels <strong>of</strong> (t2g) 3 electronic configuration, where<br />

spin-pairing has occurred within the t2g subshell. A qualitative orbital energy<br />

level diagram depicting the electron occupation <strong>of</strong> the pertinent electronic<br />

ground <strong>and</strong> excited states discussed above is provided in Fig. 1.<br />

A representative state energy level diagram for octahedral Cr(III) compounds<br />

was presented in Chap. 1 <strong>of</strong> this volume (Fig. 5b). Consistent with<br />

this diagram, for the preponderance <strong>of</strong> Cr(III) complexes the 2 Eg state has<br />

the lower energy <strong>of</strong> the two doublet excited states [1, 2]. However, exceptions<br />

occur occasionally for mixed lig<strong>and</strong> systems such as trans-[CrN4F2] + <strong>of</strong> D4h<br />

symmetry, where splitting <strong>of</strong> the 2 T1g (Oh)excitedstate(into 2 A2g <strong>and</strong> 2 Eg in<br />

D4h symmetry) results in its 2 Eg (D4h) component lying lower in energy than<br />

the components <strong>of</strong> the 2 Eg (O h) level [10, 11]. The generic Jablonski diagram<br />

shown in Chap. 1 <strong>of</strong> this volume (Fig. 6) has been modified in Fig. 2 <strong>of</strong> this<br />

chapter for the specific case <strong>of</strong> octahedral Cr(III) complexes.<br />

Fig. 1 The d orbital diagram for the electronic ground <strong>and</strong> excited states for d 3 Cr(III)<br />

complexes assuming Oh symmetry<br />

Fig. 2 Jablonski state energy level diagram for Oh Cr(III) complexes, showing the principal<br />

processes that may occur subsequent to vertical excitation to a 4 T2g or 4 T1g<br />

Franck–Condon excited state. Full arrows represent radiative processes (absorption or<br />

emission), whereas wavy arrows designate radiationless deactivation processes


40 N.A.P. Kane-Maguire<br />

The dominant lig<strong>and</strong> field b<strong>and</strong>s in the UV-visible absorption spectra are<br />

associated with the 4 A2g → 4 T2g <strong>and</strong> 4 A2g → 4 T1g transitions, since the corresponding<br />

absorptions generating the two doublet excited states are both<br />

Laporte <strong>and</strong> spin multiplicity forbidden. Photochemical <strong>and</strong> photophysical<br />

studies <strong>of</strong> Cr(III) species are, therefore, usually restricted to initial excitation<br />

into one <strong>of</strong> the quartet excited states. The spin-allowed absorption b<strong>and</strong>s for<br />

the 4 A2g → 4 T2g <strong>and</strong> 4 A2g → 4 T1g transitions are broad. This is a consequence<br />

<strong>of</strong> the 4 T2g <strong>and</strong> 4 T1g excited states both having an electron residing<br />

in an eg antibonding σ ∗ orbital (Fig. 1), which results in a large nuclear displacement<br />

relative to the ground state. For excitation into the higher lying<br />

4 T1g level, very fast internal conversion (IC) occurs to the 4 T2g state with<br />

near unit efficiency. Under normal photochemical conditions (i.e., in solution<br />

near room temperature) 4 T2g → 4 A2g fluorescence is rarely observed [12, 13],<br />

due to 4 T2g → 2 Eg intersystem crossing (ISC) being an unusually rapid process<br />

[4, 13–15] <strong>and</strong> <strong>of</strong>ten occurring with high efficiency (see Table 4 in [4]).<br />

With only occasional exceptions, Cr(III) complexes in rigid low temperature<br />

media exhibit intense, long-lived phosphorescence from the 2 Eg level generated<br />

by ISC. Very little geometric change is expected between the 4 A2g ground<br />

state <strong>and</strong> 2 Eg excited state, due their common (t2g) 3 orbital parentage. As<br />

a result, low temperature 2 Eg → 4 A2g phosphorescence spectra <strong>of</strong>ten display<br />

sharp, highly resolved fine structure, which has led to a very extensive<br />

literature (including medium <strong>and</strong> temperature effects) on the photophysical<br />

properties <strong>of</strong> these systems [2, 4, 9, 11, 16].<br />

In room temperature (rt) solution, 4 A2g → 4 T2g (or 4 A2g → 4 T1g )excitation<br />

<strong>of</strong>ten leads to facile substitution <strong>of</strong> one or more bound lig<strong>and</strong>s by solvent<br />

or an added nucleophile [1–4]. This observation does not, however, preclude<br />

the possibility <strong>of</strong> reaction out <strong>of</strong> the lower lying 2 Eg level, since this state is<br />

subsequently populated by rapid <strong>and</strong> efficient 4 T2g → 2 Eg ISC. An enormous<br />

effort has been expended over the last 30 years in an attempt to determine<br />

the relative photochemical roles <strong>of</strong> the 4 T2g <strong>and</strong> 2 Eg excited states. Importantly,<br />

a significant number <strong>of</strong> Cr(III) complexes exhibit relatively long-lived<br />

(≥ 100 ns) phosphorescence in solution near rt, <strong>and</strong> this emission can be bimolecularly<br />

quenched by added reagents. Comparing photoreaction data for<br />

experiments carried out in the presence <strong>and</strong> absence <strong>of</strong> these added reagents<br />

(an approach first pioneered by Chen <strong>and</strong> Porter [17]) has in many instances<br />

proved very informative.<br />

The most definitive quenching studies have been for the strong field hexacyano<br />

complex [Cr(CN)6] 3– [18] <strong>and</strong> the pentacyano species [Cr(CN)5(X)] n– ,<br />

where X = NH3 [19], pyridine (py) [20], <strong>and</strong> NCS – [5]. Under experimental<br />

conditions <strong>of</strong> total emission quenching, no reaction quenching was detected.<br />

Such data provide compelling evidence for the reaction proceeding exclusively<br />

out <strong>of</strong> the 4 T2g level (a similar conclusion was reached for [Cr(CN)6] 3–<br />

based on sensitization studies [21, 22]). For most Cr(III) systems, however,


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 41<br />

total phosphorescence quenching is accompanied by significant but less than<br />

total reaction quenching. Definitively establishing the precise role <strong>of</strong> the<br />

doublet level in this quenched reaction component has proven an elusive<br />

goal, with competing options including direct 2 Eg reaction, 2 Eg tunneling to<br />

a ground state intermediate (GSI) surface, or “delayed” quartet excited state<br />

reaction via thermally activated 2 Eg → 4 T2g back-intersystem crossing, BISC<br />

(Fig. 2).<br />

This debate has been exhaustively discussed elsewhere [4, 6, 23, 24], <strong>and</strong><br />

will not be a focus <strong>of</strong> this review. From the outset, it was appreciated that<br />

the 4 T2g level <strong>of</strong> (t2g) 2 (eg) 1 orbital parentage was an attractive c<strong>and</strong>idate for<br />

substitution chemistry, based on the occupation <strong>of</strong> an eg orbital which is σ ∗<br />

antibonding with respect to the metal–lig<strong>and</strong> (M – L) bond. At present, the<br />

most widely employed theoretical model for rationalizing Cr(III) photosubstitution<br />

behavior, assuming quartet reactivity, is the semi-empirical symmetry<br />

restricted angular overlap model (AOM) developed by Vanquickenborne<br />

<strong>and</strong> Ceulemans [25–27]. For mixed lig<strong>and</strong> systems it has had considerable<br />

success in predicting relative lig<strong>and</strong> labilities based on identifying the plane<br />

<strong>of</strong> labilization <strong>and</strong> assuming that the lig<strong>and</strong> with the smallest excited state<br />

M – L bond strength is preferentially substituted. A further strength <strong>of</strong> the<br />

model is the rationalization it provides for the stereochemical change that<br />

is a common feature <strong>of</strong> Cr(III) photochemistry (in contrast to their corresponding<br />

thermal behavior), especially for cases <strong>of</strong> axial lig<strong>and</strong> loss in mixed<br />

lig<strong>and</strong> systems <strong>of</strong> D4h or C4v symmetry [4, 28]. A more recent ab initio study<br />

<strong>of</strong> the photochemistry <strong>of</strong> Cr(III) ammine systems yielded results in good<br />

agreement with the earlier AOM calculations [29]. Tris-polypyridyl Cr(III)<br />

complexes may prove to be an exception to this apparent preference for quartet<br />

excited state reactivity. Early quenching studies on [Cr(phen)3] 3+ (where<br />

phen = 1,10-phenanthroline) revealed up to 95% reaction quenching in the<br />

presence <strong>of</strong> doublet quenchers such as I – <strong>and</strong> NCS – [30, 31], <strong>and</strong> the data from<br />

subsequent <strong>and</strong> more detailed investigations were most readily interpreted in<br />

terms <strong>of</strong> a direct doublet excited state reaction pathway [32–34].<br />

The remaining sections <strong>of</strong> this chapter are devoted to a discussion <strong>of</strong> developments<br />

since December 1998 in a range <strong>of</strong> different focus areas <strong>of</strong> Cr(III)<br />

photochemistry <strong>and</strong> photophysics. For convenience, the state term symbols<br />

for Oh symmetry shown in Fig. 2 are usually employed during these discussions.<br />

3<br />

Ultrafast Dynamics <strong>of</strong> Cr(III) Lig<strong>and</strong> Field Excited States<br />

Ultrafast time-resolved absorption spectroscopy constitutes one <strong>of</strong> the most<br />

exciting <strong>and</strong> promising new frontiers in transition metal photochemistry <strong>and</strong><br />

photophysics. The term ultrafast is applied to photoprocesses that occur on


42 N.A.P. Kane-Maguire<br />

the time scale <strong>of</strong> nuclear motion, <strong>and</strong> have until recently been most frequently<br />

associated with intramolecular processes such as vibrational equilibration<br />

<strong>and</strong> conformational dynamics <strong>and</strong> with medium effects such as solvation<br />

shell reorientation. The time frames involved are in the femtosecond to low<br />

picosecond range, <strong>and</strong> because <strong>of</strong> the challenging experimental dem<strong>and</strong>s, the<br />

field <strong>of</strong> ultrafast transition metal spectroscopy is still in its infancy [35–37].<br />

As depicted earlier (Fig. 2), for the Franck–Condon excited level generated<br />

by initial light absorption, evolution down to lower-energy excited states<br />

involves processes such as IC, ISC, <strong>and</strong> vibrational relaxation (VR). Conventional<br />

wisdom based on organic experience suggests that the rate constants<br />

for these three processes would normally be in the order: kVR >kIC ><br />

kISC [38]. However, the observation that Cr(III) 2 Eg → 4 A2g phosphorescence<br />

yields <strong>and</strong> emission/reaction quenching ratios showed some dependence on<br />

the wavelength chosen for 4 A2g → 4 T2g excitation led to the suggestion that<br />

4 T2g → 2 Eg ISC may compete successfully with 4 T2g vibrational cooling [39–<br />

43]. This possibility received support from picosecond pulse laser studies<br />

that showed that the rise time for 2 Eg excited state transient absorption was<br />

shorter than the instrument time response [12–14].<br />

3.1<br />

Ultrafast Dynamics <strong>of</strong> 2 Eg State Formation in Cr(acac)3<br />

The first Cr(III) femtosecond spectroscopic study addressing the questions<br />

raised above has been very recently reported in an elegant study by Juban<br />

<strong>and</strong> McCusker for the test case <strong>of</strong> [Cr(acac)3] (where acac = acetylaceto-<br />

Fig. 3 Simplified Jablonski state energy level diagram for [Cr(acac)3]


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 43<br />

Fig. 4 A one-electron representation <strong>of</strong> the orbital electron populations for the 4 LMCT<br />

<strong>and</strong> 2 LMCT excited states <strong>of</strong> [Cr(acac)3]<br />

nate) [15, 37]. The state <strong>and</strong> orbital energy level diagrams for this molecule<br />

are shown in Figs. 3 <strong>and</strong> 4, respectively.<br />

Both lig<strong>and</strong>-field 4 A2 → 4 T2 <strong>and</strong> charge transfer 4 A2 → 4 LMCT pulse excitation<br />

experiments were performed, the former being the first such study<br />

for any transition metal system. Time evolution for 2 E excited state formation<br />

in CH3CN solution at ambient temperature was followed by monitoring<br />

the transient signal associated with the 2 E → 2 LMCT transition.<br />

Data analysis yielded the rate constants shown in Fig. 4, <strong>and</strong> provides convincing<br />

evidence that 4 T2 → 2 E ISC competes effectively with vibrational<br />

relaxation in the initially formed 4 T2 state. Coupling these results with<br />

those from related studies on [Fe(tren(py)3)] 2+ (where tren is tris(2-pyridylmethyliminoethyl)amine)<br />

[44], the authors argue that for transition metal<br />

systems the relative nuclear equilibrium displacements <strong>of</strong> potential energy<br />

surfaces <strong>and</strong> the high density <strong>of</strong> states may have a larger influence on the<br />

time-course <strong>of</strong> Franck–Condon excited state relaxation than spin selection<br />

rules [15, 37].<br />

4<br />

Photosubstitution Studies<br />

A survey <strong>of</strong> the literature since 1999 reveals a marked decrease in activity in<br />

this foundation area <strong>of</strong> transition metal photochemistry, concomitant with<br />

the rapid development <strong>of</strong> the new areas <strong>of</strong> interest identified in Chap. 1 <strong>of</strong> this<br />

volume. For the case <strong>of</strong> octahedral or pseudo-octahedral Cr(III) species, only<br />

ten articles have been identified where lig<strong>and</strong> photosubstitution studies were<br />

the primary activity investigated [5, 6, 45–52]. Highlights from some <strong>of</strong> these<br />

papers are presented below.<br />

4.1<br />

[Cr(phen)3] 3+ Photoracemization/Hydrolysis<br />

The loss <strong>of</strong> optical activity <strong>of</strong> Λ-[Cr(phen)3] 3+ upon photolysis in aqueous<br />

solution exhibits a strong pH dependence [31, 32]. Under acidic conditions,


44 N.A.P. Kane-Maguire<br />

the rate <strong>of</strong> direct racemization is much larger than that <strong>of</strong> acid hydrolysis<br />

to a cis-[Cr(phen)2(H2O)2] 3+ product. In contrast, at pH ≥ 11, the rate<br />

<strong>of</strong> optical activity loss matches closely that for base hydrolysis to a cis-<br />

[Cr(phen)2(OH)2] + product,<strong>and</strong>itwassuggestedthattheloss<strong>of</strong>optical<br />

activity under basic conditions occurs primarily via the hydrolysis path. However,<br />

these data did not preclude the possibility that a substantial fraction <strong>of</strong><br />

rotation loss might occur via direct racemization, provided there is significant<br />

retention <strong>of</strong> optical configuration in the base hydrolysis reaction.<br />

A recent paper demonstrates that chiral capillary electrophoresis (CE)<br />

provides a very effective direct probe <strong>of</strong> the extent <strong>of</strong> racemization <strong>of</strong><br />

parent Λ-[Cr(phen)3] 3+ , while simultaneously determining the optical purity<br />

<strong>of</strong> the hydrolysis product [49]. The electropherogram shown in Fig. 5<br />

(electropherogram A) is for a mixture <strong>of</strong> rac-[Cr(phen)3] 3+ <strong>and</strong> cis-rac-<br />

[Cr(phen)2(H2O)2] 3+ (where 50 mM dibenzoyl-l-tartrate was employed as<br />

the capillary chiral additive), while the electropherogram for parent Λ-<br />

[Cr(phen)3] 3+ prior to photolysis is provided in Fig. 5 (electropherogram B).<br />

Fig. 5 Electropherograms <strong>of</strong> A an aqueous mixture <strong>of</strong> 1 mM rac-[Cr(phen)3] 3+ <strong>and</strong> 1 mM<br />

cis-rac-[Cr(phen)2(H2O)2] 3+ ,<strong>and</strong>B a 1 mM aqueous solution <strong>of</strong> Λ-[Cr(phen)3] 3+ ,using<br />

50 mM sodium dibenzoyl-l-tartrate as chiral additive in 25 mM borate buffer, pH 9.2<br />

(20 ◦ C). Detection wavelength = 320 nm<br />

The corresponding electropherograms for Λ-[Cr(phen)3] 3+ solutions irradiated<br />

for 24 min at 350 nm at pH 2.2 <strong>and</strong> pH 11.6, respectively, are presented<br />

in Fig. 6.<br />

The pH 2.2 data (Fig. 6) show significant formation <strong>of</strong> ∆-[Cr(phen)3] 3+ ,<br />

but no b<strong>and</strong>s (in agreement with expectation) for cis-[Cr(phen)2(H2O)2] 3+<br />

product. The corresponding pH 11.6 results (Fig. 6) are strikingly different.<br />

ThereisnoevidencefordirectΛ-[Cr(phen)3] 3+ racemization, while extensive<br />

hydrolysis is observed with no apparent retention <strong>of</strong> absolute configuration.<br />

In a control experiment, no loss <strong>of</strong> optical activity was observed on irradiating<br />

a ∆-cis-[Cr(phen)2(OH)2] + solution for 60 min at pH 11.6. These results<br />

provide direct verification that optical rotation loss for Λ-[Cr(phen)3] 3+


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 45<br />

Fig. 6 Electropherograms <strong>of</strong> 5 mM aqueous solutions <strong>of</strong> Λ-[Cr(phen)3] 3+ after 350 nm<br />

photolysis for 24 min at pH 2.2 <strong>and</strong> pH 11.6, using 50 mM sodium dibenzoyl-l-tartrate as<br />

chiral additive in 25 mM borate buffer, pH 9.2 (20 ◦ C). Detection wavelength = 320 nm<br />

under strongly basic conditions is predominantly a consequence <strong>of</strong> hydrolysis.<br />

4.2<br />

Axial Lig<strong>and</strong> Photodissociation in Cr(III) Porphyrins<br />

In a series <strong>of</strong> nanosecond laser photolysis studies, Inamo <strong>and</strong> coworkers<br />

have recently explored the details <strong>of</strong> axial lig<strong>and</strong> photosubstitution<br />

(<strong>and</strong> recombination) for a range <strong>of</strong> Cr(III) porphyrin systems <strong>of</strong> the type<br />

[Cr(porphyrin)(Cl)(L)] in toluene <strong>and</strong> dichloromethane solution [46, 47, 50,<br />

51]. Interest in this area derives from the possible biological implications <strong>and</strong><br />

the role <strong>of</strong> Cr(III) porphyrins in a variety <strong>of</strong> photocatalytic reactions.<br />

The presence <strong>of</strong> the highly conjugated porphyrin ring leads to orbital <strong>and</strong><br />

state energy level diagrams considerably different from those normally en-<br />

countered for Cr(III) complexes. Iterative extended Huckel calculations [52]<br />

predict that a vacant d 2<br />

z <strong>and</strong> half-filled dxy, dxz,<strong>and</strong>dyz orbitals <strong>of</strong> the Cr(III)<br />

center are located between the HOMO π <strong>and</strong> LUMO π∗ orbitals <strong>of</strong> the por-<br />

phyrin ring, while the empty Cr d 2<br />

x – 2<br />

y orbital lies well above the porphyrin<br />

LUMO π∗ orbital. Weak coupling <strong>of</strong> the porphyrin (π,π∗ )stateswiththe<br />

paramagnetic Cr(III) center results in the singlet ground <strong>and</strong> excited 1 (π,π∗ )<br />

states becoming 4S0 <strong>and</strong> 4S1 levels, respectively, whereas the excited triplet<br />

3 (π,π∗ )stateissplitintotripdoublet2T1, tripquartet4T1, <strong>and</strong>tripsextet6T1 levels. The resultant state energy level diagram is shown in Fig. 7 [52].<br />

In the studies by Inamo <strong>and</strong> coworkers, several porphyrin ring systems<br />

were utilized (tetraphenylporphyrin, octaethylporphyrin, <strong>and</strong> tetramesitylporphyrin)<br />

as well as a range <strong>of</strong> leaving axial lig<strong>and</strong>s (L = H2O, pyridine,<br />

piperidine, 1-methylimidazole, triphenylphosphine, <strong>and</strong> triphenylphosphine<br />

oxide). Analysis <strong>of</strong> the transient spectra observed, following initial pulse exci-


46 N.A.P. Kane-Maguire<br />

Fig. 7 State energy level diagram for [Cr(porphyrin)(Cl)(L)] complexes, showing the porphyrin<br />

(π,π ∗ ) levels following weak coupling with the d orbitals <strong>of</strong> the paramagnetic<br />

Cr(III) center<br />

tation <strong>of</strong> [Cr(porphyrin)(Cl)(L)] into the 4 S1 (π–π ∗ ) excited state, confirmed<br />

the formation <strong>of</strong> the five-coordinate complex, [Cr(porphyrin)(Cl)], produced<br />

by the photodissociation <strong>of</strong> the axial lig<strong>and</strong> L. Spectral evidence was also<br />

found for generation <strong>of</strong> the thermally equilibrated 4 T1 <strong>and</strong> 6 T1 excited states.<br />

The quantum yield, φ diss, for the photodissociation <strong>of</strong> L from [Cr(porphyrin)(Cl)(L)]<br />

0 was found to asymptotically decrease with increasing dissolved<br />

O2 concentrations towards a constant value. This suggested the presence<br />

<strong>of</strong> a quenchable dissociation pathway attributed to the longer-lived 4 T1<br />

<strong>and</strong> 6 T1 levels, <strong>and</strong> a non-quenchable reaction component associated with<br />

the short-lived (≪ 1 ns) 4 S1 level. The φ diss values were also found to vary<br />

markedly with the porphyrin ring <strong>and</strong> axial lig<strong>and</strong>s present. Figure 8 summarizes<br />

the electronic energy dissipation processes proposed for these Cr(III)<br />

porphyrin systems.<br />

Fig. 8 State energy level diagram for [Cr(porphyrin)(Cl)(L)] complexes showing the principal<br />

relaxation processes following 4 S0 → 4 S1 excitation. Full arrows represent radiative<br />

processes, whereas wavy arrows refer to radiationless decay pathways


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 47<br />

As shown in Fig. 8, the quenchable <strong>and</strong> non-quenchable dissociation pathways<br />

are both thought to proceed through lower-lying porphyrin π–Cr dπ<br />

charge transfer (CT) states [48]. For example, an increase in the electron<br />

density in a Cr dπ orbital with a z-axis component should lead to a weakening<br />

in the Cr-axial lig<strong>and</strong> bond. The rich photobehavior <strong>of</strong> these systems suggests<br />

that they would make good c<strong>and</strong>idates for future ultrafast spectroscopic<br />

studies.<br />

5<br />

Thermally Activated 2 Eg Excited State Relaxation Studies<br />

<strong>of</strong> Sterically Constrained Systems<br />

Early studies on Cr(III) complexes <strong>of</strong> the type [Cr(N4)X2] n+ where N4 is<br />

the macrocyclic lig<strong>and</strong> cyclam or tet a (see Fig. 9) <strong>and</strong> X = Cl – [53, 54],<br />

CN – [55, 56], NH3 [56–58], <strong>and</strong> F – [59] revealed a marked difference in the<br />

photobehavior <strong>of</strong> the geometric isomers. All the trans isomers are photoinert<br />

while the cis species are photoreactive. The cyano, ammine, <strong>and</strong> fluoro systems<br />

drew particular attention because <strong>of</strong> their strong emission in rt solution,<br />

with accompanying lifetimes almost identical to those reported at 77 K. These<br />

observations were attributed to the steric rigidity <strong>of</strong> the macrocyclic ring restricting<br />

access to the thermally activated photochemical relaxation channels<br />

available to their non-macrocylic analogs.<br />

Fig. 9 Macrocyclic-N4 lig<strong>and</strong>s, cyclam <strong>and</strong> tet a<br />

An extensive literature now exists on the effects <strong>of</strong> lig<strong>and</strong> steric constraint<br />

on 2 Eg excited state relaxation [4, 60–71], with studies by Endicott <strong>and</strong> coworkers<br />

being especially noteworthy. Hexaam(m)ine Cr(III) systems have been<br />

one key area <strong>of</strong> study <strong>of</strong> Endicott’s group, where a range <strong>of</strong> complexes were<br />

synthesized containing lig<strong>and</strong>s that would be trigonally strained if coordinated<br />

octahedrally to Cr(III). Their studies provided convincing evidence that<br />

the more trigonally strained lig<strong>and</strong> systems underwent more rapid 2 Eg deactivation.<br />

In an illustrative example [63], the photobehavior <strong>of</strong> [Cr(en)3] 3+ was


48 N.A.P. Kane-Maguire<br />

compared with that <strong>of</strong> the quasi-cage complex [Cr(sen)] 3+ ,where[Cr(sen)] 3+<br />

can be regarded as a [Cr(en)3] 3+ analog with a neopentyl cap bonded in a facial<br />

position.<br />

X-ray crystallographic data revealed that the CrN6 microsymmetry is virtually<br />

identical for these two complexes, with the NCrN bond angles in<br />

[Cr(sen)] 3+ being slightly closer to octahedral. These data <strong>and</strong> MM2 calculations<br />

also established considerable trigonal strain in the neopentyl cap for<br />

[Cr(sen)] 3+ .<br />

Both compounds were found to have similar 2 Eg → 4 A2g emission lifetimes<br />

at 77 K(120 µs <strong>and</strong>171 µs for the en <strong>and</strong> sen complexes, respectively),<br />

<strong>and</strong> fairly comparable quantum yields for photoaquation in rt solution (0.27<br />

<strong>and</strong> 0.10, respectively). However, the 2 Eg lifetime for [Cr(sen)] 3+ in ambient<br />

solution was a factor <strong>of</strong> 10 4 times shorter than that for [Cr(en)3] 3+ .The<br />

authors attributed this difference to a thermally activated 2 Eg deactivation<br />

channel promoted by steric factors associated with the sen complex. The general<br />

conclusion from this body <strong>of</strong> work was that large amplitude trigonal<br />

twists can facilitate thermally activated 2 Eg relaxationforarange<strong>of</strong>sterically<br />

constrained hexaam(m)ine Cr(III) complexes [64]. The authors also<br />

suggest that this relaxation pathway may have mechanistic implications for<br />

the photoracemization <strong>of</strong> Cr(III) species with D3 symmetry [63]. Such a reaction<br />

channel could, for example, facilitate the trigonal twist pathway invoked<br />

for the observed photoinversion <strong>of</strong> Λ-fac-[Cr(S-trp)3] to ∆-fac-[Cr(S-trp)3]<br />

(where S-trp is the bidentate amino acid lig<strong>and</strong> S-tryptophan) [72].<br />

Fig. 10 Quasi-cage N6 lig<strong>and</strong>, sen<br />

The earlier studies on macrocyclic cis- <strong>and</strong>trans-[Cr(N4)X2] n+ complexes<br />

(where X is NH3 or CN – )werealsoexp<strong>and</strong>edbyEndicott’sgrouptoinclude<br />

systems where stereochemical perturbations were introduced by the<br />

presence <strong>of</strong> methyl substituents in the macrocylic ring in positions near the<br />

Cr–X coordination sites [65]. Their analysis <strong>of</strong> X-ray data <strong>and</strong> MM2 calculations<br />

supported the hypothesis that the more facile thermally activated 2 Eg<br />

relaxation <strong>of</strong> the cis-[Cr(N4)X2] n+ systems is predominantly a stereochemi-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 49<br />

cal effect. It was also argued that 2 Eg back-intersystem crossing (BISC) was<br />

not likely to be an important component <strong>of</strong> 2 Eg excited state deactivation<br />

for Cr(III) complexes with N6 or N4C2 chromophores [64]. This latter conclusion<br />

has been questioned by Kirk [4, 68], <strong>and</strong> in Sect. 5.1 the subject is<br />

discussed further. In Sect. 5.2 some very recent work by the Wagenknecht<br />

group employing a new series <strong>of</strong> sterically constrained N4-macrocycles is featured<br />

[69–71].<br />

5.1<br />

[Cr(sen)] 3+ <strong>and</strong> [Cr[18]aneN6] 3+<br />

5.1.1<br />

[Cr(sen)3] 3+<br />

In a recent report, Kirk <strong>and</strong> coworkers have reinvestigated the photobehavior<br />

<strong>of</strong> [Cr(sen)] 3+ , <strong>and</strong> compared it with that for the macrocyclic complex,<br />

[Cr[18]aneN6] 3+ [68].<br />

The data obtained for [Cr(sen)] 3+ supported the earlier report <strong>of</strong> a very<br />

short doublet lifetime in rt aqueous solution, <strong>and</strong> the photoaquation quantum<br />

yield <strong>of</strong> 0.10 determined upon 4 A2g → 4 T2g excitation was in excellent<br />

agreement with that recorded earlier.<br />

However, based on a more detailed investigation <strong>of</strong> Cr(sen)] 3+ photoaquation,<br />

it was proposed that this process occurs via the 4 T2g excited state after<br />

back-intersystem crossing (BISC). The more convincing argument presented<br />

was that direct irradiation into the 2 Eg state yielded a photoaquation quantum<br />

yield 22% lower than that for 4 T2g excitation excitation. However, as<br />

noted by the authors, direct spin-forbidden doublet excitation experiments<br />

are fraught with difficulties. Their second argument was based on a deter-<br />

Fig. 11 Macrocyclic-N6 lig<strong>and</strong>, [18]aneN6


50 N.A.P. Kane-Maguire<br />

mination <strong>of</strong> the stereochemistry <strong>of</strong> the [Cr(sen-NH)(H2O)] 4+ photoaquation<br />

product via chiral capillary electrophoresis analysis (CE), employing<br />

d-tartrate as the chiral capillary additive. The primary aquation product<br />

exhibited only a single peak, consistent with the presence <strong>of</strong> trans product –<br />

a result anticipated by AOM theory assuming quartet excited state reactivity.<br />

However, a control thermal aquation experiment (where isomerism is not expected)<br />

also yielded the same data. Although an explanation was <strong>of</strong>fered for<br />

this latter result, the reviewer notes from experience [49, 73] that the CE separation<br />

<strong>of</strong> the ∆ <strong>and</strong> Λ isomers <strong>of</strong> cis hydrolysis products is <strong>of</strong>ten difficult,<br />

<strong>and</strong> a definitive assignment <strong>of</strong> a single peak to the trans isomer can be made<br />

confidently only after several chiral additives have been tested in the capillary<br />

buffer medium.<br />

5.1.2<br />

[Cr[18]aneN6] 3+<br />

No emission was detectable from this compound in rt solution, despite<br />

the presence <strong>of</strong> strong, long-lived 2 E → 4 A2g phosphorescence (162 µs) at<br />

77 K [67]. A temperature dependence study <strong>of</strong> the lifetime for this transition<br />

showed the usual low <strong>and</strong> high temperature regimes, with a single-exponential<br />

fit to the high temperature region giving an apparent activation energy <strong>of</strong><br />

34 kJ mol –1 . Interestingly, however, the compound was also photoinert in ambient<br />

solution. X-ray crystallographic data on [Cr[18]aneN6] 3+ indicated S6<br />

point group symmetry for the complex, with no evidence for trigonal twist<br />

strain in the [18]aneN6 lig<strong>and</strong>. The authors argue, therefore, that the Endicott<br />

thermally activated 2 Eg relaxation model is unlikely to be operative in<br />

this case. Instead, they propose that fast radiationless decay at rt is a consequence<br />

<strong>of</strong> the S6 distortion from octahedral geometry, which leads to a mixing<br />

<strong>of</strong> states with doublet <strong>and</strong> quartet character <strong>and</strong> a facilitation <strong>of</strong> 2 Eg ISC to<br />

the ground state. In the light <strong>of</strong> data to be presented in Sect. 5.2, one could<br />

also conjecture whether a contributing factor to the short rt lifetimes might be<br />

a non-productive reaction pathway involving transient Cr – Nbondcleavage.<br />

Finally, note is made <strong>of</strong> the recent communication by Sargeson <strong>and</strong> coworkers<br />

on the remarkable photobehavior <strong>of</strong> the caged hexamine complex,<br />

[Cr(fac-Me5-D3htricosaneN6] 3+ [74]. This photoinert compound exhibits<br />

unique photophysical behavior for an N6 chromophoric Cr(III) species<br />

in rt aqueous solution. In addition to displaying an exceptionally long 2 Eg<br />

state lifetime (τ = 235 µs), the emission shows a very strong isotope effect<br />

upon N – Hdeuteration(τ = 1.5 ms). These observations demonstrate that<br />

2 Eg excited state decay in solution at ambient temperature is dominated by<br />

2 Eg → 4 A2g radiationless deactivation, promoted by high frequency N – H<br />

stretching acceptor modes. Importantly, the results also argue against thermally<br />

activated back-intersystem crossing being a significant 2 Eg deactivation<br />

pathway for this CrN6 system.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 51<br />

5.2<br />

trans-[Cr(N4)(CN)2] + (where N4 = cyclam, 1,11-C3-cyclam, <strong>and</strong> 1,4-C2-cyclam)<br />

The Wagenknecht group has very recently studied a set <strong>of</strong> trans-dicyanochromium(III)<br />

complexes <strong>of</strong> topologically constrained tetraazamacrocycles,<br />

namely trans-[Cr(1,11-C3-cyclam)(CN)2] + <strong>and</strong> trans-[Cr(1,4-C2-cyclam)<br />

(CN)2] + (Fig. 12) to determine the effect that the additional strap has on<br />

the overall chemistry <strong>and</strong> photophysics relative to the cyclam parent complex<br />

[69–71].<br />

Fig. 12 Parent cyclam lig<strong>and</strong>, <strong>and</strong> the strapped derivatives 1,4-C2-cyclam <strong>and</strong> 1,11-C3cyclam<br />

In their initial work [69, 70], differences in thermal reactivity, UV-visible absorption<br />

spectra, <strong>and</strong> low temperature photophysics were adequately explained<br />

on the basis <strong>of</strong> steric <strong>and</strong> symmetry arguments, <strong>and</strong> differences in numbers<br />

<strong>of</strong> N – H oscillators in the molecule. However, marked differences in their rt<br />

photobehavior eluded explanation. For example, the 1,11-C3-cyclam <strong>and</strong> 1,4-<br />

C2-cyclam complexes have rt 2 Eg excited state lifetimes one <strong>and</strong> three orders <strong>of</strong><br />

magnitude lower, respectively, than the corresponding cyclam complex.<br />

Furthermore, the lifetimes for complexes with the topologically constrained<br />

lig<strong>and</strong>s are strongly temperature dependent near rt in acidified aqueous solution<br />

<strong>and</strong> the Arrhenius plots are linear [70]. Potential radiationless deactivation<br />

pathways for the 2 Eg level in these systems are depicted in Fig. 13.<br />

Of these possibilities, back-intersystem crossing (BISC) was considered<br />

unlikely on energetic grounds, while net photoreaction was rejected as a significant<br />

relaxation pathway due to the very low quantum yields for photoaquation<br />

for all three complexes. Additionally, MM2 studies suggested that<br />

neither solvent association nor symmetry destroying molecular “twists” are<br />

likely causes for the data in the temperature-dependent regime [70]. In their<br />

most recent paper [71], the authors present evidence that a photodissociation<br />

pathway involving transient Cr-macrocyclic N-bond cleavage (followed<br />

by rapid ring closure) was the most plausible explanation for the thermally<br />

activated 2 Eg relaxation. This conclusion received strong support from the<br />

observation <strong>of</strong> photodeuteration <strong>of</strong> the NH protons upon photolysis <strong>of</strong> the<br />

cyclam complex in acidified D2O (where thermal deuteration was shown to


52 N.A.P. Kane-Maguire<br />

Fig. 13 Qualitative potential energy level diagram for Oh Cr(III) complexes showing possible<br />

radiationless 2 Eg relaxation processes, including a direct deactivation to the ground<br />

state, b back-intersystem crossing (BISC), c direct doublet reaction, or d surface crossing<br />

to a ground state intermediate surface (GSI)<br />

be minimal). The proposed mechanism for this photodeuteration is shown in<br />

Fig. 14.<br />

Fig. 14 Possible mechanism for photoinitiated macrocyclic N – H deuteration in acidic<br />

aqueous solution<br />

A related paper on the corresponding difluorosystems has recently been<br />

published [75].<br />

6<br />

Energy Transfer Studies<br />

Cr(III) complexes were employed as acceptor species in room temperature<br />

energy transfer experiments between transition metal complexes as early as<br />

1972 [21, 76], <strong>and</strong> several years later the first cases appeared where the donor<br />

<strong>and</strong> acceptor were both Cr(III) compounds [77, 78]. In the survey since 1999,<br />

eight articles were identified where energy transfer studies involving Cr(III)<br />

species were the primary research activity [79–86]. The paper chosen for<br />

discussion in Sect. 6.1 describes the first report <strong>of</strong> electronic energy selfexchange<br />

between Cr(III) complexes [81].


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 53<br />

6.1<br />

Self-Exchange Energy Transfer Between Identical Chromophores<br />

As noted earlier (Chap. 1 <strong>of</strong> this volume, Sect. 4.4.2), electronic energy transfer<br />

involving Cr(III) complexes is expected to proceed via an exchange<br />

mechanism, <strong>and</strong> thus effective donor–acceptor orbital overlap is a necessity.<br />

A large number <strong>of</strong> cross-exchange energy transfer studies utilizing Cr(III)<br />

donors <strong>and</strong>/or acceptors have been undertaken with the objective <strong>of</strong> determining<br />

the relative importance <strong>of</strong> thermodynamics, electronic factors (such<br />

as orbital overlap), <strong>and</strong> nuclear factors associated with Franck–Condon restrictions<br />

[87–92].<br />

The study <strong>of</strong> self-exchange energy transfer is an attractive complementary<br />

approach, since the effect <strong>of</strong> thermodynamics on the rate is eliminated.<br />

However, monitoring self-exchange has proven a serious experimental challenge,<br />

since the absorption <strong>and</strong> emission characteristics <strong>of</strong> the donor <strong>and</strong><br />

acceptor are identical. The only prior report <strong>of</strong> virtual self-exchange involving<br />

transition metal systems is that <strong>of</strong> Balzani <strong>and</strong> coworkers on several Ru(II)<br />

polypyridyl systems [93, 94]. In the paper to be discussed [81], advantage<br />

was taken <strong>of</strong> the marked enhancements in emission lifetimes <strong>and</strong> steadystate<br />

intensities in rt solution for the Cr(III) complexes listed in Table 1 upon<br />

deuteration <strong>of</strong> the amine N – Hprotons.<br />

For each complex, the solution absorption <strong>and</strong> emission maxima <strong>of</strong> the<br />

deuterated <strong>and</strong> undeuterated compounds were essentially identical, indicating<br />

the presence <strong>of</strong> effectively identical chromophores. Irradiation <strong>of</strong> acidified<br />

mixtures <strong>of</strong> the isotopically labeled <strong>and</strong> unlabeled chromophores leads to the<br />

Table 1 Emission lifetimes <strong>of</strong> Cr(III) complexes at 20 ◦ C<br />

Complex Solvent τH τD Refs.<br />

(µs) (µs)<br />

trans-[Cr(cyclam)(CN)2] + H2O 335 1500 [55]<br />

trans-[Cr(cyclam)(NH3)2] 3+ DMSO 135 1620 [57]<br />

trans-[Cr(tet a)F2] + H2O 30 234 [59]<br />

Scheme 1 Energy transfer between long-lived (CrL) <strong>and</strong>short-lived(CrS) complexes


54 N.A.P. Kane-Maguire<br />

reaction sequence shown in Scheme 1. In this scheme, the long-lived <strong>and</strong><br />

short-lived Cr(III) species are labeled CrL <strong>and</strong> CrS, respectively, while kET<br />

<strong>and</strong> k–ET are the corresponding rate constants for forward <strong>and</strong> reverse energy<br />

transfer. Likewise, the terms kL <strong>and</strong> kS are the reciprocals <strong>of</strong> the lifetimes <strong>of</strong><br />

CrL <strong>and</strong> CrS, respectively, in the absence <strong>of</strong> energy transfer. Emission quenching<br />

<strong>of</strong> CrL <strong>and</strong> CrS couldthenbefollowedbyanalyzingthedecaypr<strong>of</strong>ile<br />

following pulsed excitation according to the mathematical formulation developed<br />

by Maharaj <strong>and</strong> Winnik [95].<br />

For the trans-dicyano <strong>and</strong> trans-diammine systems, energy transfer rate<br />

constants at 20 ◦ C(µ = 1.0) were determined to be kET ≫ 7 × 10 6 M –1 s –1<br />

<strong>and</strong> 9.7 × 10 6 M –1 s –1 , respectively. However, for trans-[Cr(tet a)F2] + no<br />

energy transfer was observed, which implied that the rate constant was<br />

≪ 3 × 10 5 M –1 s –1 . Analysis <strong>of</strong> these results using Marcus theory lead to<br />

the important conclusion that electronic effects play a significant role in<br />

determining the rates <strong>of</strong> energy transfer self-exchange for these series <strong>of</strong><br />

complexes.<br />

7<br />

Photoredox Behavior <strong>of</strong> [Cr(diimine)3] 3+ Systems<br />

Arguably, the seminal report by Gafney <strong>and</strong> Adamson in 1972 [96] that the<br />

3 MLCT excited state <strong>of</strong> [Ru(bpy)3] 2+ (where bpy is 2,2 ′ -bipyridine) could<br />

function as an electron transfer agent was the catalytic event that led to the<br />

extraordinary growth <strong>of</strong> transition metal photochemistry <strong>and</strong> photophysics<br />

over the last three decades [94, 97–99]. Today, the polypyridyl compounds<br />

<strong>of</strong> Ru(II) still hold a favored status, due to a coalescence <strong>of</strong> desirable properties<br />

including an intense, relatively long-lived luminescence signature, <strong>and</strong><br />

a remarkable thermal robustness in a range <strong>of</strong> oxidation states.<br />

The analogous polypyridyl complexes <strong>of</strong> Cr(III) are the next most investigated<br />

[M(diimine)3] n+ systems. A few years after the Gafney <strong>and</strong> Adamson<br />

article appeared, Bolletta et al. presented the first evidence that the lig<strong>and</strong>field<br />

2 Eg excited state <strong>of</strong> [Cr(bpy)3] 3+ was a strong one-electron photooxidant<br />

[100]. This involvement <strong>of</strong> polypyridyl Cr(III) species in direct bimolecular<br />

electron transfer reactions <strong>of</strong> the generic type represented in Eq. 1 has<br />

since been thoroughly documented for numerous substrates, Q [101–106]:<br />

( 2 Eg)Cr 3+ +Q→ Cr 2+ +Q + . (1)<br />

Importantly, [Cr(diimine)3] 3+ complexes are more powerful photooxidants<br />

than their [Ru(diimine)3] 2+ analogs. The oxidizing power <strong>of</strong> the Cr(III) 2 Eg<br />

excited state can be assessed from the value <strong>of</strong> the 2 Eg excited state reduction<br />

potential, E o ( ∗ Cr 3+ /Cr 2+ ). It has been shown [102, 103] that this latter quantity<br />

can be reliably estimated from the difference between the 2 Eg → 4 A2g


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 55<br />

emission energies in eV units <strong>and</strong> the ground state st<strong>and</strong>ard reduction potentials,<br />

E o (Cr 3+ /Cr 2+ ), obtained from cyclic voltammetry (CV) measurements.<br />

A representative illustration <strong>of</strong> the relevant energetics is shown in Fig. 15<br />

for the case <strong>of</strong> [Cr(bpy)3] 3+ ,fromwhichE o ( ∗ Cr 3+ /Cr 2+ ) is determined to be<br />

1.44 V versus NHE in aqueous solution [103].<br />

Fig. 15 Energetics associated with 2 Eg excited state oxidizing power<br />

In contrast, the corresponding E o ( ∗ Ru 2+ /Ru + ) value for [Ru(bpy)3] 2+<br />

(where ∗ Ru 2+ is the 3 MLCT excited state) is reported as 0.84 V [107]. The<br />

primary reason for these differences is the relatively minor energetic cost<br />

<strong>of</strong> ground state M n+ → M (n–1)+ reduction in the Cr(III) case, which leaves<br />

approximately 85% <strong>of</strong> the free energy <strong>of</strong> the 2 Eg excited state available for<br />

photoredox (as opposed to 40% for the Ru(II) analog).<br />

Another important observation is that the 2 Eg → 4 A2g emission signal<br />

<strong>of</strong> [Cr(diimine)3] 3+ complexes in ambient solution is significantly quenched<br />

by the presence <strong>of</strong> dissolved oxygen, 3 O2, as the result <strong>of</strong> an energy transfer<br />

process generating excited state singlet oxygen ( 1 O2) [103, 105, 108]:<br />

( 2 Eg)Cr 3+ + 3 O2 → ( 4 A2g)Cr 3+ + 1 O2 . (2)<br />

SingletoxygenproductioninEq.2thenprovidesanalternativemethodfor<br />

substrate oxidation, where the Cr(III) 2 Eg excited state is functioning as<br />

a photocatalyst. During the present review period, Pagliero <strong>and</strong> Argüello<br />

examined the role <strong>of</strong> O2 in the photooxidation <strong>of</strong> phenols in aqueous solution,<br />

employing [Cr(phen)3] 3+ as the photocatalyst [109]. Although direct<br />

phenol oxidation according to Eq. 1 is thermodynamically feasible, under airsaturated<br />

conditions the net photochemistry is dominated by a singlet oxygen<br />

mediated pathway leading to benzoquinone as the sole organic product. The<br />

results confirm <strong>and</strong> amplify the observations from an earlier study [110], <strong>and</strong><br />

have practical relevance to the emerging field <strong>of</strong> photoremediation <strong>of</strong> waste<br />

waters [111].<br />

Despite the large number <strong>of</strong> molecules that have been shown to quench<br />

the 2 Eg excited state <strong>of</strong> [Cr(diimine)3] 3+ complexes via Eq. 1, biological substrates<br />

have very rarely been employed in this role. Several recent studies


56 N.A.P. Kane-Maguire<br />

utilizing DNA as the potential quenching species are highlighted in the following<br />

section.<br />

7.1<br />

DNA Interactions<br />

The last 20 years have witnessed the emergence <strong>of</strong> a rich chemistry associated<br />

with the non-covalent interaction <strong>of</strong> chiral [M(diimine)3] n+ complexes<br />

with duplex DNA, with the ultimate goal <strong>of</strong> developing new diagnostic <strong>and</strong><br />

therapeutic agents [112, 113]. Of particular interest in the present context is<br />

the potential utility <strong>of</strong> [M(diimine)3] n+ systems as DNA photocleavage agents<br />

via excited state redox processes, which could lead to applications in the general<br />

field <strong>of</strong> photodynamic therapy [111]. The most widely explored systems<br />

have been [Ru(diimine)3] 2+ species. However, except for a few notable exceptions<br />

[114], values for E 0 ( ∗ Ru 2+ /Ru + )fallwellshort<strong>of</strong>the1.2 Vvalue<br />

required for direct one-electron oxidation <strong>of</strong> guanine (the most readily oxidized<br />

nucleobase [115]) via a reaction pathway analogous to Eq. 1 above.<br />

Although [Ru(diimine)3] 2+ systems are known to photoinitiate DNA oxidation,<br />

this damage normally occurs via the intermediacy <strong>of</strong> a singlet oxygen<br />

pathway analogous to Eq. 2 [116, 117].<br />

In view <strong>of</strong> the markedly higher oxidative power <strong>of</strong> the 2 Eg excited state<br />

<strong>of</strong> Cr(III) polypyridyl complexes (approximately 1.4 V versus NHE), such<br />

species would appear to be more attractive c<strong>and</strong>idates for photoinitiated direct<br />

oxidation <strong>of</strong> DNA via Eq. 1 (where Q = DNA). Another potential advantage<br />

<strong>of</strong> these d 3 systems with regard to bimolecular redox activity is their longerlived<br />

2 Eg → 4 A2g emission (<strong>of</strong>ten two orders <strong>of</strong> magnitude greater than that<br />

for the analogous Ru(II) 3 MLCT emission signals [106]). These photoredox<br />

expectations have been experimentally confirmed in several reports in the<br />

present review period [118–120].<br />

In the first <strong>of</strong> these studies, the interaction <strong>of</strong> the complexes [Cr(phen)3] 3+<br />

<strong>and</strong> [Cr(bpy)3] 3+ with duplex DNA <strong>and</strong> a range <strong>of</strong> mononucleotides was explored<br />

[118]. A key observation was that the Cr(III) emission signals were<br />

strongly quenched in the presence <strong>of</strong> guanine-containing nucleotides, but not<br />

by the mononucleotides <strong>of</strong> adenine, cytosine, or thymidine, nor by the synthetic<br />

polynucleotide, poly(dA-dT) · poly(dA-dT). A representative example <strong>of</strong><br />

Cr(III) emission quenching in shown in Fig. 16 for the case <strong>of</strong> [Cr(phen)3] 3+ in<br />

the presence <strong>of</strong> calf thymus B-DNA (which has 40%GCbasepairs).<br />

Such behavior provides strong evidence for direct oxidation <strong>of</strong> the guanine<br />

base <strong>of</strong> DNA via Eq. 1, since the corresponding oxidation <strong>of</strong> the other<br />

nucleobases is thermodynamically more difficult [115]. Since guanine oxidation<br />

has been shown in other systems to serve as a genesis point for DNA<br />

str<strong>and</strong> scission [115], these Cr(III) complexes show potential as a new class<br />

<strong>of</strong> DNA photocleavage agents (photonucleases). This expectation receives<br />

support from our recent observation <strong>of</strong> permanent DNA damage (str<strong>and</strong>


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 57<br />

Fig. 16 Quenching <strong>of</strong> [Cr(phen)3] 3+ steady-state emission in air-saturated 50 mM Tris-<br />

HCl buffer (pH 7.4) by calf thymus B-DNA (22 ◦ C)<br />

scission) from agarose gel electrophoresis studies <strong>of</strong> [Cr(phen)3] 3+ bound<br />

to supercoiled ΦX174 RF plasmid DNA following photolysis at 350 nm (see<br />

Fig. 17) (Barnett et al., unpublished observations). In such studies, the detection<br />

<strong>of</strong> open-circular DNA following photolysis is clear evidence for a single<br />

str<strong>and</strong> break in the DNA [116].<br />

Fig. 17 Photoactivated cleavage at pH 7.4 <strong>of</strong> 5 µM ΦX174 plasmid DNA by 200 µM<br />

[Cr(phen)3] 3+ following irradiation at 350 nm (Rayonet). Samples were subjected to electrophoresis<br />

on 1% agarose gels for 3 hat 70 V, followed by staining with ethidium<br />

bromide<br />

It is also noteworthy that while photodamage in the analogous Ru(II) cases<br />

is dramatically decreased in the absence <strong>of</strong> O2 (consistent with a singlet O2<br />

pathway), photodamage by [Cr(phen)3] 3+ is considerably greater under a N2<br />

atmosphere (Barnett et al., unpublished observations). Since many cancer


58 N.A.P. Kane-Maguire<br />

cells are hypoxic [111], the increase in photodamage at lower O2 levels may<br />

provide a selectivity advantage for these [Cr(diimine)3] 3+ reagents in terms<br />

<strong>of</strong> their future potential as phototherapeutic agents.<br />

In another aspect <strong>of</strong> this study [118], a mathematical analysis <strong>of</strong> the emission<br />

quenching data was undertaken. Representative steady-state intensity<br />

<strong>and</strong> lifetime Stern–Volmer (SV) plots for quenching <strong>of</strong> [Cr(phen)3] 3+ emissionbycalfthymusB-DNAareshowninFig.18.Fromthelifetimedata,<br />

a bimolecular quenching rate constant <strong>of</strong> 1.1 × 10 8 M –1 s –1 was extracted<br />

(a value close to that anticipated for a diffusion controlled process). In contrast,<br />

the steady-state SV plot showed strong upward curvature at higher DNA<br />

concentrations. This observation was attributed to the formation <strong>of</strong> a nonluminescent<br />

[Cr(phen)3] 3+ /DNA ion pair, <strong>and</strong> allowed an estimation to be<br />

made for the binding constant with DNA (KDNA ≈ 4000 M –1 ).<br />

Fig. 18 Stern–Volmer plots for [Cr(phen)3] 3+ emission quenching in air-saturated 50 mM<br />

Tris-HCl buffer (pH 7.4) by calf thymus B-DNA at 22 ◦ C: • steady-state data, � lifetime<br />

data<br />

A limitation <strong>of</strong> this initial work with [Cr(bpy)3] 3+ <strong>and</strong> [Cr(phen)3] 3+<br />

is the relatively small binding constants <strong>of</strong> these compounds with DNA.<br />

In a subsequent study [119], the photoredox behavior <strong>of</strong> the complex<br />

[Cr(phen)2(DPPZ)] 3+ with DNA was investigated, where the third diimine<br />

lig<strong>and</strong> is dipyridophenazine, DPPZ. The value <strong>of</strong> KDNA increased by two<br />

orders <strong>of</strong> magnitude, consistent with the known ability <strong>of</strong> the DPPZ lig<strong>and</strong><br />

to intercalate into DNA base stacks [113]. Perhaps more importantly, the<br />

complex was found to have an E o ( ∗ Cr 3+ /Cr 2+ )value80 mV more positive<br />

than that for [Cr(phen)3] 3+ , which placed it in the thermodynamic threshold<br />

range required for direct oxidation <strong>of</strong> the nucleobase adenine [115]. In accord<br />

with this thermodynamic argument, SV plots <strong>of</strong> the quenching <strong>of</strong> the emission<br />

lifetime <strong>of</strong> [Cr(phen)2(DPPZ)] 3+ in the presence <strong>of</strong> deoxyguanosine-5 ′ -<br />

monophosphate <strong>and</strong> deoxyadenosine-5 ′ -monophosphate yielded quenching<br />

rate constants <strong>of</strong> 2.4 × 10 9 M –1 s –1 <strong>and</strong> 1.8 × 10 7 M –1 s –1 , respectively [119].<br />

More recently [120], a report by Vaidyanathan <strong>and</strong> Nair describes nucleobase<br />

photooxidation by the terpyridine Cr(III) derivatives [Cr(ttpy)2] 3+


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 59<br />

<strong>and</strong> [Cr(Brphtpy)2] 3+ (where ttpy = p-tolylterpyridine <strong>and</strong> Brphtpy = pbromophenylterpyridine).<br />

The two complexes were reported to emit strongly<br />

in rt aqueous solution, although no emission lifetimes or spectra (except<br />

wavelength maxima) were provided. Such emission is quite remarkable in<br />

view <strong>of</strong> the exceedingly weak emission <strong>and</strong> very short lifetime (≈ 0.05 µs)<br />

found for the parent terpyridine complex, [Cr(tpy)2] 3+ [103]. Based on CV<br />

data <strong>and</strong> the reported emission spectral maxima, exceptionally high values<br />

for E o ( ∗ Cr 3+ /Cr 2+ ) were assessed for [Cr(ttpy)2] 3+ <strong>and</strong> [Cr(Brphtpy)2] 3+<br />

(1.65 V<strong>and</strong>1.85 V, respectively). Consistent with these thermodynamic observations,<br />

both complexes were demonstrated to be very powerful photooxidants.<br />

This was especially true for [Cr(Brphtpy)2] 3+ ,whereitsemission<br />

was quenched by all four mononucleotides (including deoxythymidine-<br />

5 ′ -monophosphate). This statement, however, requires that the labels for<br />

Figs. 4A <strong>and</strong> B in the paper were accidentally reversed.<br />

8<br />

Photoredox Involving Coordinated Lig<strong>and</strong>s<br />

Whereas Sect. 7 was concerned with examples <strong>of</strong> intermolecular electron<br />

transfer between Cr(III) excited states <strong>and</strong> external substrates, attention<br />

is directed in the present section to cases <strong>of</strong> intramolecular redox chemistry<br />

involving the coordinated lig<strong>and</strong>s. These studies have usually involved<br />

photoexcitation into high-energy LMCT excited states involving<br />

the lig<strong>and</strong> in question, which <strong>of</strong>ten results in the transient formation <strong>of</strong><br />

a Cr(II)/lig<strong>and</strong> radical pair. The subject has been reviewed by Kirk [4],<br />

<strong>and</strong> some representative examples <strong>of</strong> molecules previously investigated are<br />

[Cr(NH3)5Br] 2+ [121] <strong>and</strong> trans-[Cr(tfa)3] (where tfa is the anion <strong>of</strong> 1,1,1trifluoro-2,4-pentanedione)<br />

[122]. Some <strong>of</strong> the more recent contributions in<br />

this area are discussed in the following two sections.<br />

8.1<br />

Photolabilization <strong>of</strong> NO from Cr(III)-Coordinated Nitrite<br />

It has been recently established that nitric oxide (NO) regulates a number<br />

<strong>of</strong> mammalian biological processes, including blood pressure, neurotransmission,<br />

<strong>and</strong> smooth muscle relaxation [123]. Additionally, tumor cells are<br />

particularly sensitive to NO, which induces programmed cell death [124]<br />

<strong>and</strong> limits metasis [125]. In response to these findings, Ford <strong>and</strong> coworkers<br />

have developed a range <strong>of</strong> air-stable, water-soluble nitrito-Cr(III) macrocyclic<br />

complexes, which display photochemically activated NO release [126–128].<br />

The initial study involved the complex trans-[Cr(cyclam)(ONO)2] + [126],<br />

which for convenience is labeled structure I in Fig. 19.


60 N.A.P. Kane-Maguire<br />

Fig. 19 Representative trans-[Cr(macrocycle)(ONO)2] + complexes<br />

The only product <strong>of</strong> 436 nm continuous photolysis <strong>of</strong> I in deaerated pH 7<br />

aqueous solution was trans-[Cr(cyclam)(H2O)(ONO)] 2+ ,formedinasubstitution<br />

step in only low quantum yield (φaq = 0.009). However, when the<br />

same photolysis was performed in air-saturated solution, a very different<br />

product was formed in much higher yield (φO2 = 0.27). On the basis <strong>of</strong><br />

mass spectral <strong>and</strong> EPR evidence, this Cr final product was formulated as the<br />

oxo-Cr(V) species, trans-[Cr(cyclam)(O)(ONO)] 2+ . In addition, NO gas release<br />

was confirmed employing an NO-specific electrode sensor. Transient<br />

absorption spectral studies indicated that NO release occurred in a rapidly<br />

reversible earlier step (see Scheme 2) involving homolytic cleavage <strong>of</strong> coordinated<br />

nitrite ion in I to yield the transitory oxo-Cr(IV) product, trans-<br />

[Cr(cyclam)(O)(ONO)] + .<br />

Scheme 2 Photoinitiated reaction scheme for trans-[Cr(cyclam)(ONO)2] +<br />

In deaerated solution, the transient rapidly reforms the parent complex,<br />

resulting in simple photoaquation being the only observed net reaction.<br />

Under air-saturated conditions, however, the transient species is very rapidly<br />

scavenged by dissolved O2 to generate the oxo-Cr(V) final product, leading to<br />

the net release <strong>of</strong> NO gas. The overall mechanism is summarized in Scheme 2.<br />

In an effort to utilize this reaction scheme in a more practical NO delivery<br />

system, the Ford group subsequently synthesized a series <strong>of</strong> complexes where


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 61<br />

Fig. 20 Mechanism <strong>of</strong> NO release following light absorption by a pendant aromatic antennae<br />

strongly absorbing pendant aromatic chromophores were attached to the<br />

macrocyclic ring [128–130]. Two <strong>of</strong> these second-generation Cr(III)nitrito<br />

systems are shown in Fig. 19, where molecules II <strong>and</strong> III have anthracenyl<br />

<strong>and</strong> pyrenyl pendant arms, respectively. The normally strong fluorescence<br />

<strong>of</strong> these tethered aromatics was largely quenched upon Cr(III) coordination,<br />

consistent with fast intramolecular energy transfer to the lower lying<br />

Cr(III) lig<strong>and</strong> field excited states followed by NO gas release according to<br />

Scheme 2 [128]. The overall NO-generating process for these promising lightgathering<br />

antennae systems is depicted in Fig. 20 for the pyrenyl-pendant<br />

complex (molecule III).<br />

8.2<br />

Photogeneration <strong>of</strong> Nitrido Complexes from Cr(III) Coordinated Azide<br />

The complex [Cr(NH3)5N3] 2+ containing the azido lig<strong>and</strong>, N3 – ,wasthesubject<br />

<strong>of</strong> several photochemical studies during the 1970s [131–134]. The results<br />

from irradiations in the LMCT region (λ ≤ 330 nm)identifiedthepresence<strong>of</strong><br />

two competing processes, involving the formation <strong>of</strong> azide radical, N3 · <strong>and</strong><br />

nitrene, N – , intermediates (Eqs. 3 <strong>and</strong> 4, respectively) [132–134]:<br />

[CrIII (NH3)5N3] 2+ + hν → [CrII (NH3)5(N3·)] 2+ → 1.5N2<br />

[CrIII (NH3)5N3] 2+ + hν → [CrIII (NH3)5N] 2+ +N2<br />

Product analysis was complicated by the thermal reactions <strong>of</strong> the radical<br />

species generated. However, based in part on the differences expected in the<br />

N2 gas yields for the two processes, Katz <strong>and</strong> Gafney [132, 134] concluded that<br />

initial nitrene formation was the dominant reaction pathway. Although the fi-<br />

(3)<br />

(4)


62 N.A.P. Kane-Maguire<br />

nal fate <strong>of</strong> the nitrene intermediate was not established, Sriram <strong>and</strong> Endicott<br />

noted that quasi-thermodynamic calculations suggested the lowest energy<br />

product ground state for this system would be a Cr(V) species [133].<br />

More recent photochemical investigations on azido complexes <strong>of</strong> Cr(III)<br />

have focused on systems containing tetradentate lig<strong>and</strong>s such as N2O4 Schiffbases<br />

[135] <strong>and</strong> N3 or N4 macrocyclic lig<strong>and</strong>s [136, 137] as stable non-leaving<br />

groups. For many <strong>of</strong> these systems, air-stable, solid products have been isolated,<br />

<strong>and</strong> have been fully characterized by a variety <strong>of</strong> spectroscopic probes<br />

(including X-ray crystallography). These studies provide convincing evidence<br />

for the formation <strong>of</strong> stable Cr(V) complexes containing the nitrido lig<strong>and</strong>,<br />

N3– , formed via the generic photoreaction shown in Eq. 5:<br />

[CrIII – N3] 2+ + hν → [CrV ≡ N] 2+ +N2<br />

(5)<br />

Evidence for a Cr(V) product is based on the diagnostic EPR signature displayed<br />

by this d1 metal ion. The presence <strong>of</strong> a Cr ≡ Ntriplebondinthe<br />

product is also in accord with the short Cr – N bond distances observed in Xray<br />

crystallographic studies, <strong>and</strong> the presence <strong>of</strong> a strong infrared absorption<br />

in the 1020–1150 cm –1 region [135–138].<br />

During the present review period, the charge transfer photochemistry <strong>of</strong><br />

several azido-Cr(III) complexes containing new Schiff-base lig<strong>and</strong>s as the<br />

non-leaving groups were examined [139, 140]. In the first <strong>of</strong> these papers,<br />

no crystallographic evidence was presented for Cr(V)-nitrido product formation,<br />

but this product assignment was strongly supported by EPR <strong>and</strong> infrared<br />

spectral results [139]. In the second contribution, a potentially valuable<br />

biochemical application <strong>of</strong> azido-Cr(III) photochemistry is reported by Shrivastava<br />

<strong>and</strong> Nair [140]. An azido-Cr(III) Schiff-base complex was irradiated<br />

in the presence <strong>of</strong> bovine serum albumin (BSA), <strong>and</strong> the photolyte examined<br />

by sodium dodecyl sulfate-polyacrylamide disc electrophoresis (SDS-PAGE).<br />

The SDS-PAGE results revealed that the BSA protein was cleaved at multiple<br />

sites, non-specifically into smaller peptide fragments. The protein cleavage<br />

was attributed to the azido-Cr(III) complex binding at multiple sites, <strong>and</strong> being<br />

subsequently converted to the reactive nitrido-Cr(V) species upon light<br />

activation. This light-promoted protease activity bears some analogies with<br />

the photonuclease activity discussed earlier for [Cr(diimine)3] 3+ interactions<br />

with DNA (Sect. 7.1). A non-selective photonuclease could be utilized in<br />

a variety <strong>of</strong> applications, including protein sequencing.<br />

9<br />

Final Comments<br />

In this chapter, the author has attempted to provide an overview <strong>of</strong> recent<br />

progress in the field <strong>of</strong> Cr(III) photochemistry <strong>and</strong> photophysics, with a more<br />

detailed focus on certain topics <strong>of</strong> interest. For cohesiveness, it was not pos-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 63<br />

sible to cover some aspects <strong>of</strong> the field that did not naturally fit within the<br />

scope <strong>of</strong> those focus areas.<br />

Included in the material not covered, is the recent contribution by Ronco<br />

<strong>and</strong> coworkers [141], where advantage is taken <strong>of</strong> the sometimes exquisite dependence<br />

on environmental factors <strong>of</strong> the emission intensity <strong>and</strong> lifetime <strong>of</strong><br />

Cr(III) polypyridyls. This sensitivity was used to probe hydrophobic sites in<br />

anionic polyelectrolytes, where such information may provide useful guidance<br />

in attempts to enhance the rates <strong>of</strong> photoactivated electron transfer<br />

processes <strong>and</strong>/or retard recombination events. More recently [142], in an effort<br />

to more finely tune 2 Eg excited state properties, their group synthesized<br />

a range <strong>of</strong> mixed lig<strong>and</strong> polypyridyls <strong>of</strong> Cr(III), using a procedure we had<br />

developed earlier [143]. Another area not addressed is the increasing use <strong>of</strong><br />

Cr(III) complexes in spectral hole-burning experiments to overcome spectral<br />

broadening in condensed phases [144, 145]. Finally, one <strong>of</strong> the more intriguing<br />

topics omitted is a report in Nature in 2000 describing an experimental<br />

confirmation [146] <strong>of</strong> a theoretical prediction, termed magnetochiral dichroism,<br />

that a chiral medium would absorb light traveling parallel to a magnetic<br />

field differently from light traveling antiparallel [147]. The compound investigated<br />

was [Cr(oxalate)3] 3– , <strong>and</strong> a very small, strongly excitation wavelengthdependent,<br />

induction <strong>of</strong> optical activity was observed on laser irradiation in<br />

a very powerful magnetic field (up to 15 Tesla).<br />

In conclusion, it is noted that although the subject <strong>of</strong> Cr(III) photochemistry<br />

<strong>and</strong> photophysics is unlikely to reassume the degree <strong>of</strong> prominence it<br />

held up until the early 1970s, the present condition <strong>of</strong> the field is good <strong>and</strong> the<br />

long-term prognosis is excellent. Who is to say that the sign on my laboratory<br />

door which boldly states: “Chromium – The Final Frontier”, will not one day<br />

be more than just a catchy phrase?<br />

Acknowledgements The author gratefully acknowledges stimulating discussions with Paul<br />

Wagenknecht <strong>and</strong> John Wheeler during the preparation <strong>of</strong> this chapter. In early 2006, the<br />

field <strong>of</strong> Cr(III) photochemistry <strong>and</strong> photophysics lost one <strong>of</strong> its young luminaries, Marc<br />

Perkovic. This chapter is dedicated to his memory.<br />

References<br />

1. Balzani V, Carassiti V (1970) <strong>Photochemistry</strong> <strong>of</strong> coordination compounds. Academic,<br />

London<br />

2. Zinato E (1975) Substitutional photochemistry <strong>of</strong> first-row transition metal elements.<br />

In: Adamson AW, Fleischauer PD (eds) Concepts in inorganic photochemistry.<br />

Wiley, New York (Chap 4)<br />

3. Roundhill DM (1994) <strong>Photochemistry</strong> <strong>and</strong> photophysics <strong>of</strong> metal complexes.<br />

Plenum, New York<br />

4. Kirk AD (1999) Chem Rev 99:1607<br />

5. Zinato E, Riccieri P (2001) Coord Chem Rev 211:5<br />

6. Irwin G, Kirk AD (2001) Coord Chem Rev 211:25


64 N.A.P. Kane-Maguire<br />

7. Derwahl A, Wasgestian F, House DA, Robinson WT (2001) Coord Chem Rev 211:45<br />

8. Kane-Maguire NAP, Wheeler JF (2001) Coord Chem Rev 211:145<br />

9. Forster LS (2002) Coord Chem Rev 227:59<br />

10. Fucaloro AF, Forster LS, Glover SG, Kirk AD (1985) Inorg Chem 24:4242<br />

11. Forster LS (1990) Chem Rev 90:331<br />

12. Rojas G, Magde D (1983) Chem Phy Lett 102:399<br />

13. Rojas GE, Dupuy C, Sexton DA, Magde D (1986) J Phys Chem 90:87<br />

14. Kirk AD, Hoggard PE, Porter GB, Rockley MG, Windsor MW (1976) Chem Phys Lett<br />

37:199<br />

15. Juban EA, McCusker JK (2005) J Am Chem Soc 127:6857<br />

16. Forster LS (1969) Transition Met Chem 5:1<br />

17. Chen SN, Porter GB (1970) Chem Phys Lett 6:41<br />

18. Wasgestian F (1972) J Phys Chem 76:1947<br />

19. Riccieri P, Zinato E (1990) Inorg Chem 29:5035<br />

20. Riccieri P, Zinato E, Aliboni A (1997) Inorg Chem 36:2279<br />

21. Sabbatini N, Balzani V (1972) J Am Chem Soc 94:7587<br />

22. Balzani V, Moggi L, Manfrin MF, Bolletta F, Laurence GS (1975) Coord Chem Rev<br />

15:321<br />

23. Endicott JF, Ramasami T, Tamilarasan R, Lessard RB, Ryu CK (1987) Coord Chem<br />

Rev 77:1<br />

24. Endicott JF (1985) Comments Inorg Chem 3:349<br />

25. Vanquickenborne LG, Ceulemans A (1978) J Am Chem Soc 100:475<br />

26. Vanquickenborne LG, Ceulemans A (1979) Inorg Chem 18:3475<br />

27. Vanquickenborne LG, Ceulemans A (1983) Coord Chem Rev 48:157<br />

28. Kirk AD (1993) Comments Inorg Chem 14:89<br />

29. Vanquickenborne LG, Coussens B, Postelmans D, Ceulemans A, Pierloot K (1992)<br />

Inorg Chem 31:539<br />

30. Kane-Maguire NAP, Langford CH (1972) J Am Chem Soc 94:2125<br />

31. Kane-Maguire NAP, Langford CH (1976) Inorg Chem 15:464<br />

32. Jamieson MA, Serpone N, H<strong>of</strong>fman MZ (1981) Coord Chem Rev, p 121<br />

33. Bolletta F, Maestri M, Moggi L, Jamieson MA, Serpone N, Henry MS (1983) H<strong>of</strong>fman<br />

Inorg Chem 22:2502<br />

34. Lilie J, Waltz WL, Lee SH, Gregor LL (1986) Inorg Chem 25:4487<br />

35. Vleck A Jr (2000) Coord Chem Rev 933:200–202<br />

36. McCusker JK (2003) Acc Chem Res 36:876<br />

37. Juban EA, Smeigh AL, Monat JE, McCusker JK (2006) Coord Chem Rev, p 783<br />

38. Turro NJ (1991) Modern molecular photochemistry. University Science Books, Mill<br />

Valley, CA<br />

39. Kane-Maguire NAP, Richardson DE, Toney CG (1976) J Am Chem Soc 98:3996<br />

40. Kane-Maguire NAP, Phifer JE (1976) Toney Inorg Chem 15:593<br />

41. S<strong>and</strong>rini D, G<strong>and</strong>olfi MT, Moggi L, Balzani V (1978) J Am Chem Soc 100:1463<br />

42. Sasseville RLP, Langford CH (1980) Inorg Chem 19:2850<br />

43. Langford CH (1984) Acc Chem Res, p 96<br />

44. Monat JE, McCusker JK (2000) J Am Chem Soc 122:4092<br />

45. Yang G, Leng W, Zhang Y, Chen Z, Van Houten J (1999) Polyhedron 18:1273<br />

46. Inamo M, Hoshino M (1999) Photochem Photobiol 70:596<br />

47. Inamo M, Nakaba H, Nakajima K, Hoshino M (2000) Inorg Chem 39:4417<br />

48. Jeoung SC, Kim D, Cho DW, Yoon M (2000) J Phys Chem A 104:4816<br />

49. Mytykh OV, Martin SE, Wheeler JF, Kane-Maguire NAP (2000) Inorg Chim Acta<br />

311:143


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 65<br />

50. Inamo M, Eba K, Nakano K, Itoh N (2003) Inorg Chem 42:6095<br />

51. Inamo M, Matsubara N, Nakajima K, Iwayama T, Okimi H, Hoshino M (2005) Inorg<br />

Chem 44:6445<br />

52. Gouterman M, Hanson LK, Khalil GE, Leenstra WR, Buchler JW (1975) J Chem Phys<br />

62:2343<br />

53. Kutal C, Adamson AW (1971) J Am Chem Soc 93:5581<br />

54. Kutal C, Adamson AW (1973) Inorg Chem 12:1990<br />

55. Kane-Maguire NAP, Crippen WS, Miller PK (1983) Inorg Chem 22:696<br />

56. Kane-Maguire NAP, Wallace KC, Cobranchi DP, Derrick JM, Speece DG (1986) Inorg<br />

Chem 25:2101<br />

57. Kane-Maguire NAP, Wallace KC, Miller DB (1985) Inorg Chem 24:597<br />

58. Friesen DA, Lee SH, Lilie J, Waltz WL, Vincze L (1991) Inorg Chem 30:1975<br />

59. Kane-Maguire NAP, Wallace KC, Speece DG (1986) Inorg Chem 25:4650<br />

60. Ramasami T, Endicott JF, Brubaker GR (1983) J Phys Chem 87:5057<br />

61. Endicott JF, Lessard RB, Lei Y, Ryu CK, Tamilarasan R (1986) ACS Symp Ser 307:85<br />

62. Perkovic MW, Endicott JF (1990) J Phys Chem 94:1217<br />

63. Perkovic MW, Heeg MJ, Endicott JF (1991) Inorg Chem 30:3140<br />

64. Endicott JF, Perkovic MW, Heeg MJ, Ryu CK, Thompson D (1997) Adv Chem Ser<br />

253:199<br />

65. Lessard RB, Heeg MJ, Bur<strong>and</strong>a T, Perkovic MW, Schwartz CL, Rudong Y, Endicott JF<br />

(1992) Inorg Chem 31:3091<br />

66. Comba P, Mau AWH, Sargeson AM (1985) J Phys Chem 89:394<br />

67. Comba P, Creaser II, Gahan LR, Harrowfield JM, Lawrance GA, Martin LL,<br />

Mau AWH, Sargeson AM, Sasse WHF, Snow MR (1986) Inorg Chem 25:384<br />

68. Irwin G, Kirk AD, Mackay I, Nera J (2002) Inorg Chem 41:874<br />

69. Wright-Garcia K, Basinger J, Williams S, Hu C, Wagenknecht PS (2003) Inorg Chem<br />

42:4885<br />

70. Wagenknecht PS, Hu C, Ferguson D, Nathan LC, Hancock RD, Whitehead JR,<br />

Wright-Garcia K, Vagnini MT (2005) Inorg Chem 44:9518<br />

71. Vagnini MT, Kane-Maguire NAP, Wagenknecht PS (2006) Inorg Chem 45:3789<br />

72. Kane-Maguire NAP, Hanks TW, Jurs DG, Tollison RM, Heatherington AL, Ritzenthaler<br />

LM, McNulty LM, Wilson HM (1995) Inorg Chem 34:1121<br />

73. Harris JE, Desai, Seaver KE, Watson RT, Kane-Maguire NAP, Wheeler JF (2001)<br />

J Chromat A 919:427<br />

74. Brown KN, Geue RJ, Moran G, Ralph SF, Riesen H, Sargeson AM (1998) Chem<br />

Commun, p 2291<br />

75. Vagnini MT, Rutledge WC, Hu C, VanDerveer DG, Wagenknecht PS (2007) Inorg<br />

Chim Acta 360:1482<br />

76. Bolletta F, Maestri M, Moggi L, Balzani V (1973) J Am Chem Soc 95:7864<br />

77. Bolletta F, Maestri M, Balzani V (1976) J Phys Chem 80:2499<br />

78. Kane-Maguire NAP, Toney CG, Swiger B, Adamson AW, Wright RE (1977) Inorg<br />

Chim Acta 22:L11<br />

79. Chen Y, Perkovic MW (2001) Inorg Chim Acta 317:127<br />

80. Iwamura M, Otsuka T, Kaizu Y (2002) Inorg Chim Acta 333:57<br />

81. Wagenknecht PS, Kane-Maguire NAP, Speece DG, Helwic N (2002) Inorg Chem<br />

41:1229<br />

82. Von Arx ME, Hauser A (2002) J Phys Chem A 106:7106<br />

83. Subhan MDA, Nakata H, Suzuki T, Choi JH, Kaizaki S (2003) J Lumin 101:307<br />

84. Iwamura M, Otsuka T, Kaizu Y (2004) Inorg Chim Acta 357:1565<br />

85. Iwamura M, Morita M (2004) Inorg Chim Acta 357:3451


66 N.A.P. Kane-Maguire<br />

86. Otsuka T, Iwamura M, Kaizu Y (2006) Inorg Chim Acta 359:1351<br />

87. Balzani V, Indelli MT, S<strong>and</strong>rini D, Sc<strong>and</strong>ola F (1980) J Phys Chem 84:852<br />

88. G<strong>and</strong>olfi MT, Maestri M, S<strong>and</strong>rini D, Balzani V (1983) Inorg Chem 22:3435<br />

89. Endicott JF, Ramasami T, Gaswick DC, Tamilarasan GR, Heeg MJ, Brubaker GR,<br />

Pyke SC (1983) J Am Chem Soc 105:5301<br />

90. Gamache RE, Rader RA, McMillin DR (1985) J Am Chem Soc 107:1141<br />

91. Endicott JF (1985) Coord Chem Rev 64:293<br />

92. Tamilarasan R, Endicott JF (1986) J Phys Chem 90:1027<br />

93. Maestri M, S<strong>and</strong>rini D, Balzani V, Belser P, Von Zelewski A (1984) Chem Phys Lett<br />

110:611<br />

94. Balzani V, Juris A (2001) Coord Chem Rev 211:97<br />

95. Maharaj U, Winnik MA (1981) Chem Phys Lett 82:29<br />

96. Gafney HD, Adamson AW (1972) J Am Chem Soc 94:8238<br />

97. Juris A, Balzani V, Barigelletti F, Campagna S, Belser P, von Zelewsky A (1988) Coord<br />

Chem Rev 84:85<br />

98. Balzani V, Juris A, Venturi M, Campagna S, Serroni S (1996) Chem Rev 96:759<br />

99. Balzani V, Venturi M, Credi A (2003) Molecular Devices <strong>and</strong> Machines. Wiley, Weinheim<br />

100. Bolletta F, Maestri M, Moggi L, Balzani V (1975) J Chem Soc Chem Commun, p 901<br />

101. Ballardini R, Varani G, Sc<strong>and</strong>ola F (1976) J Am Chem Soc 98:7432<br />

102. Ballardini R, Varani G, Indelli MT, Sc<strong>and</strong>ola F, Balzani V (1978) J Am Chem Soc<br />

100:7219<br />

103. Brunschwig B, Sutin N (1978) J Am Chem Soc 100:7568<br />

104. Juris A, Manfrin MF, Maestri M, Serpone N (1978) Inorg Chem 17:2258<br />

105. Serpone N, Jamieson MA, Henry MS, H<strong>of</strong>fman MZ, Bolletta F, Maestri M (1979)<br />

J Am Chem Soc 101:2907<br />

106. Jamieson MA, Serpone N, H<strong>of</strong>fman MZ (1981) Coord Chem Rev 39:121<br />

107. Navin G, Sutin N (1976) Inorg Chem 15:496<br />

108. Tiyabhorn A, Zahir KO (1996) Can J Chem 74:336<br />

109. Pagliero D, Argüello GA (2001) J Photochem Photobiol A 138:207<br />

110. Pizzocaro C, Bolte M, Sun H, H<strong>of</strong>fman MZ (1994) New J Chem 18:737<br />

111. DeRosa MC, Crutchley RJ (2002) Coord Chem Rev 233–234:351<br />

112. Kirsch-De Mesmaeker A, Lecomte J-P, Kelly JM (1996) Top Curr Chem 177:25<br />

113. Erkkila KE, Odom DT, Barton JK (1999) Chem Rev 99:2777<br />

114. Lecomte J-P, Kirsch-De Mesmaeker A, Feeny MM, Kelly JM (1995) Inorg Chem<br />

34:6481<br />

115. Burrows CJ, Muller JG (1998) Chem Rev 98:1109<br />

116. Fleisher MB, Waterman KC, Turro NJ, Barton JK (1986) Inorg Chem 25:3549<br />

117. Mei H-Y, Barton JK (1988) Proc Natl Acad Sci USA 85:1339<br />

118. Watson RT, Desai N, Wildsmith J, Wheeler JF, Kane-Maguire NAP (1999) Inorg<br />

Chem 38:2683<br />

119. Barker KD, Benoit BR, Bordelon JA, Davis RJ, Delmas AS, Mytykh OV, Petty JT,<br />

Wheeler JF, Kane-Maguire NAP (2001) Inorg Chim Acta 322:74<br />

120. Vaidyanathan VG, Nair BU (2004) Eur J Inorg Chem, p 1840<br />

121. Sriram R, Endicott JF (1976) J Chem Soc Chem Commun, p 683<br />

122. Kutal C, Yang DB, Ferraudi G (1980) Inorg Chem 19:2907<br />

123. Fukuto JM, Wink DA (1999) Metal Ions Biol Syst 36:547<br />

124. Boyd CS, Cadenas E (2002) J Biol Chem 383:411<br />

125. Xie K, Huang S (2003) Free Radic Biol <strong>and</strong> Med 34:969<br />

126. DeLeo M, Ford PC (1999) J Am Chem Soc 121:1980


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Chromium 67<br />

127. DeLeo M, Ford PC (2000) Coord Chem Rev 208:47<br />

128. DeRosa F, Bu X, Ford PC (2005) Inorg Chem 44:4157<br />

129. DeRosa F, Bu X, Ford PC (2003) Inorg Chem 42:4171<br />

130. DeRosa F, Bu X, Pohaku K, Ford PC (2005) Inorg Chem 44:4166<br />

131. Vogler A (1971) J Am Chem Soc 93:5912<br />

132. Katz M, Gafney HD (1976) J Am Chem Soc 98:7458<br />

133. Sriram R, Endicott JF (1977) Inorg Chem 16:2766<br />

134. Katz M, Gafney HD (1978) Inorg Chem 17:93<br />

135. Arshankov SI, Poznjak ALZ (1981) Z Anorg Allg Chem 481:201<br />

136. Niemann A, Bossek U, Haselhorst G, Weighardt K, Nuber B (1996) Inorg Chem<br />

35:906<br />

137. Meyer K, Bendix J, Bill E, Weyhermüller T, Wieghardt K (1998) Inorg Chem<br />

138. Bendix J, Wilson SR, Prussak-Wiechowska T (1998) Acta Crystallogr C54:923<br />

139. Kanthimathi M, Nair BU (2004) Transition Met Chem 29:751<br />

140. Shrivastava HY, Nair BU (2004) J Inorg Biochem 98:991<br />

141. Ronco S, Persing B, Mortinsen R, Barber J, Shotwell S (2000) Inorg Chim Acta<br />

308:107<br />

142. Isaacs M, Sykes AG, Ronco S (2006) Inorg Chim Acta 359:3847<br />

143. Barker K, Barnett K, Connell S, Glaeser J, Wallace A, Wildsmith J, Herbert B,<br />

Wheeler JF, Kane-Maguire NAP (2001) Inorg Chim Acta 316:41<br />

144. Lewis ML, Riesen H (2002) Phys Chem Chem Phys 4:4845<br />

145. Riesen H, Wallace L (2003) Phys Chem Comm 6:9<br />

146. Rikken GLJA, Raupach E (2000) Nature 405:932<br />

147. Wagnière G, Meir A (1982) Chem Phys Lett 93:78


Top Curr Chem (2007) 280: 69–115<br />

DOI 10.1007/128_2007_128<br />

© Springer-Verlag Berlin Heidelberg<br />

Published online: 24 May 2007<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong><br />

<strong>Compounds</strong>: Copper<br />

Nicola Armaroli (✉) · Gianluca Accorsi · François Cardinali · Andrea Listorti<br />

Molecular Photoscience Group, Istituto per la Sintesi Organica e la Fotoreattività,<br />

Consiglio Nazionale delle Ricerche, Via Gobetti 101, 40129 Bologna, Italy<br />

nicola.armaroli@is<strong>of</strong>.cnr.it<br />

1 AnOverview<strong>of</strong>Copper............................. 70<br />

1.1 HistoricalNotes,CurrentUse,ConsumptionTrends ............. 70<br />

1.2 Chemical Properties, <strong>Coordination</strong> Geometries,<br />

ExcitedStates—Cu(I)vs.Cu(II) ........................ 71<br />

1.3 CopperinBiology................................ 73<br />

1.4 Cu(I)inSupramolecularChemistry ...................... 78<br />

2 Cu(I)-Bisphenanthroline Complexes ...................... 81<br />

2.1 GroundStateGeometry............................. 81<br />

2.2 AbsorptionSpectra ............................... 83<br />

2.3 Excited State Distortion: Pulsed X-ray<br />

<strong>and</strong>TransientAbsorptionSpectroscopy.................... 86<br />

2.4 EmissiveExcitedState(s)<strong>and</strong>LuminescenceSpectra............. 88<br />

2.5 Photoinduced Processes in Multicomponent Systems Based on<br />

[Cu(NN)2] + Complexes............................. 92<br />

2.6 BimolecularQuenchingProcesses ....................... 94<br />

3 Heteroleptic Diimine/Diphosphine [Cu(NN)(PP)] + Complexes ....... 95<br />

3.1 PhotophysicalProperties ............................ 95<br />

3.2 OLED<strong>and</strong>LECDevices............................. 99<br />

4 Cuprous Clusters ................................ 101<br />

4.1 CuprousHalideClusters ............................ 101<br />

4.2 CuprousIodideClusters ............................ 102<br />

4.3 OtherCopperClusters ............................. 105<br />

5 Miscellanea <strong>of</strong> Cu(I) Luminescent Complexes ................ 107<br />

6 Conclusions <strong>and</strong> Perspectives ......................... 109<br />

References ....................................... 110<br />

Abstract Cu(I) complexes <strong>and</strong> clusters are the largest class <strong>of</strong> compounds <strong>of</strong> relevant<br />

photochemical <strong>and</strong> photophysical interest based on a relatively abundant metal element.<br />

Interestingly, Nature has given an essential role to copper compounds in some biological<br />

systems, relying on their kinetic lability <strong>and</strong> versatile coordination environment. Some<br />

basic properties <strong>of</strong> Cu(I) <strong>and</strong> Cu(II) such as their coordination geometries <strong>and</strong> electronic<br />

levels are compared, pointing out the limited significance <strong>of</strong> Cu(II) compounds<br />

(d 9 configuration) in terms <strong>of</strong> photophysical properties. Well-established synthetic protocols<br />

are available to build up a variety <strong>of</strong> molecular <strong>and</strong> supramolecular architectures


70 N. Armaroli et al.<br />

(e.g. catenanes, rotaxanes, knots, helices, dendrimers, cages, grids, racks, etc.) containing<br />

Cu(I)-based centers <strong>and</strong> exhibiting photo- <strong>and</strong> electroluminescence as well as<br />

light-induced intercomponent processes. By far the largest class <strong>of</strong> copper complexes investigated<br />

to date is that <strong>of</strong> Cu(I)-bisphenanthrolines ([Cu(NN)2] + ) <strong>and</strong> recent progress<br />

in the rationalization <strong>of</strong> their metal-to-lig<strong>and</strong> charge-transfer (MLCT) absorption <strong>and</strong> luminescence<br />

properties are critically reviewed, pointing out the criteria by which it is now<br />

possible to successfully design highly emissive [Cu(NN)2] + compounds, a rather elusive<br />

goal for a long time. To this end the development <strong>of</strong> spectroscopic techniques such as<br />

light-initiated time-resolved X-ray absorption spectroscopy (LITR-XAS) <strong>and</strong> femtosecond<br />

transient absorption have been rather fruitful since they have allowed us to firmly ground<br />

the indirect pro<strong>of</strong>s <strong>of</strong> the molecular rearrangements following light absorption that had<br />

accumulated in the past 20 years. A substantial breakthrough towards highly emissive<br />

Cu(I) coordination compounds is constituted by heteroleptic Cu(I) complexes containing<br />

both N- <strong>and</strong> P-coordinating lig<strong>and</strong>s ([Cu(NN)(PP)] + ) which may exhibit luminescence<br />

quantum yields close to 30% in deaerated CH2Cl2 solution <strong>and</strong> have been successfully<br />

employed as active materials in OLED <strong>and</strong> LEC optoelectronic devices. Also copper clusters<br />

may exhibit luminescence b<strong>and</strong>s <strong>of</strong> halide-to-metal charge transfer (XMCT) <strong>and</strong>/or<br />

cluster centered (CC) character <strong>and</strong> they are briefly reviewed along with miscellaneous<br />

Cu(I) compounds that recently appeared in the literature, which show luminescence<br />

b<strong>and</strong>s ranging from the blue to the red spectral region.<br />

Keywords Clusters · Copper · Electron transfer · Energy transfer · Luminescence ·<br />

OLED · Phenanthroline<br />

1<br />

An Overview <strong>of</strong> Copper<br />

1.1<br />

Historical Notes, Current Use, Consumption Trends<br />

Copper was known to some <strong>of</strong> the oldest civilizations on record, <strong>and</strong> has a history<strong>of</strong>usethatisatleast10<br />

000 years old. A copper pendant was found in<br />

what is now northern Iraq that dates to 8700 B.C. <strong>and</strong> by 5000 B.C. there are<br />

signs <strong>of</strong> copper smelting from simple copper compounds such as malachite or<br />

azurite. This process appears to have been developed independently in several<br />

parts <strong>of</strong> the world since several centuries B.C., including Anatolia, China,<br />

Central America <strong>and</strong> West Africa. The Egyptians found that, upon addition<br />

<strong>of</strong> small amounts <strong>of</strong> tin, copper becomes easier to cast, <strong>and</strong> bronze alloys<br />

were extensively found in the Nile valley. The use <strong>of</strong> bronze was so pervasive<br />

in a certain era <strong>of</strong> civilization that the period spanning from 2500 to 600<br />

B.C. is named the Bronze Age. In Roman times, copper became known as aes<br />

Cyprium, aes being the generic Latin term for copper alloys such as bronze<br />

or other metals, <strong>and</strong> Cyprium because so much <strong>of</strong> it was mined in the isl<strong>and</strong><br />

<strong>of</strong> Cyprus. From this, the phrase was simplified to cuprum (originating the<br />

current chemical symbol) <strong>and</strong> then eventually Anglicized into copper.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 71<br />

Copper is usually found in Nature in association with sulfur. Pure copper<br />

metal is generally produced from a multistage process, beginning with the<br />

mining <strong>and</strong> concentrating <strong>of</strong> low-grade ores containing copper sulfide minerals,<br />

<strong>and</strong> followed by smelting <strong>and</strong> electrolytic refining to produce a pure<br />

copper cathode. An increasing share <strong>of</strong> this metal is produced from acid<br />

leaching <strong>of</strong> oxidized ores. Because <strong>of</strong> its properties which include high ductility,<br />

malleability, thermal <strong>and</strong> electrical conductivity, <strong>and</strong> resistance to corrosion,<br />

copper has become a major industrial metal, ranking third after iron<br />

<strong>and</strong> aluminum in terms <strong>of</strong> quantities consumed. Electrical uses <strong>of</strong> copper,<br />

including power transmission <strong>and</strong> generation, building wiring, telecommunication,<br />

<strong>and</strong> electronic products, account for about three quarters <strong>of</strong> total<br />

employment.Today,copperby-productsfrommanufacturing<strong>and</strong>obsolete<br />

copper products are readily recycled <strong>and</strong> contribute significantly to supply.<br />

This is becoming a necessity due to the increasing difficulty <strong>of</strong> production to<br />

meet current world dem<strong>and</strong> which has led to a quintuplication <strong>of</strong> the copper<br />

price during the last seven years, rising from $0.60/pound in June 1999<br />

to $3.75/pound in May 2006. In 2005 14.9 million tons <strong>of</strong> copper were mined<br />

around the world; the global world reserves, economically recoverable with<br />

current technologies, amount to 470 million tons [1]. It has been recently<br />

estimated that ca. 25% <strong>of</strong> the copper stock initially available in the lithosphere<br />

has been already placed in use or in wastes from which it will probably<br />

never be recovered. This poses concern about the sustainability <strong>of</strong> current<br />

consumption trends <strong>of</strong> such a valuable commodity in the mid-long term [2].<br />

1.2<br />

Chemical Properties, <strong>Coordination</strong> Geometries, Excited States—Cu(I) vs. Cu(II)<br />

Copper is a transition element belonging to the same group <strong>of</strong> the periodic<br />

table as gold <strong>and</strong> silver, these elements are sometimes referred to as the coinage<br />

metals in recognition <strong>of</strong> their historically widespread use in stamping coins.<br />

Copper has a single s electron outside the filled 3d shell but its properties have<br />

essentially nothing in common with alkali metals except for the possibility <strong>of</strong><br />

assuming the +1 oxidation state. The filled d shell is not very effective in shielding<br />

the s electron from the nuclear charge, so the first ionization enthalpy <strong>of</strong> Cu<br />

is higher than that <strong>of</strong> the alkali metals. Since the electrons <strong>of</strong> the d shell are also<br />

involved in metallic bonding, the heat <strong>of</strong> sublimation <strong>and</strong> the melting point <strong>of</strong><br />

Cu are also much higher than those <strong>of</strong> the alkalis. The above factors, taken together,<br />

are responsible for the noble character <strong>of</strong> copper. Indeed copper is the<br />

only industrial pure metal, used on a massive scale, exhibiting a positive electrochemical<br />

potential: for this reason it is not corroded by acids, unless they<br />

are strongly oxidizing like HNO3 <strong>and</strong> H2SO4.<br />

Copper in solution has two common oxidation states: + 1 <strong>and</strong> + 2. Because<br />

<strong>of</strong> their intrinsically superior photochemical <strong>and</strong> photophysical properties<br />

(vide infra), in this review our attention is focused on Cu(I) complexes, which


72 N. Armaroli et al.<br />

can be classified in three main categories, i.e. anionic complexes (e.g. alocomplexes),<br />

neutral clusters <strong>and</strong> cationic complexes. The photochemistry <strong>of</strong><br />

Cu(I) complexes, also related to environmental aspects, has already been<br />

reviewed [3, 4], here we will essentially focus on photophysics. Anionic complexes<br />

do not exhibit attractive photophysical properties (e.g. luminescence),<br />

unlike cluster <strong>and</strong> cationic complexes which show a very rich photophysical<br />

behavior. Among the latter, the most extensively investigated are NN-type<br />

(where NN indicates a chelating imine lig<strong>and</strong>, typically 1,10-phenanthroline)<br />

or PP-type (where PP denotes a bisphosphine lig<strong>and</strong>). Both homoleptic<br />

[Cu(NN)2] + <strong>and</strong> heteroleptic [Cu(NN)(PP)] + motifs have been investigated.<br />

The coordination behavior <strong>of</strong> Cu(I) is strictly related to its electronic<br />

configuration. The complete filling <strong>of</strong> d orbitals (d 10 configuration) leads<br />

to a symmetric localization <strong>of</strong> the electronic charge. This situation favors<br />

a tetrahedral disposition <strong>of</strong> the lig<strong>and</strong>s around the metal center in order to<br />

locate the coordinative sites far from one another <strong>and</strong> minimize electrostatic<br />

repulsions (Fig. 1). Clearly, the complete filling <strong>of</strong> d orbitals prevents d-d<br />

metal-centered electronic transitions in Cu(I) compounds. On the contrary,<br />

such transitions are exhibited by d 9 Cu(II) complexes <strong>and</strong> cause relatively intense<br />

absorption b<strong>and</strong>s in the visible (VIS) spectral window. The lowest ones<br />

extend into the near infrared (NIR) region (above 800 nm for the Cu(II) aqua<br />

ion) [5] <strong>and</strong> deactivate via ultrafast non-radiative processes. The fact that<br />

the lowest electronic states <strong>of</strong> Cu(II) complexes are ultra-short lived make<br />

them far less interesting than Cu(I) complexes from the photophysical point<br />

<strong>of</strong> view.<br />

Fig. 1 Tetrahedral coordination environment typical <strong>of</strong> Cu(I) complexes


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 73<br />

Cu(I) cluster compounds are characterized by a variety <strong>of</strong> emitting electronic<br />

levels whereas Cu(I) cationic complexes show only luminescence originating<br />

from metal-to-lig<strong>and</strong> charge-transfer (MLCT) states, as long as empty<br />

π orbitals are easily accessible in the lig<strong>and</strong>s. Such MLCT transitions, which<br />

clearly take advantage <strong>of</strong> the low oxidation potential <strong>of</strong> Cu(I), arealsocommonly<br />

observed in other classes <strong>of</strong> coordination compounds, for example<br />

those <strong>of</strong> d 6 metals like Ru(II)-bipyridines [6] <strong>and</strong> Ir(III)-phenylpyridine complexes<br />

[7].<br />

MLCT electronic transitions in coordination compounds are normally<br />

more intense when compared to MC (metal-centered) ones since they do not<br />

undergo the same prohibitions by orbital symmetry; accordingly MLCT absorption<br />

b<strong>and</strong>s exhibit relatively high molar extinction coefficients. As far as<br />

emission is concerned, when MLCT excited states are the lowest-lying, they<br />

are generally characterized by long lifetimes, <strong>and</strong> potentially intense luminescence,<br />

even though exceptions are possible (vide infra). Complexes exhibiting<br />

long-lived MLCT excited states have been extensively investigated in the last<br />

decades both for a better comprehension <strong>of</strong> fundamental phenomena [8, 9]<br />

<strong>and</strong> for potential applications related to solar light harvesting <strong>and</strong> conversion<br />

[10–12]. Among them the highest attention was probably devoted to<br />

Ru(II) [13], Os(II) [14] <strong>and</strong>, more recently, Ir(III) [7] complexes, however,<br />

economical <strong>and</strong> environmental considerations make Cu(I) compounds interesting<br />

alternatives [15].<br />

As extensively discussed in the literature, long-lived luminescent MLCT<br />

excited states <strong>of</strong> d 6 metal complexes, in particular those <strong>of</strong> Ru(II), can be<br />

strongly affected by the presence <strong>of</strong> upper lying MC levels. The latter can<br />

be partially populated through thermal activation from the MLCT states <strong>and</strong><br />

prompt non-radiative deactivation pathways <strong>and</strong> photochemical degradation<br />

[6, 16]. Closed shell d 10 copper(I) complexes cannot suffer these kinds <strong>of</strong><br />

problems, but undesired non-radiative deactivation channels <strong>of</strong> their MLCT<br />

levels can be favored by other factors, as will be discussed in detail further<br />

on in this review. An orbital diagram illustrating the electronic transitions <strong>of</strong><br />

Ru(II) <strong>and</strong> Cu(I) complexes is reported in Fig. 2.<br />

1.3<br />

Copper in Biology<br />

Copper, even if present in traces, is an essential metal for the growth <strong>and</strong> development<br />

<strong>of</strong> biological systems. Copper plays a fundamental role in cerebral<br />

activity, nervous <strong>and</strong> cardiovascular systems, oxygen transport <strong>and</strong> cell protection<br />

against oxidation. Copper is important to strengthen the bones <strong>and</strong> to<br />

guarantee the performances <strong>of</strong> the immune system [17].<br />

Metals are commonly found as natural constituents <strong>of</strong> proteins <strong>and</strong>, in<br />

thecourse<strong>of</strong>evolution,Naturehaslearnedhowtousethespecialproperties<br />

<strong>of</strong> metal ions to perform a wide variety <strong>of</strong> specific functions associated


74 N. Armaroli et al.<br />

Fig. 2 Qualitative comparison <strong>of</strong> orbitals <strong>and</strong> related electronic transitions in metal complexes<br />

having d 6 (e.g. Ru(II)) <strong>and</strong> d 10 (e.g. Cu(I)) configurations<br />

with life processes. It is puzzling that only a limited number <strong>of</strong> transition<br />

elements <strong>of</strong> the periodic table are utilized in biological systems, among them<br />

iron, copper, <strong>and</strong> zinc are <strong>of</strong> key importance. The criteria by which Nature<br />

chooses metals in biological systems are rather intriguing. One factor that


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 75<br />

seems to be quite important is their relative abundance on Earth. A second<br />

factor is related to the fact that active centers <strong>of</strong> metalloproteins consist <strong>of</strong><br />

kinetically labile <strong>and</strong> thermodynamically stable units. Kinetic lability facilitates<br />

rapid assembly/disassembly <strong>of</strong> the metal centers as well as fast association/dissociation<br />

<strong>of</strong> substrates. Both the above criteria are fulfilled by copper,<br />

which has been present in living organisms since the early stages <strong>of</strong> evolution,<br />

representing a fundamental constituent <strong>of</strong> many biological systems, particularly<br />

proteins, with the function <strong>of</strong> transporting oxygen <strong>and</strong> transferring<br />

electrons.<br />

A key diversity between Cu(I) <strong>and</strong> Cu(II) is the different preferential coordination<br />

geometry. Cu(I) prefers tetrahedral four-coordinate geometries<br />

whereas Cu(II) complexes are typically square-planar or, in some biosystems,<br />

trigonal planar; occasionally, square planar compounds bind two additional<br />

weakly bonded axial lig<strong>and</strong>s. In metalloproteins undergoing electron<br />

transfer processes, copper experiences a wealth <strong>of</strong> slightly different coordination<br />

environments: a tetrahedral lig<strong>and</strong> arrangement usually stabilizes<br />

Cu(I) over Cu(II), decreasingtheCu(II)/Cu(I) reduction potential, whereas<br />

high reduction potentials are achieved through distortion towards trigonal<br />

planar. In general, the thermodynamic stability <strong>of</strong> a metal center in biological<br />

environments is determined not only by inherent preferences <strong>of</strong> the<br />

metal for a particular oxidation state, lig<strong>and</strong> donor set, <strong>and</strong> coordination<br />

geometry, but also by the ability <strong>of</strong> the biopolymer to control, through its<br />

three-dimensional structure, the stereochemistry <strong>and</strong> the actual nearby lig<strong>and</strong><br />

available for coordination. Non-coordinating residues also contribute to<br />

shape the local environment via hydrophilic/hydrophobic effects or steric<br />

blocking <strong>of</strong> coordination sites [17]. The complex pattern <strong>of</strong> factors occurring<br />

in biological systems make it possible to reach coordination geometries,<br />

such as trigonal planar, which can hardly be reproduced via synthetic<br />

chemistry.<br />

Two examples <strong>of</strong> copper containing metalloproteins, namely the blue copper<br />

site <strong>and</strong> the mixed-valence binuclear CuA center, can be illustrated to<br />

better underst<strong>and</strong> how Nature organizes metal complexed centers, with the<br />

aim <strong>of</strong> optimizing their properties for a specific function, in this case electron<br />

transfer [18].<br />

In the blue copper site, which occurs in the plastocyanin that couples photosystem<br />

I with photosystem II through electron transfer (ET) [19], the X-ray<br />

geometrical structure <strong>of</strong> the Cu(II) center is distorted tetrahedral <strong>and</strong> not<br />

square planar, as normally observed for cupric complexes. The coordination<br />

environment is provided by two histidine nitrogen atoms giving 2.05 ˚A long<br />

N-Cubonds,onethiolatesulfur<strong>of</strong>cysteinewithashortCu– Sbond<strong>of</strong><br />

≈ 2.1 ˚A length <strong>and</strong> one thioether methionine at a longer distance (S – Cu<br />

≈ 2.9 ˚A), Fig. 3.<br />

The unusual geometry <strong>and</strong> ligation are responsible for the unique spectroscopic<br />

features <strong>of</strong> the blue copper site. In contrast to the weak d-d tran-


76 N. Armaroli et al.<br />

Fig. 3 Thebluecoppersiteinsideplastocyanin<br />

sitions <strong>of</strong> normal tetragonal Cu(II) complexes with ε ≈ 40 M –1 cm –1 at ca.<br />

16 000 cm –1 (≈ 620 nm),thebluecoppersitehasanintenseabsorptionb<strong>and</strong><br />

at 16 000 cm –1 with ε ≈ 5000 M –1 cm –1 Fig. 4 [20]. This result is a consequence<br />

<strong>of</strong> an inversion <strong>of</strong> the lig<strong>and</strong>-to-metal charge transfer (LMCT) pattern for<br />

the blue copper site that rises from its particular lig<strong>and</strong>s distribution. As can<br />

be seen in Fig. 5 variation <strong>of</strong> the typical overlapping between the orbitals <strong>of</strong><br />

copper <strong>and</strong> those <strong>of</strong> the lig<strong>and</strong>s leads to an inversion <strong>of</strong> the relative absorption<br />

intensity, the final result is an enhancement <strong>of</strong> the absorption on the low<br />

energy side [21].<br />

Fig. 4 Absorption spectrum <strong>of</strong> the blue copper site in plastocyanin (Reprinted from [20]<br />

with permission, © (2006) American Chemical Society)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 77<br />

Fig. 5 Inverted intensity pattern <strong>of</strong> lig<strong>and</strong>-to-metal charge transfer absorption transitions<br />

for the blue copper site compared to a regular Cu(II) complex. L = generic organic lig<strong>and</strong>;<br />

S = sulfur coordinating site <strong>of</strong> a cysteine residue<br />

In photosynthesis, plastocyanin functions as an electron transfer relay<br />

between cytochrome f (inside cytochrome b6f complex) <strong>and</strong> P700 + .Cytochrome<br />

b6f complex (from photosystem II) <strong>and</strong> P700 + (from photosystem<br />

I) are both membrane-bound proteins with exposed residues on the lumenside<br />

<strong>of</strong> the thylakoid membrane <strong>of</strong> chloroplasts. Cytochrome f acts as an<br />

electron donor while P700+ accepts electrons from reduced plastocyanin<br />

(Fig. 6) [18].<br />

Fig. 6 The so-called Z-scheme <strong>of</strong> photosynthesis<br />

Plastocyanin (Cu 2+ Pc) is reduced by cytochrome f to Cu + Pc which eventually<br />

diffuses through the lumen until recognition/binding occurs with P700 + ,<br />

which oxidizes Cu + Pc back to Cu 2+ Pc. The electronic structure <strong>of</strong> the blue<br />

copper is crucial for an efficient electron transfer in which Cu(II) is reduced<br />

to Cu(I). The tetrahedral organization <strong>of</strong> the Cu(II) site minimizes the reorganizational<br />

energy λ increasing the rate <strong>of</strong> the process, according to Marcus<br />

theory [22].


78 N. Armaroli et al.<br />

In the CuA mixed-valence binuclear site (Fig. 7), present for example in<br />

the terminal aerobic respiration enzyme (cytochrome c oxidase), a similar<br />

situation occurs [23]. The efficient electron transfer is promoted by the threedimensional<br />

organization <strong>and</strong> by electronic factors. The presence <strong>of</strong> two<br />

coordinated anionic cysteine thiolates in a monomeric Cu complex, however,<br />

would severely decrease the rate <strong>of</strong> the electron transfer process by stabilizing<br />

the oxidized Cu 2+ state <strong>and</strong> making the reduction potential too negative<br />

(Fig. 7).<br />

Fig. 7 Schematic structure <strong>of</strong> the CuA mixed valence dinuclear site, present in some<br />

terminal aerobic respiration enzymes<br />

In CuA this is avoided by weakening axial bonding interactions <strong>and</strong> by<br />

delocalizing the charge over two Cu ions [24].<br />

In conclusion, Nature makes extensive use <strong>of</strong> the coordination flexibility<br />

<strong>of</strong> copper complexes <strong>and</strong> <strong>of</strong> the related tuning <strong>of</strong> electronic properties, to<br />

optimize processes <strong>of</strong> crucial importance in living organisms.<br />

1.4<br />

Cu(I) in Supramolecular Chemistry<br />

In the frame <strong>of</strong> the spectacular development <strong>of</strong> synthetic supramolecular<br />

chemistry over the last two decades [25], coordination chemistry has played<br />

a primary role [26] <strong>and</strong>, in this context, bisphenanthroline Cu(I) complexes<br />

(hereafter indicated as [Cu(NN)2] + ) have been major players [15]. Cu(I)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 79<br />

has a strong tendency to bind phenanthroline-type lig<strong>and</strong>s [27] originating<br />

a wealth <strong>of</strong> simple tetrahedral [Cu(NN)2] + complexeswithhighyields.<br />

The parent compound [Cu(phen)2] + (phen = 1,10-phenanthroline) has been<br />

scarcely studied, probably due to the lack <strong>of</strong> long-lived electronic excited<br />

states in solution, a problem that is partially avoided in the solid state, where<br />

some emission has been detected [28]. The most common [Cu(NN)2] + complexes<br />

are those 2,9 or 4,7 disubstituted phenanthrolines, due to an easier<br />

synthetic accessibility <strong>of</strong> the related lig<strong>and</strong>s.<br />

The development <strong>of</strong> sophisticated synthetic strategies, which take advantage<br />

<strong>of</strong> this metal-lig<strong>and</strong> affinity, has afforded a number <strong>of</strong> complicated<br />

molecular architectures like catenanes [29–31], rotaxanes [7, 32, 33],<br />

Fig. 8 Selected examples <strong>of</strong> multicomponent arrays containing Cu(I)-bisphenanthroline<br />

centers: A catenane, B rotaxane (R=4-[tris-(4-tert-butyl-phenyl)-methyl]-phenolato),<br />

C grid, D dendrimer (R=C8H17)


80 N. Armaroli et al.<br />

pseudo-rotaxanes [34, 35], knots [36, 37], dendrimers [38, 39], helices [40–<br />

42], polynuclear hosts [43] etc. as originally developed by Sauvage, Dietrich-<br />

Buchecker <strong>and</strong> coworkers [44]. Some <strong>of</strong> these fascinating structures are depicted<br />

in Fig. 8.<br />

Most notably, some suitably engineered supramolecular architectures<br />

based on [Cu(NN)2] + cores are able to carry out motion at the molecular level<br />

upon chemical [45], or electrochemical/photochemical stimulation [32, 46]<br />

behaving as molecular machine prototypes [47, 48]. For instance a [2]-catenate<br />

made <strong>of</strong> two different rings, one with a phenanthroline fragment the<br />

other bearing both a phenanthroline <strong>and</strong> a terpyridine unit, undergoes spontaneous<br />

<strong>and</strong> reversible molecular rearrangements (Fig. 9 steps (B) <strong>and</strong> (D))<br />

upon oxidation (step A) <strong>and</strong> subsequent reduction (step C). Rearrangements<br />

are driven by the different preferential coordination geometries <strong>of</strong> Cu(I) <strong>and</strong><br />

Cu(II), i.e. tetra- vs. pentacoordination [46].<br />

Fig. 9 Electrochemically induced molecular motions in a catenane containing a [Cu(NN)2] +<br />

center <strong>and</strong> a free tpy lig<strong>and</strong>. The spontaneous motion is driven by the different preferential<br />

coordination geometry <strong>of</strong> Cu(I) vs. Cu(II)<br />

More recently, Schmittel <strong>and</strong> coworkers have made new supramolecular<br />

systems based on [Cu(NN)2] + -type building blocks such as racks [49],<br />

grids [50], boxes [51], <strong>and</strong> macrocycles [52]. Control <strong>of</strong> the sophisticated<br />

heteroleptic architectures has been achieved by exploiting the HETPHEN<br />

(HETeroleptic bisPHENanthroline) concept [53]. This approach is based on<br />

the kinetic control <strong>of</strong> the metal complexation equilibrium <strong>and</strong>, in the target<br />

complex, the Cu(I) ionturnsouttobeboundtoasimple<strong>and</strong>abulky<br />

phenanthroline lig<strong>and</strong>; this concept is schematically illustrated in Fig. 10.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 81<br />

Fig. 10 Kinetic control in the formation <strong>of</strong> heteroleptic [CuL 1 L 2 ] + complexes (HETPHEN<br />

approach)<br />

Simple <strong>and</strong> supramolecular Cu(I)-bisphenanthroline complexes exhibit<br />

very interesting <strong>and</strong> largely tunable photophysical properties that will be illustrated<br />

in the next sections.<br />

2<br />

Cu(I)-Bisphenanthroline Complexes<br />

2.1<br />

Ground State Geometry<br />

Cu(I)-bisphenanthroline complexes generally display distorted tetrahedral<br />

geometries. The distortion from D2d symmetry can be visualized with the aid<br />

<strong>of</strong> Fig. 11 [54].<br />

θx, θy, <strong>and</strong>θz define the interlig<strong>and</strong> angles based on the CuN4 core <strong>of</strong> the<br />

complex. When a molecule possesses a perfect tetrahedral geometry (D2d), θx<br />

= θy = θz = 90 ◦ ,whereasthesquareplanargeometryD2 implies θx = θy = 90 ◦<br />

<strong>and</strong> θz = 0 ◦ .Practically,θz is the dihedral angle between the lig<strong>and</strong> planes <strong>and</strong><br />

a decrease from 90 ◦ indicates a flattening distortion <strong>of</strong> the molecule that progressively<br />

lowers its symmetry to D2.Theθx <strong>and</strong> θy values indicate the degree<br />

<strong>of</strong> “rocking” <strong>and</strong> “wagging” distortions [55].<br />

These distortions are due to intra- <strong>and</strong> intermolecular (in solid state crystals)<br />

π - stacking interactions which also cause considerable displacement<br />

from D2d symmetry. In practice, combinations <strong>of</strong> various types <strong>of</strong> distortions<br />

occur in [Cu(NN)2] + complexes <strong>and</strong> their extent is dictated by the size,<br />

chemical nature, <strong>and</strong> positions <strong>of</strong> the phenanthroline substituents. Recently,<br />

the parameter ξCD has been proposed to quantify the degree <strong>of</strong> distortion as


82 N. Armaroli et al.<br />

Fig. 11 Relative orientation <strong>of</strong> the two lig<strong>and</strong> planes (xz <strong>and</strong> yz) <strong>and</strong> <strong>of</strong> the reference θx<br />

(rocking), θy (wagging), <strong>and</strong> θz (flattening) angles. θz is on the plane <strong>of</strong> the sheet; the<br />

grey circles represent the N phenanthroline atoms; the Cu ion is centered on the origin<br />

<strong>of</strong> the reference axes; the vector ξ lies on the yz plane, the vector η is perpendicular to it.<br />

In this schematic representation the phenanthroline on the left-h<strong>and</strong> side is assumed to<br />

be placed on the plane <strong>of</strong> the sheet (yz) <strong>and</strong> that on the right-h<strong>and</strong> side on the xz plane<br />

perpendicular to it<br />

a combination <strong>of</strong> θx, θy <strong>and</strong> θz (Eq. 1) [56]:<br />

� �� �� �<br />

90◦ + θx 90◦ + θy 90◦ + θz<br />

ξCD =<br />

1803 where CD st<strong>and</strong>s for “combined distortion”.<br />

Detailed X-ray crystallographic studies have shown that the specific geometry<br />

in the solid state is dictated by packing forces <strong>and</strong> considerable variation<br />

is found for the same complex as a function <strong>of</strong> the counteranion. In the case<br />

<strong>of</strong> [Cu(1)2] (Fig. 12) θz varies from 88◦ in the tetrafluoroborate <strong>and</strong> tosylate<br />

salts to 73◦ in the picrate [57].<br />

As an example, the θz dihedral angles between the two phenanthroline<br />

planes in [Cu(1)2]PF6, [Cu(2)2]PF6 <strong>and</strong> [Cu(3)2]PF6 are 79.4◦ [57], 87.5◦ [56,<br />

58] <strong>and</strong> 79.8◦ [59], respectively, according to X-ray crystal structures; lig<strong>and</strong>s<br />

1, 2 <strong>and</strong> 3 are depicted in Fig. 12.<br />

(1)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 83<br />

Fig. 12 Lig<strong>and</strong>s 1 (2,9-dimethyl-1,10-phenanthroline), 2 (2,9-ditrifluoromethyl-1,10-phenanthroline),<br />

3 (2,9-diphenyl-1,10-phenanthroline)<br />

Substantial geometric distortion is possible with small methyl residues<br />

in [Cu(1)2]PF6, whereas bulkier trifluoromethyl groups force a more rigid<br />

quasi-tetrahedral arrangement in [Cu(2)2]PF6. Inthecase<strong>of</strong>[Cu(3)2]PF6,<br />

instead, the distortion is dictated by intramolecular π-π stacking interactions<br />

between phenyl rings <strong>and</strong> phenanthroline cores <strong>of</strong> the other lig<strong>and</strong>. The<br />

strong <strong>and</strong> sometimes hardly predictable effects <strong>of</strong> steric <strong>and</strong>/or electronic<br />

factors on the ground state geometry <strong>of</strong> [Cu(NN)2] + complexesisreflectedin<br />

the wide tuning <strong>of</strong> electronic absorption spectra.<br />

2.2<br />

Absorption Spectra<br />

In Fig. 13 are depicted the absorption spectra <strong>of</strong> three homoleptic complexes<br />

<strong>of</strong> 2,9-disubstituted phenanthrolines in CH2Cl2 solution, namely [Cu(1)2] + ,<br />

[Cu(4)2] + ,<strong>and</strong>[Cu(5)2] + . Lig<strong>and</strong>s 1, 4, <strong>and</strong>5 (Figs. 12 <strong>and</strong> 14) have alkyl- or<br />

aryl-type substituents <strong>and</strong> serve as paradigmatic cases.<br />

The ultraviolet (UV) portion <strong>of</strong> the spectra are characterized by the intense<br />

lig<strong>and</strong>-centered (LC) b<strong>and</strong>s typical <strong>of</strong> the ππ transitions <strong>of</strong> the phenanthroline<br />

lig<strong>and</strong>s [60]; the molar absorption coefficients (ε) are<strong>of</strong>theorder<strong>of</strong><br />

50 000–60 000 M –1 cm –1 . Some mixing with MLCT states cannot be excluded<br />

according to density functional theory (DFT) calculations [61]. The b<strong>and</strong>s<br />

lying in the VIS spectral region are much weaker than those in the UV (ε<br />

typically below 10 000 M –1 cm –1 ) <strong>and</strong> have been assigned to metal-to-lig<strong>and</strong><br />

charge-transfer (MLCT) electronic transitions [61–64]. These levels occur at<br />

low energy because the Cu + ion can be easily oxidized [65] <strong>and</strong> the phentype<br />

lig<strong>and</strong>s possess low-energy empty π orbitals. Direct evidence <strong>of</strong> the<br />

localized nature <strong>of</strong> the lowest-lying MLCT state <strong>of</strong> Cu(I)-bisphenanthrolines<br />

was achieved via resonance Raman [66] <strong>and</strong> transient absorption spectroscopy<br />

[67].<br />

In a number <strong>of</strong> papers McMillin et al. [68–71] <strong>and</strong> others [72–74] have presented<br />

<strong>and</strong> discussed the absorption spectra <strong>of</strong> several mononuclear Cu(I)-


84 N. Armaroli et al.<br />

Fig. 13 Absorption spectra <strong>of</strong> [Cu(1)2] + (full black line), [Cu(4)2] + (dashed line) <strong>and</strong><br />

[Cu(5)2] + (full grey line) in CH2Cl2 solution at room temperature<br />

Fig. 14 Lig<strong>and</strong>s 4 (2,9-di-p-tolyl-1,10-phenanthroline), 5 (2,9-bis(3,5-di-tert-butyl-4methoxyphenyl)-1,10-phenanthroline),<br />

6 (2,9-Diphenyl-3,4,7,8-tetramethyl-1,10-phenanthroline)<br />

bisphenanthrolines. In general, at least three MLCT b<strong>and</strong>s can be identified<br />

in the VIS spectral region [68]. They are termed b<strong>and</strong> I (above 500 nm), b<strong>and</strong><br />

II (maximum around 430–480 nm, the most prominent, attributed to S0→S3<br />

transitions [61]), <strong>and</strong> b<strong>and</strong> III (390–420 nm, <strong>of</strong>ten hidden by the onset <strong>of</strong><br />

b<strong>and</strong> II). The envelope <strong>of</strong> such MLCT b<strong>and</strong>s defines the shape <strong>of</strong> the VIS absorption<br />

spectrum (400–700 nm). Spectral intensities are strictly related to<br />

the symmetry <strong>of</strong> the complex (D2d vs. D2, Fig.15)that,inturn,isaffectedby<br />

the distortion from the tetrahedral geometry (see above).


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 85<br />

Fig. 15 Schematic orbital splitting diagrams for [Cu(NN)2] + . Left, D2d symmetry; right,<br />

D2 symmetry. Grey arrows represent the transitions leading to b<strong>and</strong>s I (— —), b<strong>and</strong> II<br />

(···), <strong>and</strong> b<strong>and</strong> III (– – –)<br />

A simple picture able to rationalize the MLCT absorption patterns <strong>of</strong><br />

these complexes in solution is difficult, nevertheless some general trends have<br />

been found by analyzing the MLCT absorption maxima <strong>and</strong> molar extinction<br />

coefficients <strong>of</strong> series <strong>of</strong> [Cu(NN)2] + compounds <strong>and</strong> have been reported<br />

elsewhere [15].<br />

The spectrum <strong>of</strong> [Cu(4)2] + in Fig. 13 exhibits a well-established fingerprint<br />

for complexes with 2,9-aryl substitution <strong>of</strong> the phenanthroline ring, i.e.<br />

a pronounced low-energy shoulder extending down to 650 nm (b<strong>and</strong> I, see<br />

above), which is practically absent in the case <strong>of</strong> [Cu(1)2] + .Thispatternisrelated<br />

to the above mentioned intramolecular π-stacking interactions, which<br />

make the transition corresponding to b<strong>and</strong> I more permitted. The absorption<br />

spectrum <strong>of</strong> [Cu(5)2] + is somewhat different if compared to both [Cu(4)2] +<br />

<strong>and</strong> [Cu(1)2] + . The low energy b<strong>and</strong> is wider <strong>and</strong> more intense <strong>and</strong> the peak<br />

around 440, which is typically the most intense in [Cu(NN)2] + complexes, appears<br />

as just a weak shoulder [74]. This trend is due to the presence <strong>of</strong> the<br />

cumbersome tert-butyl groups on the phenyl residues, that limit π-π stacking<br />

interactions <strong>and</strong> make the structure <strong>of</strong> the complex particularly rigid.<br />

The MLCT absorption pr<strong>of</strong>ile <strong>of</strong> [Cu(5)2] + is found to be similar to that <strong>of</strong><br />

aCu(I)-catenate complex made <strong>of</strong> two interlocked 27-membered rings containing<br />

a phenanthroline moiety <strong>and</strong> exhibiting a very rigid coordination<br />

environment [75]. The structural peculiarity <strong>of</strong> [Cu(5)2] + , as revealed by the<br />

absorption spectrum, is in accord with its extraordinary kinetic inertness towards<br />

demetallation [74].<br />

Interestingly, it has been found that the complex [Cu(6)2] + (Fig. 14) exhibits<br />

a very weak b<strong>and</strong> above 500 nm despite the presence <strong>of</strong> phenyl rings in<br />

the 2 <strong>and</strong> 9 positions [71]. Indeed this confirms the above described model:<br />

the methyl groups in the 3 <strong>and</strong> 8 position contrast the flattening distortion,


86 N. Armaroli et al.<br />

leading to a coordination geometry (<strong>and</strong> a spectrum) very similar to that <strong>of</strong><br />

the much simpler [Cu(1)2] + complex.<br />

Some Cu(I)-bisphenanthroline compounds have also been utilized as receptors<br />

for dicarboxylic acids [76] <strong>and</strong> spherical inorganic anions [77].<br />

The recognition motif is established by suitably functionalizing one phenyl<br />

residue in the 2 or 9 position <strong>of</strong> the phenanthroline chelating units, which<br />

are made able to pinch the pertinent substrate. The formation <strong>of</strong> the supramolecular<br />

adducts is monitored through the substantial changes <strong>of</strong> the MLCT<br />

absorption b<strong>and</strong>s that occur as a consequence <strong>of</strong> the modification <strong>of</strong> the coordination<br />

geometry <strong>and</strong> related symmetry [76].<br />

2.3<br />

Excited State Distortion: Pulsed X-ray <strong>and</strong> Transient Absorption Spectroscopy<br />

Upon light excitation <strong>of</strong> [Cu(NN)2] + complexes the lowest MLCT excited state<br />

is populated, thus the metal center changes its formal oxidation state from<br />

Cu(I) to Cu(II) [15]. The Cu(I) MLCT excited complex undergoes further flattening<br />

compared to its ground state <strong>and</strong> assumes a geometry similar to that<br />

<strong>of</strong> ground state Cu(II)-bisphenanthroline complexes [59]. In this excited state<br />

flattened tetrahedral structure a fifth coordination site is made available for<br />

the Cu(II) d 9 ion, that can be filled by nucleophilic species such as solvent<br />

molecules <strong>and</strong> counterions. The intermediate species thus obtained is termed<br />

“pentacoordinated exciplex”. The process, which is schematically depicted in<br />

Fig. 16, had been proposed by McMillin <strong>and</strong> coworkers nearly 20 years ago<br />

on the basis <strong>of</strong> classical photochemical experiments on series <strong>of</strong> [Cu(NN)2] +<br />

complexes with increasingly nucleophilic counteranions [78].<br />

Fig. 16 Flattening distortion <strong>and</strong> subsequent nucleophilic attack by solvent, counterion,<br />

or other molecules following light excitation in Cu(I)-phenanthrolines. The size (<strong>and</strong> position)<br />

<strong>of</strong> the R substituents is <strong>of</strong> paramount importance in determining both the extent <strong>of</strong><br />

the distortion <strong>and</strong> the protection <strong>of</strong> the newly formed Cu(II) ion from nucleophiles<br />

In recent years this hypothesis has been nicely confirmed thanks to the<br />

development <strong>of</strong> light-initiated time-resolved X-ray absorption spectroscopy


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 87<br />

(LITR-XAS) [79]. This pump-<strong>and</strong>-probe technique allows one to catch the<br />

transient oxidation state <strong>of</strong> a metal atom as well as its surrounding structures<br />

following photoexcitation via an ultrafast laser source; accordingly, it is particularly<br />

suited to investigating the process depicted in Fig. 16. Practically, the<br />

information obtained via LITR-XAS is a sort <strong>of</strong> snapshot <strong>of</strong> electronic excited<br />

states in disordered media (e.g. solution), which are generated via a UV-VIS<br />

femtosecond pump laser <strong>and</strong> subsequently probed with a 30–100 ps intense<br />

X-ray pulse produced by 3rd generation large synchrotron facilities [80].<br />

This technique has been applied in solution studies to [Cu(1)2] + <strong>and</strong> unambiguously<br />

confirmed that, in the thermally equilibrated MLCT excited<br />

state, the copper ion is pentacoordinated both in poorly donor (toluene) [81]<br />

or highly donor (CH3CN) [82] solvents; in addition, the copper ion has<br />

thesameoxidationstateasthecorrespondinggroundstateCu(II) complex<br />

in both cases. Analogous investigations have been carried out also on<br />

Cu(I) complexes as solid crystals (“photocrystallography”) [55, 83]. LITR-<br />

XAS studies <strong>of</strong> [Cu(1)2] + in solution have been complemented by optical<br />

time-resolved spectroscopy, which evidenced spectroscopic features in the ps<br />

timescale, associated to excited state structural rearrangements, possibly flattening<br />

distortion [82].<br />

We have also carried out femtosecond transient absorption studies on<br />

[Cu(7)2] + <strong>and</strong> [Cu(8)2] + in CH2Cl2 (Fig. 17) [84]. These complexes are characterized<br />

by alkyl- <strong>and</strong> more cumbersome phenyl-residues in the 2 <strong>and</strong> 9<br />

position <strong>of</strong> the phenanthroline lig<strong>and</strong>, which imparts rather different photophysical<br />

properties (i.e. shape <strong>of</strong> UV-VIS absorption, luminescence spectra,<br />

excited state lifetime) [85]. Despite this diversity, femtosecond transient absorption<br />

spectra have revealed a dynamic process lasting 15 ps in both cases<br />

Fig. 18.<br />

Fig. 17 Lig<strong>and</strong>s 7 (2,9-bis(4-n-butylphenyl)-1,10-phenanthroline) <strong>and</strong> 8 (2,9-di-n-hexyl-<br />

1,10-phenanthroline)<br />

Specific assignment <strong>of</strong> the observed spectral variation to (i) flattening distortion<br />

or (ii) extra lig<strong>and</strong> pick-up, two processes that might also occur sim-


88 N. Armaroli et al.<br />

Fig. 18 Transient absorption spectral changes observed for [Cu(8)2] + in CH2Cl2 at λexs<br />

= 400 nm (150 fs Ti:Sapphire laser pulse). The spectra were recorded at delays <strong>of</strong> 2, 5, 10,<br />

25, 100, 1000 ps following the excitation pulse. In the inset are depicted spectral decays at<br />

two selected wavelengths, λobs = 520 (full circles) <strong>and</strong> 590 (half-empty circles)<br />

ultaneously, is not straightforward. Access to the fifth coordinating position<br />

is likely to be less favored for the more congested phenyl–phenanthroline<br />

complex [Cu(7)2] + . Hence the identical rate constant observed for the two<br />

compounds in CH2Cl2, aswellasinCH3CN for another [Cu(NN)2] + complex<br />

[82], is likely to be associated with the flattening distortion which is<br />

expected to be less solvent- <strong>and</strong> lig<strong>and</strong>-dependent than picking up an external<br />

unit for coordination expansion. Transient absorption studies leading<br />

to similar results have also been carried out with monophenanthroline complexes<br />

[86].<br />

2.4<br />

Emissive Excited State(s) <strong>and</strong> Luminescence Spectra<br />

The first report on [Cu(NN)2] + luminescence in fluid solution dates back to<br />

1980, when it was shown that, upon excitation into the MLCT b<strong>and</strong> region,<br />

[Cu(1)2] + exhibits a luminescence spectrum peaking around 700 nm <strong>and</strong> an<br />

excited state lifetime <strong>of</strong> 54 ns in air-equilibrated CH2Cl2 [87]. Luminescence<br />

from [Cu(NN)2] + complexes is observed in poorly electron donor solvents,<br />

typically CH2Cl2. At room temperature, the emission b<strong>and</strong> is wide <strong>and</strong> exhibit<br />

λmax peaking between 680 <strong>and</strong> 740 nm, with rather low quantum yield<br />

(Φem 10 –3 –10 –4 ) [15]. Excited state lifetimes in CH2Cl2 solution are strongly<br />

dependent on the degree <strong>of</strong> excited state distortion <strong>and</strong> the protection to-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 89<br />

wards exciplex quenching. It has been hypothesized that emission stems from<br />

the pentacoordinated exciplex itself that deactivates to a pentacoordinated<br />

groundstatespecieswhicheventuallylosesthe“fifth”nucleophiliclig<strong>and</strong>to<br />

regenerate the initial ground state pseudotetrahedral complex [81]. In recent<br />

years it has also been evidenced that some [Cu(NN)2] + complexes exhibit<br />

a prompt luminescence signal with a lifetime <strong>of</strong> the order <strong>of</strong> 13–16 ps, which<br />

is attributed to deactivation <strong>of</strong> 1 MLCT [88], whereas the long-lived component<br />

(above 50 ns) would be phosphorescence from 3 MLCT, which borrows<br />

intensity from upper lying singlet levels [89].<br />

The shortest <strong>and</strong> longest values reported to date (oxygen-free CH2Cl2<br />

solution, longer-lived component) are 80 [68] <strong>and</strong> 930 ns [71] <strong>and</strong> refer to homoleptic<br />

[Cu(NN)2] + complexes <strong>of</strong> the two phenanthroline lig<strong>and</strong>s depicted<br />

in Fig. 19.<br />

Fig. 19 Lig<strong>and</strong> 9 (2,9-dimethyl-4,7-diphenyl-1,10-phenanthroline) <strong>and</strong> 10 (2,9-di-n-butyl-<br />

3,4,7,8-tetramethyl-1,10-phenanthroline). Lifetime values <strong>of</strong> their [Cu(NN)2] + complexes<br />

differ by more than one order <strong>of</strong> magnitude<br />

The value for [Cu(9)2] + is very similar to that <strong>of</strong> [Cu(1)2] + under the<br />

same conditions (80 ns), suggesting that electronic delocalization <strong>of</strong> the lig<strong>and</strong>s<br />

[73], very strong for 9, is less important than steric factors in determining<br />

the lifetimes value. Notably, substitution on the 3 <strong>and</strong> 8 position<br />

with a simple methyl residue in 10 is particularly effective to limit exciplex<br />

quenching <strong>and</strong> yields a lifetime <strong>of</strong> almost 1 µs. In general, the large majority<br />

<strong>of</strong> [Cu(NN)2] + homoleptic complexes exhibit excited state lifetimes in the<br />

range 80–350 ns in oxygen-free solution [15]. A longer lifetime (730 ns, Φem<br />

= 0.01, oxygen-free CH2Cl2) has been found with the suitably designed heteroleptic<br />

complex [Cu(1)(11)] + [90], Fig. 20, in which excited state distortion<br />

is strongly limited by the cumbersome tert-butyl substituent. Interestingly,<br />

steric contraints make the formation <strong>of</strong> the homoleptic analogue [Cu(11)2] +<br />

very difficult.<br />

The picture describing the luminescent excited states <strong>of</strong> Cu(I)-bisphenanthrolinesisnotstraightforwardalthoughParkeretal.,inlight<strong>of</strong>theob-


90 N. Armaroli et al.<br />

Fig. 20 Lig<strong>and</strong> 11 2,9-di-tert-butyl-1,10-phenanthroline<br />

served large Stokes shift (over 5000 cm –1 )hadattributedittothelowest<br />

3 MLCT excited state [91], likewise the popular family <strong>of</strong> octahedral Ru(II)polypyridines<br />

[6]. McMillin <strong>and</strong> coworkers, instead, suggested that emission<br />

<strong>of</strong> [Cu(NN)2] + compoundsarisesfromtwoMLCTexcitedstatesinthermal<br />

equilibrium, i.e. a singlet ( 1 MLCT)<strong>and</strong>atriplet( 3 MLCT) [92]. The energy<br />

gap between these states was found to be about 1500–2000 cm –1 <strong>and</strong>, at room<br />

temperature, the population <strong>of</strong> the lower lying 3 MLCT level exceeds that <strong>of</strong><br />

1 MLCT. At 77 K where the excited state population is largely frozen in the<br />

triplet, the emission b<strong>and</strong> is red-shifted <strong>and</strong> much weaker compared to room<br />

temperature, a rather unusual trend. Recent studies have confirmed <strong>and</strong> refined<br />

this rationale [81, 85, 89].<br />

A few years ago our group discussed detailed temperature-dependent luminescence<br />

studies <strong>of</strong> a series <strong>of</strong> [Cu(NN)2] + complexes <strong>of</strong> 2,9-disubstituted<br />

phenanthroline lig<strong>and</strong>s [85]. The above-described two-level model, which implies<br />

red-shift <strong>and</strong> intensity decrease <strong>of</strong> the emission b<strong>and</strong> upon temperature<br />

lowering,isalwaysobeyedexceptwhenlongalkylchainsareutilizedassubstituents<br />

<strong>of</strong> the phenanthroline chelating agent. In this case the “regular”<br />

trend is obeyed only until the matrix remains fluid (T>150 K) but, when the<br />

matrix becomes rigid (T


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 91<br />

Fig. 21 Temperature dependence <strong>of</strong> the luminescence spectra <strong>of</strong> [Cu(7)2] + (top) <strong>and</strong><br />

[Cu(8)2] + (bottom) in CH2Cl2 MeOH 1:1 (v/v). In the fluid domain (up to 170 K) emission<br />

intensity decrease <strong>and</strong> spectral red-shifting is observed by lowering temperature in<br />

both cases. By contrast when the solvent matrix becomes rigid (around 120 K), the two<br />

compounds behave differently. For the 2,9-dialkylphenanthroline complex (bottom panel)<br />

a complete reversal <strong>of</strong> the previous trend is observed with intensity recovery <strong>and</strong> blue<br />

shift. At 96 K a very strong luminescence b<strong>and</strong> is recorded


92 N. Armaroli et al.<br />

Fig. 22 Lig<strong>and</strong> 12, 2-(3,5-di-tert-butyl-4-methoxyphenyl)-9-(2,4,6-trimethylphenyl)-1,10phenanthroline<br />

(Fig. 22). This finding gives further support to the notion that steric factors<br />

are by far more important than electronic factors.<br />

The room temperature lifetime <strong>of</strong> [Cu(12)2] + in oxygen free solution (285<br />

ns) is scarcely affected under air-equilibrated conditions (266 ns) <strong>and</strong> is remarkably<br />

high also in CH3OH (182 ns), a solvent where most [Cu(NN)2] +<br />

do not show any luminescence. This rather unusual solvent insensitivity <strong>of</strong><br />

the lifetime is observed also for [Cu(5)2] + (Fig. 14) suggesting that tert-butyl<br />

groups are very effective in preventing the formation <strong>of</strong> the pentacoordinated<br />

exciplex at room temperature [74]. On the other h<strong>and</strong>, in striking contrast<br />

with [Cu(12)2] + , [Cu(5)2] + is virtually non-luminescent in 77 K rigid matrix,<br />

highlighting once again the subtle factors affecting the molecular structure<br />

<strong>and</strong>, accordingly, the emission performance. Probably, in [Cu(12)2] + ,thekey<br />

structural features causing good luminescence performance are the methyl<br />

residues on the 2 <strong>and</strong> 6 position <strong>of</strong> a phenyl substituent (Fig. 22).<br />

Solid state measurements confirm the strong dependence <strong>of</strong> the excited<br />

state lifetimes <strong>of</strong> [Cu(NN)2] + complexes on geometric distortions. In particular<br />

it has been found that there is a good linear correlation between the<br />

ξCD parameter (see Sect. 2.1, Eq. 1) <strong>and</strong> the measured lifetime both at room<br />

temperature <strong>and</strong> at 17 K: the smaller the distortion from the ideal pseudotetrahedral<br />

geometry (high ξCD parameter), the longer the lifetime [56].<br />

2.5<br />

Photoinduced Processes in Multicomponent Systems Based on<br />

[Cu(NN)2] + Complexes<br />

Templated synthesis <strong>of</strong> Cu(I)-bisphenanthrolines has prompted the design<br />

<strong>and</strong> construction <strong>of</strong> sophisticated multicomponent architectures comprising<br />

one or more [Cu(NN)2] + centers in t<strong>and</strong>em with other chromophores, typ-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 93<br />

ically porphyrins <strong>and</strong> fullerenes. Intercomponent light-induced energy- <strong>and</strong><br />

electron-transfer processes have been investigated in [2]- [93, 94] <strong>and</strong> [3]catenanes<br />

[29, 95, 96], dinuclear knots [97], fullerene- [98] <strong>and</strong> porphyrinstoppered<br />

rotaxanes [99–108], dendrimers with [Cu(NN)2] + cores [38], <strong>and</strong><br />

helicates with Cu(I)-complexed cores <strong>and</strong> peripheral methano- [109] or<br />

bismethano-fullerenes [41, 109, 110]. The latter, which constitute the most recently<br />

investigated systems, are depicted in Fig. 23.<br />

Fig. 23 Fullerohelicates based on a dinuclear [Cu(NN)2] + complex (Z=C8H17,R=C12H25).<br />

The peripheral moieties are different in the two cases, namely bismethano (left) vs.<br />

methan<strong>of</strong>ullerenes (right)<br />

Upon excitation <strong>of</strong> the Cu(I)-complexed moiety <strong>and</strong> population <strong>of</strong> the related<br />

MLCT level, electron transfer to the fullerene subunit is observed for<br />

both compounds shown in Fig. 23. By contrast, although the same process<br />

is thermodynamically allowed also by populating the fullerene lowest singlet<br />

state, it is observed only in the case <strong>of</strong> the methan<strong>of</strong>ullerene system. This<br />

is related to the inherently different electronic structure <strong>of</strong> the two fullerene<br />

derivatives. By means <strong>of</strong> an analysis <strong>of</strong> their fluorescence spectra, which<br />

are substantially different, it was possible to conclude that the singlet excited<br />

state <strong>of</strong> methan<strong>of</strong>ullerenes is more prone to undergo electron transfer than<br />

that <strong>of</strong> bismethan<strong>of</strong>ullerenes, thanks to the associated smaller internal reorganization<br />

energy [109]. In addition, methan<strong>of</strong>ullerenes are slightly easier to<br />

reduce than bismethan<strong>of</strong>ullerenes, giving also a thermodynamic advantage<br />

for electron transfer in multicomponent arrays containing the mon<strong>of</strong>unc-


94 N. Armaroli et al.<br />

tionalized derivative. The combined effect <strong>of</strong> these two factors (kinetic <strong>and</strong><br />

thermodynamic) can explain the different <strong>and</strong> unexpected trend in photoprocesses<br />

<strong>of</strong> multicomponent arrays containing Cu(I)-phenanthrolines linked to<br />

methan<strong>of</strong>ullerenes vs. bismethan<strong>of</strong>ullerenes, which has been found in a variety<br />

<strong>of</strong> molecular architectures such as dendrimers [38], rotaxanes [98] <strong>and</strong><br />

s<strong>and</strong>wich-type dyads [110].<br />

Exhaustive review articles presenting photophysical investigations on<br />

fullerene- <strong>and</strong> porphyrin-type arrays built-up around [Cu(NN)2] + centers<br />

have been published recently <strong>and</strong> we suggest the reader refers to these papers<br />

for a comprehensive <strong>and</strong> updated overview on this topic [15, 25, 111, 112].<br />

2.6<br />

Bimolecular Quenching Processes<br />

Excited state electrochemical potentials can be obtained from the ground<br />

state monoelectronic electrochemical potentials <strong>and</strong> the spectroscopic energy<br />

(E ◦◦ in eV units, to be considered divided by a unitary charge) related to the<br />

involved transition, according to Eqs 2 <strong>and</strong> 3 [6]:<br />

E(A + / ∗ A)=E(A + /A)–E ◦◦<br />

E( ∗ A/A – )=E(A/A – )+E ◦◦<br />

Hence the variation <strong>of</strong> the electron-donating or accepting capability <strong>of</strong> a given<br />

molecule A, upon light excitation, can be easily assessed. In Eqs 2 <strong>and</strong> 3: ∗ A<br />

denotes the lowest-lying electronically excited state <strong>of</strong> A <strong>and</strong> its spectroscopic<br />

energy (E ◦◦ ) can be estimated from the onset <strong>of</strong> emission spectra [6].<br />

Oxidation from Cu(I) to Cu(II) is easily accomplished <strong>and</strong> the MLCT<br />

excited states <strong>of</strong> Cu(I)-bisphenanthrolines are, therefore, potent reductants.<br />

For example [Cu(3)2] + is a more powerful reductant than the very popular<br />

photosensitizer [Ru(bpy)3] 2+ (A + /A = – 1.11 <strong>and</strong> – 0.85 V, respectively)<br />

owing to its more favorable ground state 2+/+ potential (+ 0.69 vs. + 1.27 V),<br />

that largely compensates the lower content <strong>of</strong> excited state energy (1.80 vs.<br />

2.12 eV) [15]. By contrast reduction <strong>of</strong> Cu(I)-bisphenanthrolines is strongly<br />

disfavored <strong>and</strong> they are mild excited state oxidants; accordingly, only a few examples<br />

<strong>of</strong> reductive quenching <strong>of</strong> [Cu(NN)2] + complexes are reported in the<br />

literature, with ferrocenes as donors [113, 114].<br />

Oxidative quenching <strong>of</strong> [Cu(NN)2] + ’s by Co(III) <strong>and</strong> Cr(III) complexes as<br />

well as nitroaromatic compounds <strong>and</strong> viologens has been reported <strong>and</strong> comprehensively<br />

reviewed [115]. Some attempts to sensitize wide b<strong>and</strong>-gap semiconductors<br />

with Cu(I) complexes were also carried out [115] but so far they<br />

do not seem to be competitive in terms <strong>of</strong> stability <strong>and</strong> efficiency with those<br />

based on Ru(II) complexes [12]. Energy transfer quenching to molecules<br />

possessing low-lying triplets such as anthracene has been demonstrated via<br />

transient absorption spectroscopy [116, 117], whereas oxygen quenching,<br />

(2)<br />

(3)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 95<br />

which in principle can occur both via energy- <strong>and</strong> electron transfer, was evidenced<br />

by monitoring sensitized singlet oxygen luminescence in the NIR<br />

region [29, 36]. Optically pure dicopper trefoil knots with [Cu(NN)2] + -type<br />

cores have been reported to quench the emission <strong>of</strong> the Λ or ∆ forms <strong>of</strong><br />

Tb(III) <strong>and</strong> Eu(III) complexes, a very rare example <strong>of</strong> enantioselective luminescence<br />

quenching [118].<br />

[Cu(NN)2] + complexes have also been used as substrates for DNA binding,<br />

trying to take advantage <strong>of</strong> the sensitivity <strong>of</strong> the luminescence <strong>of</strong> Cu(I)phenanthrolines<br />

to the local environment [63]. The structure <strong>of</strong> the associates<br />

has not been clarified: both electrostatic binding <strong>and</strong> intercalation <strong>of</strong> the aromatic<br />

lig<strong>and</strong>s between adjacent bases are possible. Cu(I)-porphyrins seem to<br />

be more promising substrates for DNA [63].<br />

3<br />

Heteroleptic Diimine/Diphosphine [Cu(NN)(PP)] + Complexes<br />

3.1<br />

Photophysical Properties<br />

Heteroleptic Cu(I) complexes containing both N- <strong>and</strong> P-coordinating lig<strong>and</strong>s,<br />

[Cu(NN)(PP)] + , have been studied since the late 1970s [119]. The replacement<br />

<strong>of</strong> one N-N lig<strong>and</strong> with a P-P unit is <strong>of</strong>ten aimed at improving the<br />

emission properties. Accordingly, the relentless quest for highly performing<br />

luminescent metal complexes [7] has sparked revived interest in these compounds<br />

in recent years [120–122].<br />

The absorption <strong>and</strong> luminescence spectrum <strong>of</strong> [Cu(dbp)(POP)] + (dbp =<br />

2,9-butyl-1,10-phenanthroline <strong>and</strong> POP = bis[2-(diphenylphosphino)phenyl]<br />

ether) is reported in Fig. 24, as a representative example for this class <strong>of</strong> compounds<br />

[123]. Substantial blue-shifts <strong>of</strong> the lower-energy b<strong>and</strong>s are observed<br />

compared to typical spectra <strong>of</strong> [Cu(NN)2] + compounds (see Sect. 2).<br />

UV spectral features above 350 nm are due to lig<strong>and</strong>-centered transitions<br />

whereas those in the 350–450 nm window are attributed to MLCT<br />

levels. [Cu(NN)(PP)] + complexes are subject to dramatic oxygen quenching,<br />

as deduced from the strong difference in excited state lifetimes passing<br />

from air-equilibrated to oxygen-free CH2Cl2 solution, 250 ns <strong>and</strong> 17 600 ns<br />

in the case <strong>of</strong> [Cu(dbp)(POP)] + [123]. The character <strong>of</strong> the emitting state<br />

in [Cu(NN)(PP)] + complexes has been discussed since their first characterization<br />

[119] <strong>and</strong> now its MLCT nature is established experimentally <strong>and</strong><br />

theoretically [120, 124, 125]. The electron-withdrawing effect <strong>of</strong> the P–P unit<br />

on the metal center tends to disfavor the Cu(I)→N–N electron donation, as<br />

also reflected by the higher oxidation potential <strong>of</strong> the Cu(I) center compared<br />

to [Cu(NN)2] + compounds [126], leading to a blue shift <strong>of</strong> MLCT transitions.<br />

This, according to the energy gap law [127], explains the emission enhance-


96 N. Armaroli et al.<br />

Fig. 24 Absorption <strong>and</strong> (inset) emission spectra <strong>of</strong> [Cu(dbp)(POP)] + in CH2Cl2<br />

ment <strong>of</strong> [Cu(NN)(PP)] + , that typically falls in the green spectral window,<br />

compared to weaker red-emitting [Cu(NN)2] + complexes.<br />

The luminescence efficiency <strong>of</strong> MLCT excited states in [Cu(NN)(PP)] +<br />

compounds is strongly solvent- <strong>and</strong> oxygen-dependent because it can be<br />

decreased by exciplex quenching [128, 129], in line with what is observed<br />

for the [Cu(NN)2] + analogues (see above). Therefore, the geometry <strong>of</strong><br />

[Cu(NN)(PP)] + complexes plays a central role in addressing the extent <strong>of</strong><br />

luminescence efficiency, even though this is hard to predict a priori.<br />

A variety <strong>of</strong> bidentate phosphine lig<strong>and</strong>s has been prepared to coordinate<br />

Cu(I) in t<strong>and</strong>em with phenanthroline-type units: bis[2-(diphenylphosphino)<br />

phenyl]ether (POP), triphenylphosphine (PPh3), bis(diphenylphosphino)<br />

ethane (dppe), <strong>and</strong> bis(diphenyl-phosphino)methane (dppm), represent<br />

some recent examples [120, 122, 123, 130, 131], Fig. 25.<br />

Among them, the family <strong>of</strong> mononuclear [Cu(phen)(POP)] + complexes<br />

proposed by McMillin (see Fig. 25 for the PP-type lig<strong>and</strong>s), where phen indicates<br />

a variably substituted 1,10 phenanthroline, shows an impressive emission<br />

efficiency compared to [Cu(NN)2] + compounds [120]. Especially on<br />

passing from pristine phenanthroline to dimethyl- or diphenyl-substituted<br />

analogues, <strong>and</strong> thanks to the efficient steric <strong>and</strong> electron-withdrawing effects<br />

<strong>of</strong> the POP lig<strong>and</strong>, remarkable emission quantum yields (Φem ∼ 0.15<br />

in CH2Cl2 oxygen-free solution) <strong>and</strong> long lifetimes (∼ 15 µs) have been<br />

measured. On the contrary, the replacement <strong>of</strong> the POP lig<strong>and</strong> with two<br />

PPh3 units, gives less remarkable results due to the lower geometric rigidity<br />

which leads to weak <strong>and</strong> red-shifted emissions comparable to those <strong>of</strong>


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 97<br />

Fig. 25 Some lig<strong>and</strong>s typically used as P–P units in [Cu(NN)(PP)] + complexes<br />

bis-phenanthroline-type complexes. Importantly, the P–Cu–P angle decreases<br />

from 122.7 ◦ in [Cu(dmp)(PPh3)2] + to 116.4 ◦ in [Cu(dmp)(POP)] + also allowing<br />

easier access for exciplex quenching over the fifth coordination position<br />

in the former. This example highlights the importance <strong>of</strong> having both conditions<br />

(i.e. steric protection <strong>and</strong> increased electron-withdrawing character<br />

<strong>of</strong> the P–P lig<strong>and</strong>) simultaneously satisfied for optimized photoluminescence<br />

performance <strong>of</strong> [Cu(NN)(PP)] + compounds.<br />

The importance <strong>of</strong> the choice <strong>of</strong> the P–P lig<strong>and</strong> for the coordination <strong>of</strong><br />

the metal ion is evidenced also by the systems recently investigated by Wang<br />

et al. [121], in which lig<strong>and</strong>s other than phenanthroline have been utilized<br />

(Fig. 26).<br />

By keeping the N–N lig<strong>and</strong> unchanged, the luminescence properties <strong>of</strong> the<br />

complexes (solid matrix, RT) increase on passing from dppe to POP to, sur-<br />

Fig. 26 General structure <strong>of</strong> [Cu(ppb)(P)2] complexes (pbb = 2-(2′-pyridyl)benzimidazolylbenzene)


98 N. Armaroli et al.<br />

prisingly, PPh3. Although POP provides the best emission performance when<br />

combined with phenanthroline lig<strong>and</strong>s, a better result is found here for PPh3.<br />

This shows that subtle <strong>and</strong> combined steric <strong>and</strong> electronic effects <strong>of</strong> both P–P<br />

<strong>and</strong> N–N lig<strong>and</strong>s are crucial for an enhanced light output, highlighting that<br />

general rules able to predict the photophysical behavior <strong>of</strong> [Cu(NN)(PP)] +<br />

complexes are not easy to draw. From an electronic point <strong>of</strong> view, both POP<br />

<strong>and</strong> PPh3 units promote the usual blue shift <strong>of</strong> the MLCT state compared to<br />

[Cu(NN)2] + compounds, as predictable by the substantially higher oxidation<br />

potential <strong>of</strong> the Cu(I) center <strong>of</strong> heteroleptic [Cu(NN)(PP)] + complexes.<br />

Dinuclear Cu(I) complexes have also been synthesized <strong>and</strong> investigated,<br />

two <strong>of</strong> them (A <strong>and</strong> B) are depicted in Fig. 27. Despite the presence <strong>of</strong> a P–<br />

P-type lig<strong>and</strong>, complex A shows a luminescence b<strong>and</strong> peaked at 700 nm with<br />

a lifetime <strong>of</strong> 320 ns in the solid state [132]. The X-ray crystal structure indicates<br />

a distorted tetrahedral geometry which, combined to the scarcely<br />

protective 2,5-bppz N–N lig<strong>and</strong> (2,5-bis(2-pyridil)pyrazine) leads to a weakly<br />

red-emitting compound. By changing the N-N lig<strong>and</strong> (Fig. 27, B), a stronger<br />

Fig. 27 Chemical structures <strong>of</strong> heteroleptic Cu(I) complexes A <strong>and</strong> B


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 99<br />

green emission (λmax = 550 nm) is detected with a quantum yield <strong>of</strong> 0.17 at<br />

77 KinCH2Cl2 matrix [133]. For complex B, althoughthegeometricstructure<br />

is similar to A, the oxidation potentials <strong>of</strong> the N–N lig<strong>and</strong> are substantially<br />

greater, pushing the MLCT levels at higher energy.<br />

3.2<br />

OLED <strong>and</strong> LEC Devices<br />

The outst<strong>and</strong>ing photophysical performances <strong>of</strong> some [Cu(NN)(PP)] + complexes<br />

make them potentially attractive for optoelectronic devices requiring<br />

highly luminescent materials [134–137]. This interest is also related to the<br />

lower cost <strong>and</strong> higher relative abundance <strong>of</strong> copper compared to more classical<br />

emitting metals such as europium or iridium. Some research groups have<br />

recently fabricated OLED devices with [Cu(NN)(PP)] + complexes [138]. It has<br />

been shown that they can be pr<strong>of</strong>itably used as electrophosphorescent emitters<br />

<strong>and</strong> provide device efficiency comparable to that <strong>of</strong> Ir(III) complexes in<br />

similar device structures (11.0 cd/A at1.0 mA/cm 2 , 23% wtCu(I)-complex<br />

dispersed in PVK matrix). Also Li et al. have obtained a highly efficient electrophosphorescent<br />

OLED with the complex reported in Fig. 28 [139].<br />

Fig. 28 The complex [Cu(Dicnq)(POP)] + BF4 used by Li et al. to make a highly efficient<br />

electroluminescent OLED device (Dicnq = 6,7-Dicyanodipyrido[2,2-d:2 ′ ,3 ′ -f ]quinoxaline)<br />

The performances <strong>of</strong> OLEDs fabricated by the vacuum vapor deposition<br />

technique with this complex are among the best reported for devices incorporating<br />

Cu(I) complexes as emitters. A low turn-on voltage <strong>of</strong> 4 V, a maximum<br />

current efficiency up to 11.3 cd/A, <strong>and</strong> a peak brightness <strong>of</strong> 2322 cd/m 2 have<br />

been achieved.


100 N. Armaroli et al.<br />

A different type <strong>of</strong> electroluminescent device is a light-emitting electrochemical<br />

cell (LEC). LECs are substantially different from OLEDs due to<br />

the fact that mobile ions in the electroluminescent layer drift towards the<br />

electrodes when a voltage is applied over the device, thereby facilitating<br />

charge-carrier injection from the electrodes. This results in two important<br />

advantages compared to traditional OLEDs: (i) thick electroactive layers can<br />

be used without severe voltage penalties <strong>and</strong> shorts can be eliminated even for<br />

large-area pixels; (ii) matching <strong>of</strong> the work function <strong>of</strong> the electrodes with the<br />

energy levels <strong>of</strong> the electroluminescent material is not required.<br />

We have recently described novel Cu(I) complexes with excellent PL performance<br />

(Q.Y. up to 0.28 in oxygen-free CH2Cl2) <strong>and</strong>thefirstLECdevice<br />

made with a Cu(I) complex, Fig. 29 [123].<br />

Fig. 29 Chemical structure <strong>of</strong> the complex used to make the first LEC device based<br />

on a Cu(I)-complex, R = n-butyl. A schematic representation <strong>of</strong> the device structure is<br />

also depicted; in the electroluminescent layer the complex is dispersed in a polymethylmetacrylate<br />

matrix (PMMA)<br />

The device efficiency turned out to be moderate but comparable to LEC<br />

devices made with Ru(II)-type compounds [134]. Wang et al. used the same<br />

complex but, changing experimental conditions, could make a more efficient<br />

green light emitting device (CIE coordinates: 0.25, 0.60) with a maximum<br />

current efficiency <strong>of</strong> 56 cd/A at 4.0 V, corresponding to an external quantum<br />

yield <strong>of</strong> 16% [140]. This work notes the importance <strong>of</strong> the optimization<br />

<strong>of</strong> LEC device parameters such as the response time, which greatly depends<br />

on the counterion, driving voltage, <strong>and</strong> thickness <strong>of</strong> the emitting layer. Further<br />

efforts are needed to substantially improve the device stability <strong>and</strong> light<br />

output in order to take advantage <strong>of</strong> the low-cost <strong>and</strong> limited environmental<br />

damaging effects <strong>of</strong> copper materials.<br />

Finally, it is worth pointing out that also the family <strong>of</strong> cuprous cluster (described<br />

in the next paragraph) has been tested in devices. In the late 1990s<br />

Ma et al. described the electroluminescence properties <strong>of</strong> a LED containing<br />

a tetranuclear Cu(I) cluster as the active component contributing to broaden<br />

the pool <strong>of</strong> electroluminescent materials outside the traditional boundaries <strong>of</strong><br />

organic dyes <strong>and</strong> polymers [141].


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 101<br />

4<br />

Cuprous Clusters<br />

4.1<br />

Cuprous Halide Clusters<br />

Cluster compounds contain a group <strong>of</strong> two or more metal atoms where direct<br />

<strong>and</strong> substantial metal-metal bonding is present. Cuprous halide clusters have<br />

been known for about 100 years [142], their general formula is CunXnLm(X<br />

=Cl – ,Br – or I – ; L = N or P belonging to an organic molecule). For instance,<br />

in solution, mixtures <strong>of</strong> Cu(I) salts, iodine (I) <strong>and</strong> pyridine-type molecules<br />

(py) are primarily present as tetrahedral clusters Cu4I4py4 <strong>and</strong> give origin to<br />

mononuclear or dinuclear structures only if forced by mass action law under<br />

high pyridine concentration. Normally, copper(I) complexes in solution are<br />

quite labile towards lig<strong>and</strong> substitution <strong>and</strong> the formation <strong>of</strong> new species is<br />

driven by thermodynamic stability rather than kinetic control.<br />

Fig. 30 Illustrations <strong>of</strong> Cu4I4py4 (A), Cu2I2py4 (B), [Cu3(µ-dppm)3(µ3 – η 1 -CΞC-benzo-<br />

15-crown-5)2] + (C), <strong>and</strong> the repeating unit <strong>of</strong> a “stairstep” polymer [CuIpy]n (D)redrawn<br />

from the structural data. Black circles =copperatoms;grey circles =iodine;white rings =<br />

pyridine residues


102 N. Armaroli et al.<br />

Cuprous clusters display many structural formats that are characterized by<br />

largely different emission behavior (vide infra). The variety <strong>of</strong> structural motifs<br />

<strong>and</strong> stoichiometries is related to the remarkable flatness <strong>of</strong> their ground<br />

state potential energy surfaces [143]. The most extensive studies carried<br />

out on these Cu(I) complexes concern cubane-type clusters <strong>of</strong> general formula<br />

[CuXL]4 [144, 145]. The solid state structural variety observed among<br />

cuprous clusters can be illustrated by examining some crystallographic data<br />

reported by Ford <strong>and</strong> co-workers, who found that compounds containing<br />

Cu(I), iodine, <strong>and</strong> pyridine generate different structures depending on the<br />

reaction stoichiometry, Fig. 30.<br />

For a 1:1:1 Cu:I:L ratio (Fig. 30 A), the most common motif is the tetranuclear<br />

“cubane” structure (Cu4I4L4) in which a tetrahedron <strong>of</strong> copper atoms is<br />

included by a larger I4 tetrahedron where each iodine is placed on a triangular<br />

face <strong>of</strong> the Cu4 cluster <strong>and</strong> the fourth coordination site <strong>of</strong> each copper is occupied<br />

by the lig<strong>and</strong> (L). For stoichiometry 1:1:2 (Fig. 30 B), the most common<br />

structure is an isolated rhombohedron <strong>of</strong> Cu2I2 with alternating copper <strong>and</strong><br />

halide atoms. Sometimes, clusters with stoichiometry (1:1:1) exist in more<br />

than one crystalline structure. For example (Fig. 30 C) Cu4I4py4 can also give<br />

rise to a polymeric “stair” made <strong>of</strong> an infinite chain <strong>of</strong> steps [146].<br />

4.2<br />

Cuprous Iodide Clusters<br />

The interest in the luminescence properties <strong>of</strong> Cu(I) iodide clusters goes back<br />

to the pioneering work <strong>of</strong> Hardt <strong>and</strong> co-workers [144]. They found that the<br />

emission spectra <strong>of</strong> solid samples <strong>of</strong> [CuxIy(py)z] are markedly temperaturedependent<br />

<strong>and</strong> defined the term “luminescent thermochromism”.<br />

In some cases cuprous iodide clusters exhibit two emission b<strong>and</strong>s termed<br />

HE (high energy) <strong>and</strong> LE (low energy), which sharply change their relative intensities<br />

upon temperature variation. As an example, in Fig. 31 are depicted<br />

the temperature-dependent emission spectra <strong>of</strong> Cu4I4(4 – phenylpyridine)4<br />

[147]. The LE b<strong>and</strong> dominates at room temperature, while the HE b<strong>and</strong> is by<br />

far the strongest at temperatures below 80 K.<br />

The HE b<strong>and</strong> dominating at low temperature has been attributed, on the<br />

basis <strong>of</strong> ab initio calculations <strong>and</strong> experimental work, to lig<strong>and</strong>-to-lig<strong>and</strong><br />

(I – → phenylpyridine) charge transfer states, also indicated as XLCT. The LE<br />

emission dominating at room temperature has been assigned to an excited<br />

state <strong>of</strong> mixed halide-to-metal charge transfer (XMCT) <strong>and</strong> d → s,p metalcentered<br />

character which is usually referred to as “cluster-centered” (CC).<br />

This term was introduced to highlight that these transitions are localized on<br />

the Cu4I4 cluster <strong>and</strong> are essentially independent on lig<strong>and</strong> L. The Cu–Cu distance<br />

is a fundamental parameter to allow the presence <strong>of</strong> CC b<strong>and</strong>s <strong>and</strong> must<br />

be shorter than the orbital interaction radius, estimated to be 2.8 ˚A. Ifthe<br />

distance between the two metal centers exceeds this critical value the metal


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 103<br />

Fig. 31 Temperature dependence <strong>of</strong> the emission spectrum <strong>of</strong> Cu4I4(4 – phenylpyridine)4<br />

in toluene solution with relative intensities normalized to 1 in each case (Reprinted<br />

from [147] with permission, © (1991) American Chemical Society)<br />

orbitals do not interact <strong>and</strong> the CC emission b<strong>and</strong>s are not observed [148]. In<br />

Fig. 32 the relative position <strong>of</strong> the above-described excited states is schematically<br />

represented by means <strong>of</strong> potential energy curves.<br />

When the amine aromatic lig<strong>and</strong> py is replaced by the aliphatic one piperidine<br />

(pip), the corresponding cluster Cu4I4(pip)4 preserves the CC b<strong>and</strong> but<br />

does not display the high energy XLCT emission owing to the absence <strong>of</strong> lig<strong>and</strong>s<br />

possessing π orbitals. Thus, luminescence thermochromism is the norm<br />

for Cu4I4L4 clusters, but only when L is π-unsaturated.<br />

Another factor contributing to the complicated pattern <strong>of</strong> the luminescence<br />

properties <strong>of</strong> Cu4I4(4-phenylpyridine)4 is the dramatic red-shift <strong>of</strong> the<br />

CC b<strong>and</strong> in going from solid or frozen solution samples to fluid solutions, indicating<br />

that also rigidochromism effects are operative. Ab initio calculations<br />

<strong>and</strong> the Stokes shifts for Cu4I4(4-phenylpyridine)4 (up to 16 300 cm –1 for the<br />

CC b<strong>and</strong> in 296 Ktoluenesolution;7600 cm –1 for the XLCT b<strong>and</strong> under the


104 N. Armaroli et al.<br />

Fig. 32 Schematic representation describing the relative positions <strong>of</strong> the potential energy<br />

curves related to the emitting states <strong>of</strong> Cu4I4py4<br />

same conditions) also indicate that CC excited states are quite distorted relative<br />

to both the GS <strong>and</strong> XLCT levels. Such distortion was proposed as the<br />

cause for the lack <strong>of</strong> communication between CC <strong>and</strong> XLCT states, for the<br />

rigidochromism <strong>of</strong> the low CC energy b<strong>and</strong>, <strong>and</strong> for the large reorganization<br />

energies <strong>of</strong> electron transfer reactions involving CC excited states [143].<br />

In an interesting recent work on clusters <strong>of</strong> general formula CuXL (L = N–<br />

heteroaromatic lig<strong>and</strong>s), it was shown that the energy <strong>of</strong> the emitting level<br />

can be finely tuned [149]. The emission is attributed to a MLCT charge transfer<br />

state, because no clear correlation between Cu–Cu distances <strong>and</strong> emission<br />

maxima was observed <strong>and</strong> also because the effects <strong>of</strong> bridging halides were<br />

smaller than those <strong>of</strong> N-heteroaromatic lig<strong>and</strong>s, therefore the position <strong>of</strong> the<br />

luminescence b<strong>and</strong> can be varied by increasing the electron-accepting character<br />

<strong>of</strong> the lig<strong>and</strong> L, Fig. 33. In these compounds Cu–Cu distances are in<br />

the range 2.9–3.3 ˚A, accordingly the weak metal interaction prevent clustercentered<br />

luminescence. Very recently, density functional theory calculations<br />

have confirmed the involvement <strong>of</strong> the triplet cluster-centered (CC) <strong>and</strong><br />

triplet XLCT excited states as the origin <strong>of</strong> the dual emission [151].<br />

It must be pointed out, however, that short Cu–Cu separation does not automatically<br />

imply the establishment <strong>of</strong> metal-metal bonds <strong>and</strong> the effect <strong>of</strong><br />

the bridging lig<strong>and</strong>s has to be taken into account. For example Cotton et al.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 105<br />

Fig. 33 Emission spectra <strong>of</strong> clusters CuXL (X = Br – ) at room temperature. The emission<br />

maxima <strong>of</strong> the complexes cover a wide range, from 450 to 740 nm depending on the Nheteroaromatic<br />

lig<strong>and</strong>s. bpy = 4,4-bipyridine; pyz = pyrazine; pym = pymidine; pip =<br />

piperazine; 1,5-nap = 1,5-naphthyridine; 1,6-nap = 1,6-naphthyridine; quina = quinazoline;<br />

dmap = N,N-dimethyl-amino-pyridine; 3-bzpy = 3-benzoyl-pyridine; 4-bzpy =<br />

4-benzoyl-pyridine. (Reprinted from [149] with permission, © (2006) American Chemical<br />

Society)<br />

have carried out DTF calculations on [Cu2(hpp)2], (where hpp – = 1,3,4,6,7,8hexahydro-2Hpyrimido[1,2-a]pyrimidinate)<br />

to investigate the possibility <strong>of</strong><br />

metal-metal bonding in a complex where short metal–metal separations are<br />

present [dCu···Cu = 2.497(2) ˚A]. They concluded that there is no Cu–Cu<br />

bond <strong>and</strong> the short intermetal distance is related to the strong Cu–N bonds<br />

<strong>and</strong> the small bite angle <strong>of</strong> the bridging lig<strong>and</strong> [150].<br />

4.3<br />

OtherCopperClusters<br />

Recently, the synthesis <strong>of</strong> several polynuclear copper(I) alkynyl clusters has<br />

been reported <strong>and</strong> their luminescence properties investigated in detail [152,<br />

153]; these compounds exhibit intense <strong>and</strong> long-lived luminescence upon<br />

photoexcitation. For instance, the tetranuclear copper(I) alkynyl complex<br />

[Cu4(PPh3)4(L)3]PF6, in Fig. 34 is characterized by an unusual open-cube<br />

structure, <strong>and</strong> exhibits a strong structured emission with two different max-


106 N. Armaroli et al.<br />

Fig. 34 A tetranuclear copper(I) alkynyl “open-cube” cluster<br />

ima at 445 <strong>and</strong> 630 nm in solid state at 298 K <strong>and</strong> a single b<strong>and</strong> with λmax<br />

= 445 nm in rigid matrix at 77 K [154, 155].<br />

For some <strong>of</strong> these open-cube compounds, an additional low-energy<br />

emission b<strong>and</strong> at λ > 623–665 nm was observed in the solid-state spectra,<br />

similarly to what was observed for [Cu4I4L4] systems described above.<br />

In dichloromethane solution at ambient temperature they exhibit only<br />

an orange phosphorescence <strong>and</strong> the spectrum <strong>of</strong> [Cu4(PPh3)4(L)3]PF6 (L<br />

= p – nOctC6H4) is depicted in Fig. 35 as a representative example for this<br />

class <strong>of</strong> compounds.<br />

Fig. 35 Emission spectrum <strong>of</strong> [Cu4(PPh3)4(L)3]PF6 (L = p – nOctC6H4) in degassed<br />

dichloromethane at 298 K (Reprinted from [155] with permission, © (2006) Wiley)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 107<br />

There are other examples <strong>of</strong> luminescent clusters, for instance trinuclear<br />

copper(I) pyrazolates displaying emission b<strong>and</strong>s over a wide spectral<br />

range [156], or others with a core made <strong>of</strong> four Cu(I) <strong>and</strong> four sulfur<br />

atoms [157, 158]. There are examples <strong>of</strong> Cu(I) luminescent clusters with<br />

a higher nuclearity, some <strong>of</strong> them are heterometallic (6–8 metal centers) [159]<br />

while others are homometallic but with a higher nuclearity (16–20 metal centers).<br />

In the latter case, which bear alkynyl lig<strong>and</strong>s [160], an emission b<strong>and</strong> is<br />

observed in the UV spectral region. Finally, the possible use <strong>of</strong> copper cluster<br />

units to assemble polymeric compounds with a wide range <strong>of</strong> possible<br />

structures, from one- to three-dimensional should be noted [161, 162].<br />

5<br />

Miscellanea <strong>of</strong> Cu(I) Luminescent Complexes<br />

Intheprevioussectionswehavepresentedthethreemainclasses<strong>of</strong>Cu(I)<br />

compounds exhibiting interesting photophysical properties, namely Cu(I)bisphenanthrolines,<br />

[Cu(NN)(PP)] + complexes <strong>and</strong> cuprous clusters. However,<br />

especially in recent years, a growing number <strong>of</strong> Cu(I) luminescent complexes<br />

with less conventional lig<strong>and</strong>s have appeared in the literature <strong>and</strong> some<br />

<strong>of</strong> them will be now briefly presented.<br />

The homoleptic Cu(I) complexes <strong>of</strong> the benzo[h]quinoline lig<strong>and</strong>s (BHQ)<br />

depicted in Fig. 36 exhibit excellent luminescence properties in CH2Cl2<br />

withquantumyieldsashighas0.10<strong>and</strong>τ = 5.3 µs (lig<strong>and</strong>C in Fig. 36),<br />

Fig. 36 Benzo[h]quinoline lig<strong>and</strong>s which, upon complexation with Cu(I), provide highly<br />

luminescent complexes


108 N. Armaroli et al.<br />

which are values comparable to those <strong>of</strong> [Cu(NN)(POP)] + compounds (see<br />

Sect. 3.1) [163].<br />

This relevant result has been rationalized assuming that the specific complex<br />

structure imposes minimal geometrical changes in the deactivation<br />

process <strong>of</strong> the luminescent excited state back to the ground state, while maintaining<br />

a significant energy gap. The BHQ lig<strong>and</strong>s accomplish exactly this<br />

requirement by providing considerable steric congestion in the vicinity <strong>of</strong><br />

chelation, which stabilizes Cu(I), while also providing interlig<strong>and</strong> π-stacking<br />

that distorts the ground-state geometry <strong>and</strong> favors the (formal) oxidation <strong>of</strong><br />

the metal with little structural change [163]. The same authors later proposed<br />

Cu(I) complexes made <strong>of</strong> bisphenanthroline lig<strong>and</strong>s with structures identical<br />

to A, B <strong>and</strong> C (Fig. 36), but no emission data were presented [164].<br />

2-Hydroxy-1,10-phenanthroline (Hophen) is a novel kind <strong>of</strong> substituted<br />

phenanthroline that was recently proposed (Fig. 37). With Cu(I) it gives origin<br />

to several compounds, as evidenced by X-ray crystallography, including<br />

an unusual neutral dinuclear complex [Cu2(ophen)2] which exists in three<br />

supramolecular isomeric forms <strong>and</strong> exhibits a broad <strong>and</strong> weak luminescenceb<strong>and</strong>centeredaround630<br />

nm, which has been tentatively attributed<br />

to deactivation <strong>of</strong> an MLCT state [165]. The same lig<strong>and</strong>, which is characterized<br />

by complicated coordination modes involving both the regular nitrogen<br />

sites <strong>and</strong> oxygen binding another metal ion (Fig. 37), was also used<br />

to make metal complexes <strong>of</strong> other d 10 metal ions, i.e. Zn(II), Cd(II) <strong>and</strong><br />

Hg(II), which show lig<strong>and</strong> centered (LC) emission b<strong>and</strong>s in the blue-green<br />

region [166].<br />

Fig. 37 The proposed ketone <strong>and</strong> hydroxy tautomers <strong>of</strong> Hophen<br />

Vogler et al. have made several Cu(I) complexes exhibiting emission in different<br />

regions <strong>of</strong> the VIS spectral window, including blue <strong>and</strong> red, <strong>and</strong> having<br />

pure MLCT or mixed MLCT/LLCT character [167, 168]. For instance the complex<br />

depicted in Fig. 38 which is easily accessible from commercially available<br />

productsshowsaweakbutdistinctMLCTredluminescencepeakingat600<br />

nm [167].<br />

Several other unconventional lig<strong>and</strong>s have been utilized recently to make<br />

luminescent Cu(I) complexes such as thia-calix[3]pyridine (orange MLCT


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 109<br />

Fig. 38 The red-emitting complex [Cu(BTA)(hfac)] with BTA = bis(trimethylsilyl)acetylene<br />

<strong>and</strong> hfac = 1,1,1,5,5,5-hexa-fluoroacetyl-acetonate<br />

phosphorescence) [169], mononuclear <strong>and</strong> dinuclear azaindole-containing<br />

systems (blue LC) [170, 171], heteroleptic diphosphine/diisocyanide lig<strong>and</strong>s<br />

(blue MLCT) [172], pyrazolylborates (blue LC phosphorescence) [173].<br />

6<br />

Conclusions <strong>and</strong> Perspectives<br />

In recent years, the rationalization <strong>of</strong> the electronic <strong>and</strong> photophysical properties<br />

<strong>of</strong> Cu(I) compounds has made considerable progress, in parallel with<br />

a significant implementation <strong>of</strong> synthetic protocols to prepare both simple<br />

<strong>and</strong> complex Cu(I)-based structures with satisfactory yields. Now we know<br />

key design principles that allow one to make highly luminescent Cu(I) compounds<br />

<strong>and</strong> supramolecular architectures with programmable cascades <strong>of</strong><br />

photoinduced processes. Such trends have taken Cu(I) complexes among the<br />

key players in the realm <strong>of</strong> photoactive complexes, where other metals such<br />

as Ru(II) <strong>and</strong>, more recently Ir(III) have traditionally played a prominent<br />

role. The relentless quest for luminescent metal compounds to be utilized in<br />

a variety <strong>of</strong> applications can find interesting answers among some classes <strong>of</strong><br />

Cu(I) complexes, as pointed out in this review article. Obviously, the need for<br />

abundant, cheap, <strong>and</strong> environmentally friendly smart materials makes copper<br />

compounds an attractive alternative to more traditional choices based on<br />

precious metals. The current trends in literature <strong>and</strong> patenting [174] indeed<br />

suggest that Cu(I) complexes are attracting increasing attention for technological<br />

applications (e.g. OLEDs) <strong>and</strong>, although we are still at the level <strong>of</strong><br />

prototypes <strong>and</strong> pro<strong>of</strong>s <strong>of</strong> principles, further important breakthroughs may be<br />

anticipated in the years to come.


110 N. Armaroli et al.<br />

Acknowledgements We thank the CNR (Progetto “Sistemi nanoorganizzati con proprietà<br />

elettroniche, fotoniche e magnetiche, commessa PM.P04.010 (MACOL)”) <strong>and</strong> the<br />

EC through the Integrated Project OLLA (contract no. IST-2002-004607) for financial<br />

support. Over the years we worked on several collaborative projects related to Cu(I)<br />

complexes <strong>and</strong>, in this regard, we wish to thank Jean-François Nierengarten (Toulouse,<br />

France), Jean-Pierre Sauvage (Strasbourg, France) <strong>and</strong> Michael Schmittel (Siegen, Germany)<br />

along with many other colleagues from their research groups, whose names are<br />

cited in the references.<br />

References<br />

1. US Geological Survey (2006) Mineral Commodity Summaries http://minerals.er.usgs.<br />

gov/minerals/pubs/commodity/copper/, last visited: January 2006<br />

2. Gordon RB, Bertram M, Graedel TE (2006) Proc Natl Acad Sci USA 103:1209–1214<br />

3. Horvath O (1994) Coord Chem Rev 135:303–324<br />

4. Sykora J (1997) Coord Chem Rev 159:95–108<br />

5. Jørgensen CK (1963) Adv Chem Phys 5:33–146<br />

6. Juris A, Balzani V, Barigelletti F, Campagna S, Belser P, von Zelewsky A (1988) Coord<br />

Chem Rev 84:85–277<br />

7. Lowry MS, Bernhard S (2006) Chem Eur J 12:7970–7977<br />

8. Balzani V, Juris A, Venturi M, Campagna S, Serroni S (1996) Chem Rev 96:759–833<br />

9. Roundhill DM (1994) <strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> Metal Complexes.<br />

Plenum Press, NY<br />

10. Bignozzi CA, Argazzi R, Kleverlaan CJ (2000) Chem Soc Rev 29:87–96<br />

11. Balzani V, Ceroni P, Juris A, Venturi M, Campagna S, Puntoriero F, Serroni S (2001)<br />

Coord Chem Rev 219:545–572<br />

12. Grätzel M (2005) Inorg Chem 44:6841–6851<br />

13. Vos JG, Kelly JM (2006) Dalton Trans pp 4869–4883<br />

14. Kober EM, Caspar JV, Lumpkin RS, Meyer TJ (1986) J Phys Chem 90:3722–3734<br />

15. Armaroli N (2001) Chem Soc Rev 30:113–124<br />

16. Maestri M, Armaroli N, Balzani V, Constable EC, Thompson A (1995) Inorg Chem<br />

34:2759–2767<br />

17. Lippard SJ, Berg JM (1994) Principles <strong>of</strong> Bioinorganic Chemistry. University Science<br />

Books, Mill Valley, California<br />

18. Holm RH, Kennepohl P, Solomon EI (1996) Chem Rev 96:2239–2314<br />

19. Colman PM, Freeman HC, Guss JM, Murata M, Norris VA, Ramshaw JAM, Venkatappa<br />

MP (1978) Nature 272:319–324<br />

20. Solomon EI (2006) Inorg Chem 45:8012–8025<br />

21. Gewirth AA, Solomon EI (1988) J Am Chem Soc 110:3811–3819<br />

22. Marcus RA, Sutin N (1985) Biochim Biophys Acta 811:265–322<br />

23. Babcock GT, Wikstrom M (1992) Nature 356:301–309<br />

24. Gamelin DR, R<strong>and</strong>all DW, Hay MT, Houser RP, Mulder TC, Canters GW, de Vries S,<br />

Tolman WB, Lu Y, Solomon EI (1998) J Am Chem Soc 120:5246–5263<br />

25. Armaroli N (2003) Photochem Photobiol Sci 2:73–87<br />

26. Schmittel M, Kalsani V (2005) Top Curr Chem 245:1–53<br />

27. Sammes PG, Yahioglu G (1994) Chem Soc Rev 23:327–334<br />

28. Cunningham CT, Moore JJ, Cunningham KLH, Fanwick PE, McMillin DR (2000)<br />

Inorg Chem 39:3638–3644


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 111<br />

29. Armaroli N, Balzani V, Barigelletti F, De Cola L, Flamigni L, Sauvage JP, Hemmert C<br />

(1994) J Am Chem Soc 116:5211–5217<br />

30. Dietrich-Buchecker C, Colasson B, Fujita M, Hori A, Geum N, Sakamoto S, Yamaguchi<br />

K, Sauvage JP (2003) J Am Chem Soc 125:5717–5725<br />

31. Frey J, Kraus T, Heitz V, Sauvage JP (2005) Chem Commun, pp 5310–5312<br />

32. Armaroli N, Balzani V, Collin JP, Gaviña P, Sauvage JP, Ventura B (1999) J Am Chem<br />

Soc 121:4397–4408<br />

33. Weber N, Hamann C, Kern JM, Sauvage JP (2003) Inorg Chem 42:6780–6792<br />

34. Baran<strong>of</strong>f E, Griffiths K, Collin JP, Sauvage JP, Ventura B, Flamigni L (2004) New<br />

J Chem 28:1091–1095<br />

35. Kraus T, Budesinsky M, Cvacka JC, Sauvage JP (2006) Angew Chem Int Ed 45:258–<br />

261<br />

36. Dietrich-Buchecker CO, Nierengarten JF, Sauvage JP, Armaroli N, Balzani V, De<br />

Cola L (1993) J Am Chem Soc 115:11237–11244<br />

37. Perret-Aebi LE, von Zelewsky A, Dietrich-Buchecker CD, Sauvage JP (2004) Angew<br />

Chem Int Ed 43:4482–4485<br />

38. Armaroli N, Boudon C, Felder D, Gisselbrecht JP, Gross M, Marconi G, Nicoud JF,<br />

Nierengarten JF, Vicinelli V (1999) Angew Chem Int Ed 38:3730–3733<br />

39. Gumienna-Kontecka E, Rio Y, Bourgogne C, Elhabiri M, Louis R, Albrecht-Gary AM,<br />

Nierengarten JF (2004) Inorg Chem 43:3200–3209<br />

40. Heuft MA, Fallis AG (2002) Angew Chem Int Ed 41:4520–4523<br />

41. Cardinali F, Mamlouk H, Rio Y, Armaroli N, Nierengarten JF (2004) Chem Commun,<br />

pp 1582–1583<br />

42. Zong RF, Thummel RP (2005) Inorg Chem 44:5984–5986<br />

43. Ziessel R, Charbonniere L, Cesario M, Prange T, Nierengarten H (2002) Angew<br />

Chem Int Ed 41:975–979<br />

44. Sauvage J-P, Dietrich-Buchecker CO (eds) (1999) Molecular Catenanes, Rotaxanes<br />

<strong>and</strong> Knots. A Journey through the World <strong>of</strong> Molecular Topology. Wiley-VCH, Weinheim,<br />

Germany<br />

45. Jimenez-Molero MC, Dietrich-Buchecker C, Sauvage JP (2002) Chem Eur J 8:1456–<br />

1466<br />

46. Livoreil A, Sauvage JP, Armaroli N, Balzani V, Flamigni L, Ventura B (1997) J Am<br />

Chem Soc 119:12114–12124<br />

47. Sauvage JP (2005) Chem Commun, pp 1507–1510<br />

48. Bonnet S, Collin JP, Koizumi M, Mobian P, Sauvage JP (2006) Adv Mater 18:1239–<br />

1250<br />

49. Kalsani V, Bodenstedt H, Fenske D, Schmittel M (2005) Eur J Inorg Chem 1841–1849<br />

50. Schmittel M, Kalsani V, Fenske D, Wiegrefe A (2004) Chem Commun, pp 490–491<br />

51. Schmittel M, Ammon H, Kalsani V, Wiegrefe A, Michel C (2002) Chem Commun,<br />

pp 2566–2567<br />

52. Kalsani V, Ammon H, Jäckel F, Rabe JP, Schmittel M (2004) Chem-Eur J 10:5481–<br />

5492<br />

53. Schmittel M, Ganz A (1997) Chem Commun, pp 999–1000<br />

54. Dobson JF, Green BE, Healy PC, Kennard CHL, Pakawatchai C, White AH (1984)<br />

Aust J Chem 37:649–659<br />

55. Coppens P, Vorontsov II, Graber T, Kovalevsky AY, Chen YS, Wu G, Gembicky M,<br />

Novozhilova IV (2004) J Am Chem Soc 126:5980–5981<br />

56. Kovalevsky AY, Gembicky M, Coppens P (2004) Inorg Chem 43:8282–8289<br />

57. Kovalevsky AY, Gembicky M, Novozhilova IV, Coppens P (2003) Inorg Chem<br />

42:8794–8802


112 N. Armaroli et al.<br />

58. Miller MT, Gantzel PK, Karpishin TB (1998) Angew Chem Int Ed Engl 37:1556–1558<br />

59. Miller MT, Gantzel PK, Karpishin TB (1998) Inorg Chem 37:2285–2290<br />

60. Armaroli N, De Cola L, Balzani V, Sauvage JP, Dietrich-Buchecker CO, Kern JM<br />

(1992) J Chem Soc Faraday Trans 88:553–556<br />

61. Zgierski MZ (2003) J Chem Phys 118:4045–4051<br />

62. McMillin DR, Buckner MT, Ahn BT (1977) Inorg Chem 16:943–945<br />

63. McMillin DR, McNett KM (1998) Chem Rev 98:1201–1219<br />

64. Scaltrito DV, Thompson DW, O’Callaghan JA, Meyer GJ (2000) Coord Chem Rev<br />

208:243–266<br />

65. Federlin P, Kern JM, Rastegar A, Dietrich-Buchecker C, Marnot PA, Sauvage JP<br />

(1990) New J Chem 14:9–12<br />

66. Gordon KC, McGarvey JJ (1991) Inorg Chem 30:2986–2989<br />

67. Armaroli N, Rodgers MAJ, Ceroni P, Balzani V, Dietrich-Buchecker CO, Kern JM,<br />

Bailal A, Sauvage JP (1995) Chem Phys Lett 241:555–558<br />

68. Ichinaga AK, Kirchh<strong>of</strong>f JR, McMillin DR, Dietrich-Buchecker CO, Marnot PA,<br />

Sauvage JP (1987) Inorg Chem 26:4290–4292<br />

69. Phifer CC, McMillin DR (1986) Inorg Chem 25:1329–1333<br />

70. Everly RM, McMillin DR (1991) J Phys Chem 95:9071–9075<br />

71. Cunningham CT, Cunningham KLH, Michalec JF, McMillin DR (1999) Inorg Chem<br />

38:4388–4392<br />

72. Miller MT, Gantzel PK, Karpishin TB (1999) Inorg Chem 38:3414–3422<br />

73. Miller MT, Karpishin TB (1999) Inorg Chem 38:5246–5249<br />

74. Kalsani V, Schmittel M, Listorti A, Accorsi G, Armaroli N (2006) Inorg Chem<br />

45:2061–2067<br />

75. Gushurst AKI, McMillin DR, Dietrich-Buchecker CO, Sauvage JP (1989) Inorg Chem<br />

28:4070–4072<br />

76. Goodman MS, Hamilton AD, Weiss J (1995) J Am Chem Soc 117:8447–8455<br />

77. Amendola V, Boiocchi M, Colasson B, Fabbrizzi L (2006) Inorg Chem 45:6138–6147<br />

78. Everly RM, McMillin DR (1989) Photochem Photobiol 50:711–716<br />

79. Chen LX (2005) Annu Rev Phys Chem 56:221–254<br />

80. Chen LX (2004) Angew Chem Int Ed 43:2886–2905<br />

81. Chen LX, Jennings G, Liu T, Gosztola DJ, Hessler JP, Scaltrito DV, Meyer GJ (2002)<br />

J Am Chem Soc 124:10861–10867<br />

82. ChenLX,ShawGB,NovozhilovaI,LiuT,JenningsG,Attenk<strong>of</strong>erK,MeyerGJ,Coppens<br />

P (2003) J Am Chem Soc 125:7022–7034<br />

83. Coppens P (2003) Chem Commun pp 1317–1320<br />

84. Gunaratne T, Rodgers MAJ, Felder D, Nierengarten JF, Accorsi G, Armaroli N (2003)<br />

Chem Commun pp 3010–3011<br />

85. Felder D, Nierengarten JF, Barigelletti F, Ventura B, Armaroli N (2001) J Am Chem<br />

Soc 123:6291–6299<br />

86. Cody J, Dennisson J, Gilmore J, VanDerveer DG, Henary MM, Gabrielli A, Sherrill<br />

CD, Zhang YY, Pan CP, Burda C, Fahrni CJ (2003) Inorg Chem 42:4918–4929<br />

87. Blaskie MW, McMillin DR (1980) Inorg Chem 19:3519–3522<br />

88.WilliamsRM,DeColaL,HartlF,LagrefJJ,PlaneixJM,DeCianA,HosseiniMW<br />

(2002) Coord Chem Rev 230:253–261<br />

89. Siddique ZA, Yamamoto Y, Ohno T, Nozaki K (2003) Inorg Chem 42:6366–6378<br />

90. Miller MT, Gantzel PK, Karpishin TB (1999) J Am Chem Soc 121:4292–4293<br />

91. Parker WL, Crosby GA (1989) J Phys Chem 93:5692–5696<br />

92.Kirchh<strong>of</strong>fJR,GamacheRE,BlaskieMW,DelPaggioAA,LengelRK,McMillinDR<br />

(1983) Inorg Chem 22:2380–2384


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 113<br />

93. Cardenas DJ, Collin JP, Gaviña P, Sauvage JP, De Cian A, Fischer J, Armaroli N,<br />

Flamigni L, Vicinelli V, Balzani V (1999) J Am Chem Soc 121:5481–5488<br />

94.FlamigniL,TalaricoAM,ChambronJC,HeitzV,LinkeM,FujitaN,SauvageJP<br />

(2004) Chem-Eur J 10:2689–2699<br />

95. Armaroli N, Balzani V, Barigelletti F, Decola L, Sauvage JP, Hemmert C (1991) J Am<br />

Chem Soc 113:4033–4035<br />

96. Armaroli N, Balzani V, De Cola L, Hemmert C, Sauvage JP (1994) New J Chem<br />

18:775–782<br />

97. Dietrich-Buchecker CO, Sauvage JP, Armaroli N, Ceroni P, Balzani V (1996) Angew<br />

Chem Int Ed Engl 35:1119–1121<br />

98. Armaroli N, Diederich F, Dietrich-Buchecker CO, Flamigni L, Marconi G, Nierengarten<br />

JF, Sauvage JP (1998) Chem-Eur J 4:406–416<br />

99. S<strong>and</strong>anayaka ASD, Watanabe N, Ikeshita KI, Araki Y, Kihara N, Furusho Y, Ito O,<br />

Takata T (2005) J Phys Chem B 109:2516–2525<br />

100. Li K, Bracher PJ, Guldi DM, Herranz MA, Echegoyen L, Schuster DI (2004) J Am<br />

Chem Soc 126:9156–9157<br />

101. Li K, Schuster DI, Guldi DM, Herranz MA, Echegoyen L (2004) J Am Chem Soc<br />

126:3388–3389<br />

102. Watanabe N, Kihara N, Furusho Y, Takata T, Araki Y, Ito O (2003) Angew Chem Int<br />

Ed 42:681–683<br />

103. Linke M, Chambron SC, Heitz V, Sauvage SP, Encinas S, Barigelletti F, Flamigni L<br />

(2000) J Am Chem Soc 122:11834–11844<br />

104. Andersson M, Linke M, Chambron JC, Davidsson J, Heitz V, Hammarström L,<br />

Sauvage JP (2002) J Am Chem Soc 124:4347–4362<br />

105. Andersson M, Linke M, Chambron JC, Davidsson J, Heitz V, Sauvage JP, Hammarström<br />

L (2000) J Am Chem Soc 122:3526–3527<br />

106. Flamigni L, Armaroli N, Barigelletti F, Chambron JC, Sauvage JP, Solladié N (1999)<br />

New J Chem 23:1151–1158<br />

107. Chambron JC, Harriman A, Heitz V, Sauvage JP (1993) J Am Chem Soc 115:6109–6114<br />

108. Chambron JC, Harriman A, Heitz V, Sauvage JP (1993) J Am Chem Soc 115:7419–<br />

7425<br />

109. Holler M, Cardinali F, Mamlouk H, Nierengarten JF, Gisselbrecht JP, Gross M, Rio Y,<br />

Barigelletti F, Armaroli N (2006) Tetrahedron 62:2060–2073<br />

110. Rio Y, Enderlin G, Bourgogne C, Nierengarten JF, Gisselbrecht JP, Gross M, Accorsi<br />

G, Armaroli N (2003) Inorg Chem 42:8783–8793<br />

111. Clifford JN, Accorsi G, Cardinali F, Nierengarten JF, Armaroli N (2006) C R Chim<br />

9:1005–1013<br />

112. Flamigni L, Heitz V, Sauvage JP (2006) Struct Bond 121:217–261<br />

113. Cunningham KL, McMillin DR (1998) Inorg Chem 37:4114–4119<br />

114. Cunningham KL, Hecker CR, McMillin DR (1996) Inorg Chim Acta 242:143–147<br />

115. Ruthkosky M, Kelly CA, Castellano FN, Meyer GJ (1998) Coord Chem Rev 171:309–<br />

322<br />

116. Ruthkosky M, Castellano FN, Meyer GJ (1996) Inorg Chem 35:6406–6412<br />

117. Castellano FN, Ruthkosky M, Meyer GJ (1995) Inorg Chem 34:3–4<br />

118. Meskers SCJ, Dekkers H, Rapenne G, Sauvage JP (2000) Chem-Eur J 6:2129–2134<br />

119. Buckner MT, McMillin DR (1978) J Chem Soc-Chem Commun, pp 759–761<br />

120. Cuttell DG, Kuang SM, Fanwick PE, McMillin DR, Walton RA (2002) J Am Chem Soc<br />

124:6–7<br />

121. McCormick T, Jia WL, Wang SN (2006) Inorg Chem 45:147–155<br />

122. Tsukuda T, Nakamura A, Arai T, Tsubomura T (2006) Bull Chem Soc Jpn 79:288–290


114 N. Armaroli et al.<br />

123. Armaroli N, Accorsi G, Holler M, Moudam O, Nierengarten JF, Zhou Z, Wegh RT,<br />

Welter R (2006) Adv Mater 18:1313–1316<br />

124. Yang L, Feng JK, Ren AM, Zhang M, Ma YG, Liu XD (2005) Eur J Inorg Chem 1867–<br />

1879<br />

125. Howell SL, Gordon KC (2004) J Phys Chem A 108:2536–2544<br />

126. Kuang SM, Cuttell DG, McMillin DR, Fanwick PE, Walton RA (2002) Inorg Chem<br />

41:3313–3322<br />

127. Englman R, Jortner J (1970) Mol Phys 18:145–164<br />

128. Rader RA, McMillin DR, Buckner MT, Matthews TG, Casadonte DJ, Lengel RK, Whittaker<br />

SB, Darmon LM, Lytle FE (1981) J Am Chem Soc 103:5906–5912<br />

129. Palmer CEA, McMillin DR, Kirmaier C, Holten D (1987) Inorg Chem 26:3167–3170<br />

130. Tsubomura T, Takahashi N, Saito K, Tsukuda T (2004) Chem Lett 33:678–679<br />

131. Saito K, Arai T, Takahashi N, Tsukuda T, Tsubomura T (2006) Dalton Trans pp 4444–<br />

4448<br />

132. Tsubomura T, Enoto S, Endo S, Tamane T, Matsumoto K, Tsukuda T (2005) Inorg<br />

Chem 44:6373–6378<br />

133. Jia WL, McCormick T, Tao Y, Lu JP, Wang SN (2005) Inorg Chem 44:5706–5712<br />

134. Slinker J, Bernards D, Houston PL, Abruna HD, Bernhard S, Malliaras GG (2003)<br />

Chem Commun, pp 2392–2399<br />

135. Schubert EF, Kim JK (2005) Science 308:1274–1278<br />

136. Bolink HJ, Cappelli L, Coronado E, Gavina P (2005) Inorg Chem 44:5966–5968<br />

137. Holder E, Langeveld BMW, Schubert US (2005) Adv Mater 17:1109–1121<br />

138. Zhang QS, Zhou QG, Cheng YX, Wang LX, Ma DG, Jing XB, Wang FS (2004) Adv<br />

Mater 16:432–436<br />

139. Che GB, Su ZS, Li WL, Chu B, Li MT, Hu ZZ, Zhang ZQ (2006) Appl Phys Lett<br />

89:103511<br />

140. Zhang QS, Zhou QG, Cheng YX, Wang LX, Ma DG, Jing XB, Wang FS (2006) Adv<br />

Funct Mater 16:1203–1208<br />

141. Ma YG, Che CM, Chao HY, Zhou XM, Chan WH, Shen JC (1999) Adv Mater 11:852–<br />

857<br />

142. Raston CL, White AH (1976) J Chem Soc Dalton Trans 21:2153–2156<br />

143. Vitale M, Ford PC (2001) Coord Chem Rev 219:3–16<br />

144. Hardt HD, Pierre A (1973) Z Anorg Allg Chem 402:107<br />

145. Ford PC, Cariati E, Bourassa J (1999) Chem Rev 99:3625–3647<br />

146. Eitel E, Oelkrug D, Hiller W, Strahle J (1980) Z Naturforsch (B) 35:1247–1253<br />

147. Kyle KR, Ryu CK, DiBenedetto JA, Ford PC (1991) J Am Chem Soc 113:2954–2965<br />

148. Rath NP, Holt EM, Tanimura K (1986) J Chem Soc-Dalton Trans pp 2303–2310<br />

149. Araki H, Tsuge K, Sasaki Y, Ishizaka S, Kitamura N (2005) Inorg Chem 44:9667–9675<br />

150. Cotton FA, Feng XJ, Timmons DJ (1998) Inorg Chem 37:4066–4069<br />

151. De Angelis F, Fantacci S, Sgamellotti A, Cariati E, Ugo R, Ford PC (2006) Inorg Chem<br />

45:10576–10584<br />

152. Yam VWW, Lo KKW, Wong KMC (1999) J Organomet Chem 578:3–30<br />

153. Yam VWW (2002) Acc Chem Res 35:555–563<br />

154. Yam VWW, Choi SWK, Chan CL, Cheung KK (1996) Chem Commun, pp 2067–2068<br />

155. Chan CL, Cheung KK, Lam WH, Cheng ECC, Zhu N, Choi SWK, Yam VWW (2006)<br />

Chem-Asian J 1–2:273<br />

156. Dias HVR, Diyabalanage HVK, Eldabaja MG, Elbjeirami O, Rawashdeh-Omary MA,<br />

Omary MA (2005) J Am Chem Soc 127:7489–7501<br />

157. Che CM, Xia BH, Huang JS, Chan CK, Zhou ZY, Cheung KK (2001) Chem-Eur J<br />

7:3998–4006


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Copper 115<br />

158. Kharenko OA, Kennedy DC, Demeler B, Maroney MJ, Ogawa MY (2005) J Am Chem<br />

Soc 127:7678–7679<br />

159. Wei QH, Yin GQ, Zhang LY, Shi LX, Mao ZW, Chen ZN (2004) Inorg Chem 43:3484–<br />

3491<br />

160. Baxter CW, Higgs AC, Jones AC, Parsons S, Bailey PJ, Tasker PA (2002) J Chem Soc<br />

Dalton Trans 4395–4401<br />

161. Peng R, Li D, Wu T, Zhou XP, Ng SW (2006) Inorg Chem 45:4035–4046<br />

162. He X, Lu CZ, Wu CD, Chen LJ (2006) Eur J Inorg Chem, pp 2491–2503<br />

163. Riesgo EC, Hu YZ, Bouvier F, Thummel RP, Scaltrito DV, Meyer GJ (2001) Inorg<br />

Chem 40:3413–3422<br />

164. Riesgo EC, Hu YZ, Thummel RP (2003) Inorg Chem 42:6648–6654<br />

165. Zhang XM, Tong ML, Gong ML, Lee HK, Luo L, Li KF, Tong YX, Chen XM (2002)<br />

Chem-Eur J 8:3187–3194<br />

166. Zheng SL, Zhang JP, Chen XM, Huang ZL, Lin ZY, Wong WT (2003) Chem-Eur J<br />

9:3888–3896<br />

167. Kunkely H, Vogler A (2003) Inorg Chem Commun 6:543–545<br />

168. Pawlowski V, Knor G, Lennartz C, Vogler A (2005) Eur J Inorg Chem 3167–3171<br />

169. Kinoshita I, Hamazawa A, Nishioka T, Adachi H, Suzuki H, Miyazaki Y, Tsuboyama<br />

A, Okada S, Hoshino M (2003) Chem Phys Lett 371:451–457<br />

170. Song DT, Jia WL, Wu G, Wang SN (2005) Dalton Trans pp 433–438<br />

171. Zhao SB, Wang RY, Wang SN (2006) Inorg Chem 45:5830–5840<br />

172. Fournier E, Lebrun F, Drouin M, Decken A, Harvey PD (2004) Inorg Chem 43:3127–<br />

3135<br />

173. Omary MA, Rawashdeh-Omary MA, Diyabalanage HVK, Rasika Dias HV (2003)<br />

Inorg Chem 42:8612–8614<br />

174. Tsuboyama A, Okada S, Takiguchi T, Igawa S, Kamatani J, Furugori M, Canon KK<br />

(2005) JP Patent n. US Patent 2 005 014 024


Top Curr Chem (2007) 280: 117–214<br />

DOI 10.1007/128_2007_133<br />

© Springer-Verlag Berlin Heidelberg<br />

Published online: 27 June 2007<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong><br />

<strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium<br />

Sebastiano Campagna 1 (✉)·FaustoPuntoriero 1 ·FrancescoNastasi 1 ·<br />

Giacomo Bergamini 2 · Vincenzo Balzani 2<br />

1Dipartimento di Chimica Inorganica, Chimica Analitica e Chimica Fisica,<br />

Università di Messina, Via Sperone 31, 98166 Messina, Italy<br />

campagna@unime.it<br />

2Dipartimento di Chimica “G. Ciamician”, Università di Bologna, Via Selmi 2,<br />

40126 Bologna, Italy<br />

1 Introduction .................................. 118<br />

2 Structure, Bonding, <strong>and</strong> Excited States <strong>of</strong> Ru(II) Polypyridine Complexes 119<br />

3 [Ru(bpy)3] 2+ : The Prototype ......................... 123<br />

3.1 AbsorptionSpectrum ............................. 123<br />

3.2 Deactivation<strong>of</strong>UpperExcitedStates..................... 124<br />

3.3 EmissionProperties .............................. 124<br />

3.4 Photosubstitution<strong>and</strong>PhotoracemizationProcesses ............ 127<br />

3.5 Quenching <strong>of</strong> the 3 MLCT Excited State:<br />

Energy<strong>and</strong>ElectronTransferProcesses ................... 128<br />

3.6 Chemiluminescence<strong>and</strong>ElectrochemiluminescenceProcesses....... 131<br />

4 Some Important Features <strong>of</strong> Ru(II) Polypyridine Complexes ....... 133<br />

4.1 Nonradiative Decay Rate Constants <strong>and</strong> Emission Spectral Pr<strong>of</strong>iles<br />

<strong>of</strong>Ru(II)PolypyridineComplexes ...................... 133<br />

4.2 Ultrafast Time-Resolved Spectroscopy<br />

<strong>and</strong>Localization/DelocalizationIssues .................... 135<br />

4.3 Ru(II)ComplexesBasedonTridentatePolypyridineLig<strong>and</strong>s........ 136<br />

4.4 Interplay Between Multiple Low-Lying MLCT States<br />

InvolvingaSinglePolypyridineLig<strong>and</strong>.................... 138<br />

5 Ruthenium <strong>and</strong> Supramolecular <strong>Photochemistry</strong> .............. 141<br />

5.1 Photoinduced Electron/Energy Transfer Across Molecular Bridges<br />

inDinuclearMetalComplexes......................... 142<br />

5.2 Photoactive Multinuclear Ruthenium Species<br />

ExhibitingParticularTopologies ....................... 153<br />

5.2.1 Racks<strong>and</strong>Grids ................................ 153<br />

5.2.2 Dendrimers................................... 155<br />

5.3 Donor–Chromophore–AcceptorTriads.................... 164<br />

5.4 PolyadsBasedonOligoprolineAssemblies.................. 170<br />

5.5 Multi-ruthenium Assemblies Based on Derivatized Polystyrene . . . . . . 172<br />

5.6 Photoinduced Collection <strong>of</strong> Electrons<br />

intoaSingleSite<strong>of</strong>aMetalComplex..................... 174<br />

5.7 PhotoinducedMultiholeStorage:MixedRu–MnComplexes ........ 177<br />

5.8 PhotocatalyticProcessesOperatedbySupramolecularSpecies....... 180


118 S. Campagna et al.<br />

5.8.1 Photogeneration<strong>of</strong>Hydrogen......................... 180<br />

5.8.2 OtherPhotocatalyticSystems ......................... 182<br />

5.9 Photoactive Molecular Machines Able to Perform Nuclear Motions . . . . 183<br />

6 Ruthenium Complexes <strong>and</strong> Biological Systems ............... 185<br />

7 Dye-Sensitized Photoelectrochemical Solar Cells .............. 188<br />

7.1 GeneralConcepts................................ 188<br />

7.2 Ruthenium-SensitizedPhotoelectrochemicalSolarCells .......... 191<br />

7.3 SupramolecularSensitizers .......................... 193<br />

8 Miscellanea ................................... 196<br />

References ....................................... 200<br />

Abstract Ruthenium compounds, particularly Ru(II) polypyridine complexes, are the<br />

class <strong>of</strong> transition metal complexes which has been most deeply investigated from<br />

a photochemical viewpoint. The reason for such great interest stems from a unique<br />

combination <strong>of</strong> chemical stability, redox properties, excited-state reactivity, luminescence<br />

emission, <strong>and</strong> excited-state lifetime. Ruthenium polypyridine complexes are indeed good<br />

visible light absorbers, feature relatively intense <strong>and</strong> long-lived luminescence, <strong>and</strong> can<br />

undergo reversible redox processes in both the ground <strong>and</strong> excited states. This chapter<br />

presents some general concepts on the photochemical properties <strong>of</strong> Ru(II) polypyridine<br />

complexes <strong>and</strong> gives an overview <strong>of</strong> various research topics involving ruthenium photochemistry<br />

which have emerged in the last 15 years. In particular, aspects connected to<br />

supramolecular photochemistry <strong>and</strong> photophysics are discussed, such as multicomponent<br />

systems for light harvesting <strong>and</strong> photoinduced charge separation, systems for photoinduced<br />

multielectron/hole storage, <strong>and</strong> photocatalytic processes based on supramolecular<br />

Ru(II) polypyridine species. Interaction with biological systems <strong>and</strong> dye-sensitized photoelectrochemical<br />

cells are also briefly discussed.<br />

Keywords Ruthenium · Luminescence · Electron transfer · Energy transfer ·<br />

Solar energy conversion · Light-powered molecular machines · Dye-sensitized solar cells<br />

1<br />

Introduction<br />

The photochemistry <strong>of</strong> ruthenium complexes has undergone an impressive<br />

growth in the last few decades. The prototype compound [Ru(bpy)3] 2+ (bpy<br />

=2,2 ′ -bipyridine) has certainly been one <strong>of</strong> the molecules most extensively<br />

studied <strong>and</strong> widely used in research laboratories during the last 30 years.<br />

A unique combination <strong>of</strong> chemical stability, redox properties, excited-state reactivity,<br />

luminescence emission, <strong>and</strong> excited-state lifetime has attracted the<br />

attention <strong>of</strong> many researchers, first on this molecule <strong>and</strong> then on some hundreds<br />

<strong>of</strong> its derivatives. The study <strong>of</strong> this class <strong>of</strong> complexes has stimulated the<br />

growth <strong>of</strong> several branches <strong>of</strong> chemistry. In particular, Ru(II) polypyridine<br />

complexes have played <strong>and</strong> are still playing a key role in the development <strong>of</strong>


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 119<br />

photochemistry, photophysics, photocatalysis, electrochemistry, photoelectrochemistry,<br />

chemi- <strong>and</strong> electrochemiluminescence, <strong>and</strong> electron <strong>and</strong> energy<br />

transfer. Mostly in the last 15 years, Ru(II) polypyridine complexes have<br />

also contributed highly to the development <strong>of</strong> supramolecular photochemistry,<br />

<strong>and</strong> in particular to its aspects related to photoinduced electron <strong>and</strong> energy<br />

transfer processes within multicomponent (supramolecular) assemblies,<br />

including luminescent polynuclear metal complexes, light-active dendrimers,<br />

artificial light-harvesting antennae, photoinduced charge-separation devices,<br />

luminescent sensors, <strong>and</strong> light-powered molecular machines.<br />

Because <strong>of</strong> the enormous number <strong>of</strong> Ru(II) complexes investigated from<br />

a photochemical viewpoint <strong>and</strong> the variety <strong>of</strong> multicomponent structures<br />

prepared <strong>and</strong> light-based functions explored, it is impossible to make an exhaustive<br />

review. In this chapter, we recall some basic concepts on ruthenium<br />

photochemistry <strong>and</strong> discuss in some detail a few selected topics, particularly<br />

those that have developed or emerged during the last 15 years. In this way we<br />

also hope to give an overview <strong>of</strong> some research directions which ruthenium<br />

photochemistry allows to be explored. An exhaustive review [1] published<br />

about 20 years ago collects photochemical, photophysical, <strong>and</strong> redox data <strong>of</strong><br />

several hundreds <strong>of</strong> Ru(II) polypyridine complexes. Another extensive review<br />

was published about 10 years ago [2], dealing with the luminescence<br />

properties <strong>of</strong> polynuclear transition metal complexes, most <strong>of</strong> them containing<br />

Ru(II) polypyridine subunits (interestingly, in the former review [1] less<br />

than ten polynuclear Ru complexes were reported). A review focused on the<br />

photophysical properties <strong>of</strong> Ru(II) complexes with tridentate polypyridine<br />

lig<strong>and</strong>s [3] has also been published. All these review articles contain more or<br />

less comprehensive tables <strong>of</strong> data. Enlightening articles on some basic properties<br />

<strong>of</strong> Ru(II) polypyridine complexes are also available [4–8].<br />

The very large majority <strong>of</strong> photochemical investigations on ruthenium<br />

complexes deal with Ru(II) polypyridine species. For such a reason, as also<br />

implicitly suggested above, we will limit our discussion to these species. Other<br />

photoactive compounds containing ruthenium metals, including ruthenium<br />

porphyrins, are not included in this article.<br />

2<br />

Structure, Bonding, <strong>and</strong> Excited States <strong>of</strong> Ru(II) Polypyridine Complexes<br />

Ru2+ is a d6 system <strong>and</strong> the polypyridine lig<strong>and</strong>s are usually colorless molecules<br />

possessing σ donor orbitals localized on the nitrogen atoms <strong>and</strong><br />

π donor <strong>and</strong> π∗ acceptor orbitals more or less delocalized on aromatic rings.<br />

Following a single-configuration one-electron description <strong>of</strong> the excited state<br />

in octahedral symmetry (Fig. 1a), promotion <strong>of</strong> an electron from a πM metal<br />

lig<strong>and</strong> orbitals gives rise to metal-to-lig<strong>and</strong> charge transfer<br />

orbital to the π∗ L<br />

(MLCT) excited states, whereas promotion <strong>of</strong> an electron from πM to σ∗ M or-


120 S. Campagna et al.<br />

Fig. 1 a Simplified molecular orbital diagram for Ru(II) polypyridine complexes in octahedral<br />

symmetry showing the three types <strong>of</strong> electronic transitions occurring at low<br />

energies. b Detailed representation <strong>of</strong> the MLCT transition in D3 symmetry<br />

bitals gives rise to metal-centered (MC) excited states. Lig<strong>and</strong>-centered (LC)<br />

excited states can be obtained by promoting an electron from πL to π ∗ L .All<br />

these excited states may have singlet or triplet multiplicity, although spin–<br />

orbit coupling causes large singlet–triplet mixing, particularly in MC <strong>and</strong><br />

MLCT excited states [6, 9–11].<br />

The prototype [Ru(bpy)3] 2+ (Fig. 2), as well as most <strong>of</strong> the Ru(LL)3 2+ complexes<br />

(LL = bidentate polypyridine lig<strong>and</strong>), exhibits a D3 symmetry [12].<br />

Following Orgel’s notation [13], the π ∗ orbitals may be symmetrical (χ) or<br />

Fig. 2 Molecular structural formula <strong>of</strong> [Ru(bpy)3] 2+


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 121<br />

antisymmetrical (Ψ ) with respect to rotation around the C2 axis retained by<br />

each Ru(bpy) unit. A more detailed picture <strong>of</strong> the highest occupied molecular<br />

orbitals (HOMOs) <strong>and</strong> lowest unoccupied molecular orbitals (LUMOs) is<br />

shown in Fig. 1b [14–16]. The HOMOs are πMa1(d) <strong>and</strong> πMe(d), which are<br />

mainly localized on the metal; the LUMOs are π ∗ L a2(Ψ )<strong>and</strong>π ∗ L<br />

e(Ψ ), which<br />

are mainly localized on the lig<strong>and</strong>s. The ground state <strong>of</strong> the complex is a singlet,<br />

derived from the πMe(d) 4 πMa1(d) 2 electronic configuration.<br />

According to Kasha’s rule, only the lowest excited state <strong>and</strong> the upper states<br />

that can be populated on the basis <strong>of</strong> the Boltzmann equilibrium distribution<br />

may play a role in determining the photochemical <strong>and</strong> photophysical properties.<br />

The MC excited states <strong>of</strong> d 6 octahedral complexes are strongly displaced<br />

with respect to the ground-state geometry along metal–lig<strong>and</strong> vibration coordinates<br />

[17, 18].<br />

When the lowest excited state is MC, it undergoes fast radiationless deactivation<br />

to the ground state <strong>and</strong>/or lig<strong>and</strong> dissociation reactions (Fig. 3). As<br />

a consequence, at room temperature the excited-state lifetime is very short,<br />

no luminescence emission can be observed [19], <strong>and</strong> very rarely bimolecular<br />

(or supramolecular) reactions can take place. LC <strong>and</strong> MLCT excited states<br />

are usually not strongly displaced compared to the ground-state geometry.<br />

Thus, when the lowest excited state is LC or MLCT (Fig. 3) it does not undergo<br />

fast radiationless decay to the ground state <strong>and</strong> luminescence can usually be<br />

observed. The radiative deactivation rate constant is somewhat higher for<br />

3 MLCT than for 3 LC because <strong>of</strong> the larger spin–orbit coupling effect. For this<br />

reason, the 3 LC excited states are longer lived at low temperature in a rigid<br />

matrix <strong>and</strong> the 3 MLCT excited states are more likely to exhibit luminescence<br />

at room temperature in fluid solution.<br />

Fig. 3 Schematic representation <strong>of</strong> two limiting cases for the relative positions <strong>of</strong> 3 MC <strong>and</strong><br />

3 LC (or 3 MLCT) excited states


122 S. Campagna et al.<br />

From the above discussion, it is clear that the excited-state properties <strong>of</strong><br />

a complex are related to the energy ordering <strong>of</strong> its low-energy excited states<br />

<strong>and</strong>, particularly, to the orbital nature <strong>of</strong> its lowest excited state. The energy<br />

positions <strong>of</strong> the MC, MLCT, <strong>and</strong> LC excited states depend on the lig<strong>and</strong> field<br />

strength, the redox properties <strong>of</strong> metal <strong>and</strong> lig<strong>and</strong>s, <strong>and</strong> intrinsic properties<br />

<strong>of</strong> the lig<strong>and</strong>s, respectively [1, 2, 6]. Thus, in a series <strong>of</strong> complexes <strong>of</strong> the same<br />

metal ion, the energy ordering <strong>of</strong> the various excited states, <strong>and</strong> particularly<br />

the orbital nature <strong>of</strong> the lowest excited state, can be controlled by the choice <strong>of</strong><br />

suitable lig<strong>and</strong>s [1, 2, 5, 6]. It is therefore possible to design complexes having,<br />

at least to a certain degree, desired properties.<br />

For most Ru(II) polypyridine complexes, the lowest excited state is<br />

a 3 MLCT level (or, better, a cluster [6] <strong>of</strong> closely spaced 3 MLCT levels, see<br />

later) which undergoes relatively slow radiationless transitions <strong>and</strong> thus exhibits<br />

relatively long lifetime <strong>and</strong> intense luminescence emission. Such a state<br />

is obtained by promoting an electron from a metal πM orbital to a lig<strong>and</strong> π∗ L<br />

orbital (Fig. 1). The same π∗ L orbital is usually involved in the one-electron<br />

reduction process. For a long time it has been discussed whether in homoleptic<br />

complexes the emitting 3 MLCT state is best described with a multichelate<br />

ring-delocalized orbital (Fig. 4a) or a single chelate ring-localized orbital with<br />

a small amount <strong>of</strong> interlig<strong>and</strong> interaction (Fig. 4b) [20]. This problem has<br />

been tackled with a variety <strong>of</strong> techniques on both reduced <strong>and</strong> excited complexes.<br />

Compelling evidence for “spatially isolated” [21] redox orbitals has<br />

been obtained from low-temperature cyclic voltammetry [22, 23], electron<br />

spin resonance [24], electronic absorption spectra <strong>of</strong> reduced species [25, 26],<br />

nuclear magnetic resonance [27], resonance Raman spectra [28, 29], <strong>and</strong><br />

time-resolved infrared spectroscopy [30]. In the last 10 years, with the coming<br />

into play <strong>of</strong> ultrafast spectroscopic techniques, it has also been possible to<br />

investigate the nature <strong>of</strong> the Franck–Condon state <strong>and</strong> the rate constants <strong>of</strong><br />

the localization/delocalization processes, as well as the interlig<strong>and</strong> hopping<br />

(sometimes called “r<strong>and</strong>omization <strong>of</strong> the excitation”) in the MLCT excited<br />

state. These issues will be discussed in more detail later.<br />

Fig. 4 Pictorial description <strong>of</strong> the electron promoted to the π ∗ L<br />

delocalized orbital; b single chelate ring-localized orbital<br />

orbital: a multichelate ring


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 123<br />

3<br />

[Ru(bpy)3] 2+ : The Prototype<br />

To discuss the general properties <strong>of</strong> Ru(II) polypyridine complexes, it is convenient<br />

to refer to the properties <strong>of</strong> the prototype <strong>of</strong> this class <strong>of</strong> compounds,<br />

that is, [Ru(bpy)3] 2+ .<br />

3.1<br />

Absorption Spectrum<br />

The absorption spectrum <strong>of</strong> [Ru(bpy)3] 2+ is shown in Fig. 5 along with the<br />

proposed assignments [1, 4, 6, 14–16, 31]. The b<strong>and</strong>s at 185 nm (not shown in<br />

the figure) <strong>and</strong> 285 nm have been assigned to spin-allowed LC π → π ∗ transitions<br />

by comparison with the spectrum <strong>of</strong> protonated bipyridine [32]. The<br />

two remaining intense b<strong>and</strong>s at 240 <strong>and</strong> 450 nm have been assigned to spinallowed<br />

MLCT d → π ∗ transitions. The shoulders at 322 <strong>and</strong> 344 nm might<br />

be MC transitions. In the long-wavelength tail <strong>of</strong> the absorption spectrum<br />

ashoulderispresentatabout550 nm (ε ∼ 600 M –1 cm –1 )inanethanol–<br />

methanol glass at 77 K [33]. This absorption feature is thought to be due to<br />

spin-forbidden MLCT transition(s).<br />

In spite <strong>of</strong> the presence <strong>of</strong> the heavy Ru atom, it has been established that<br />

it is reasonable to assign the electronic transitions <strong>of</strong> [Ru(bpy)3] 2+ as being<br />

due to “singlet” or “triplet” states. In particular, a singlet character ≤ 10%<br />

has been estimated [10, 34] for the lowest-lying excited states <strong>of</strong> [Ru(bpy)3] 2+ .<br />

The maximum <strong>of</strong> the 1 MLCT b<strong>and</strong> at ∼ 450 nm is slightly sensitive to solvent,<br />

suggesting an instantaneous sensing <strong>of</strong> the formation <strong>of</strong> the dipolar excitedstate<br />

[Ru 3+ (bpy)2(bpy) – ] 2+ [35].<br />

Fig. 5 Electronic absorption spectrum <strong>of</strong> [Ru(bpy)3] 2+ in alcoholic solution


124 S. Campagna et al.<br />

3.2<br />

Deactivation <strong>of</strong> Upper Excited States<br />

As mentioned in the Introduction, the upper excited states <strong>of</strong> transition<br />

metal complexes usually undergo radiationless deactivation to the lowest<br />

excited state. For [Ru(bpy)3] 2+ the lowest excited state (or, better, the cluster<br />

<strong>of</strong> lowest excited states) is relatively long livedm <strong>and</strong> its formation <strong>and</strong><br />

disappearance can be easily monitored by flash spectroscopy <strong>and</strong> luminescence<br />

decay. The absorption spectrum <strong>of</strong> the lowest excited state is shown in<br />

Fig. 6 [36–39].<br />

Fig. 6 Electronic absorption spectrum <strong>of</strong> the lowest excited state <strong>of</strong> [Ru(bpy)3] 2+ in alcoholic<br />

solution<br />

The risetime <strong>of</strong> the lowest excited state upon excitation <strong>of</strong> spin-allowed<br />

excited states was initially estimated to be ≪ 1 ns for [Ru(bpy)3] 2+ [40, 41];<br />

successively, available ultrafast spectroscopic techniques demonstrated that<br />

intersystem crossing occurs in the subpicosecond timescale (see later). The<br />

efficiency <strong>of</strong> formation <strong>of</strong> the lowest excited state, Φ( 3 MLCT) (<strong>and</strong> thus, the<br />

efficiency <strong>of</strong> intersystem crossing from the upper singlets obtained by excitation<br />

to the lowest triplet, ηisc), is essentially unity [36, 37, 42–44].<br />

3.3<br />

Emission Properties<br />

Excitation <strong>of</strong> [Ru(bpy)3] 2+ in any <strong>of</strong> its absorption b<strong>and</strong>s leads to a luminescence<br />

emission (Fig. 7) whose intensity, lifetime, <strong>and</strong> energy position<br />

are more or less temperature dependent. Detailed studies on the temperature<br />

dependence [1, 4, 6, 45–49] <strong>of</strong> the luminescence lifetime <strong>and</strong> quantum<br />

yield in the temperature range 2–70 K showed that luminescence originates<br />

from a set <strong>of</strong> three closely spaced levels (∆E, 10, <strong>and</strong> 61 cm –1 )inther-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 125<br />

Fig. 7 Emission spectrum <strong>of</strong> ∗ [Ru(bpy)3] 2+ in alcoholic solution at 77 K(solid line) <strong>and</strong><br />

at room temperature (dashed line)<br />

mal equilibrium. This cluster <strong>of</strong> luminescent, closely spaced excited states<br />

will be indicated in the following by ∗ [Ru(bpy)3] 2+ or as the 3MLCT state.<br />

∗ [Ru(bpy)3] 2+ has a substantial triplet character <strong>and</strong> a single lig<strong>and</strong> localized<br />

excitation.<br />

In rigid glass at 77 K the emission lifetime <strong>of</strong> ∗ [Ru(bpy)3] 2+ is ∼ 5 µs <strong>and</strong><br />

the emission quantum yield is ∼ 0.4 [1, 6, 8]. Taken together with the unitary<br />

intersystem crossing efficiency, these figures yield a value <strong>of</strong> ∼ 13 µs for<br />

the radiative lifetime. Values <strong>of</strong> this order <strong>of</strong> magnitude have been found for<br />

MLCT excited states <strong>of</strong> other transition metal complexes [50–53]. LC excited<br />

states <strong>of</strong> transition metal complexes usually exhibit radiative lifetimes in the<br />

millisecond range [1, 6, 49, 50, 53–60].<br />

Fig. 8 Temperature dependence <strong>of</strong> the emission lifetime <strong>of</strong> ∗ [Ru(bpy)3] 2+ in nitrile solution


126 S. Campagna et al.<br />

With increasing temperature, the emission lifetime (Fig. 8) <strong>and</strong> quantum<br />

yield decrease [1, 4, 6, 8, 32, 61–79]. This behavior may be accounted for by<br />

a stepwise term <strong>and</strong> two Arrhenius terms [1–3]:<br />

B<br />

1/τ =k0 +<br />

1+exp[C(1/T –1/TB)] + A1 exp(– ∆E1/RT)<br />

+ A2 exp(– ∆E2/RT). (1)<br />

The value <strong>of</strong> the various parameters is somewhat dependent on the nature<br />

<strong>of</strong> the solvent. In propionitrile–butyronitrile (4 : 5 v/v) the values are as<br />

follows [70, 71]: k0 = 2 × 10 5 s –1 ; B = 2.1 × 10 5 s –1 ; A1 = 5.6 × 10 5 s –1 ; ∆E1 =<br />

90 cm –1 ; A2 = 1.3 × 10 14 s –1 ; ∆E2 = 3960 cm –1 .Includedink0 are the radiative<br />

k0(r) <strong>and</strong> nonradiative k0(nr) rate constants at 84 K. The stepwise term B is<br />

due to the melting <strong>of</strong> the matrix (100–150 K) <strong>and</strong> corresponds to the coming<br />

into play <strong>of</strong> vibrations capable <strong>of</strong> facilitating radiationless deactivation [8, 71].<br />

Inthesametemperaturerangearedshift<strong>of</strong>∼ 1000 cm –1 is observed in the<br />

maximum <strong>of</strong> the emission b<strong>and</strong>, <strong>and</strong> it is mainly attributed to reorganization<br />

<strong>of</strong> solvent molecules around the excited state in fluid solution before<br />

emission takes place [8, 71]. The Arrhenius term with A1 = 5.6 × 10 5 s –1 <strong>and</strong><br />

∆E1 = 90 cm –1 is thought to correspond to the thermal equilibration with<br />

a level lying at slightly higher energy <strong>and</strong> having the same electronic nature<br />

(so it would be a fourth MLCT state [6], considering the lowest-lying MLCT<br />

state is made <strong>of</strong> three sublevels as described before). The second Arrhenius<br />

term corresponds to a thermally activated surface crossing to an upper-lying<br />

3 MC level which undergoes fast deactivation. Identification <strong>of</strong> this higher<br />

level as a 3 MC state is based upon the observed photosubstitution behavior<br />

at elevated temperatures [61], consistent with established photoreactivity<br />

patterns for d 6 metal complexes [17, 52].<br />

Experiments carried out with [Ru(bpy)3] 2+ <strong>and</strong> [Ru(bpy-d8)3] 2+ in H2O<br />

<strong>and</strong> D2O [61, 80, 81] indicate that k0(nr) is sensitive to deuteration, as expected<br />

for a weak-coupled radiationless process [6, 82–84]. By contrast, A2<br />

is insensitive to deuteration, supporting a strong-coupled (surface crossing)<br />

deactivation pathway, which may be related to the observed photosensitivity.<br />

It should be noted that the decrease in lifetime on melting has also been explained<br />

on the basis <strong>of</strong> the energy gap law because <strong>of</strong> the corresponding red<br />

shift in the emission b<strong>and</strong> [6].<br />

Finally, it should be noted that at 77 K the emission spectrum <strong>of</strong><br />

[Ru(bpy)3] 2+ , as well as that <strong>of</strong> most Ru(II) polypyridine complexes, exhibits<br />

a vibrational structure (see Fig. 7). This structure is assigned to the vibrational<br />

progression, <strong>and</strong> its energy spacing is about 1300 cm –1 ,equivalentto<br />

the C – N<strong>and</strong>C– C stretching energy <strong>of</strong> the aromatic rings, thus indicating<br />

that such stretchings are the dominant accepting modes for deactivation <strong>of</strong><br />

the 3 MLCT state.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 127<br />

3.4<br />

Photosubstitution <strong>and</strong> Photoracemization Processes<br />

Although [Ru(bpy)3] 2+ is normally considered as photochemically inert toward<br />

lig<strong>and</strong> substitution, this is not strictly true [1, 6, 14, 15]. In aqueous<br />

solution the quantum yield <strong>of</strong> [Ru(bpy)3] 2+ disappearance is in the range<br />

10 –5 –10 –3 , depending on both the pH <strong>of</strong> the solution <strong>and</strong> temperature [1].<br />

In chlorinated solvents such as CH2Cl2, the photochemistry <strong>of</strong> [Ru(bpy)3]X2<br />

(X = Cl – ,Br – ,NCS – ) is well behaved [64, 85], giving rise to [Ru(bpy)2X2] as<br />

the final product. The quantum yields are in the range 10 –1 –10 –2 .ThePF – 6 salt<br />

is photoinert. A substantial difference between aqueous <strong>and</strong> CH2Cl2 solutions<br />

is that salts <strong>of</strong> [Ru(bpy)3] 2+ are completely ion-paired in the latter medium.<br />

A detailed mechanism for the lig<strong>and</strong> photosubstitution reaction <strong>of</strong><br />

[Ru(bpy)2X2] has been proposed [6, 64] (Fig. 9). According to this mechanism,<br />

thermally activated formation <strong>of</strong> a 3 MC excited state (vide supra)<br />

leads to the cleavage <strong>of</strong> a Ru – N bond, with formation <strong>of</strong> a five-coordinate<br />

square pyramidal species. In the absence <strong>of</strong> coordinating ions, as with the<br />

PF – 6 salt, this square pyramidal species returns to [Ru(bpy)3] 2+ .Whencoordinating<br />

anions are present, as in the Cl – salt, a hexacoordinated monodentate<br />

bpy intermediate is formed. Once formed, this monodentate bpy species can<br />

undergo loss <strong>of</strong> bpy <strong>and</strong> formation <strong>of</strong> [Ru(bpy)2X2], or a “self-annealing”<br />

process (chelate ring closure), with re-formation <strong>of</strong> [Ru(bpy)3] 2+ .The“selfannealing”<br />

protective step is favored in aqueous solution, presumably because<br />

Fig. 9 Scheme <strong>of</strong> the proposed mechanism for lig<strong>and</strong> photosubstitution reactions <strong>of</strong><br />

[Ru(bpy)3]X2


128 S. Campagna et al.<br />

<strong>of</strong> stabilization <strong>of</strong> the cationic [Ru(bpy)3] 2+ species, whereas formation <strong>of</strong><br />

neutral [Ru(bpy)2X2] complexes is favored in low-polarity solvents. Photoracemization<br />

<strong>of</strong> [Ru(bpy)3] 2+ [86] also occurs with low quantum yield<br />

(2.9 × 10 –4 in water at 25 ◦ C). This process can be accounted for by a rearrangement<br />

<strong>of</strong> the square pyramidal primary photoproduct into a trigonal<br />

bipyramidal intermediate which can lead back to either the ∆ or the Λ isomer<br />

[64].<br />

Lig<strong>and</strong> photodissociation is, <strong>of</strong> course, a drawback for the use <strong>of</strong> [Ru<br />

(bpy)3] 2+ in practical applications. To avoid lig<strong>and</strong> photodissociation one<br />

should prevent population <strong>of</strong> 3 MC <strong>and</strong>/or lig<strong>and</strong> dissociation from 3 MC.<br />

Population <strong>of</strong> 3 MC can be prevented or at least reduced by: (a) addition <strong>of</strong><br />

sufficient quencher to capture 3 MLCT before surface crossing to 3 MC can<br />

occur; (b) working at low temperature; (c) increasing the energy gap between<br />

3 MLCT <strong>and</strong> 3 MC; <strong>and</strong> (d) increasing pressure [87, 88]. Lig<strong>and</strong> dissociation<br />

from 3 MC can also be reduced by (e) avoiding coordinating anions in solvent<br />

<strong>of</strong> low dielectric constant <strong>and</strong> (f) linking together the three bpy lig<strong>and</strong>s so<br />

as to form a single caging lig<strong>and</strong> which encapsulates the metal ion. Point (a)<br />

is experimentally difficult, since thermal equilibration is quite a fast process.<br />

Points (c) <strong>and</strong> (f) are particularly interesting <strong>and</strong> much effort has been<br />

made along such directions [1, 89, 90]. It should be considered that in most<br />

<strong>of</strong> the [Ru(bpy)3] 2+ derivatives, the 3 MLCT state is shifted to lower energies<br />

[1], whereas the energy <strong>of</strong> the 3 MC state usually does not change. This<br />

leads to an increased energy gap between MLCT <strong>and</strong> MC states <strong>and</strong> decreased<br />

photolability. As a consequence, photosubstitution is a minor problem in<br />

most ruthenium polypyridine complexes. It should be considered, however,<br />

that decreasing the energy <strong>of</strong> the 3 MLCT level increases the Franck–Condon<br />

factors for radiationless decay to the ground state, leading to decreased luminescence<br />

lifetimes <strong>and</strong> quantum yields. The rate <strong>of</strong> radiationless decay can<br />

be decreased by extending the delocalization <strong>of</strong> the promoted electron on<br />

suitable aromatic lig<strong>and</strong>s [78, 91, 92].<br />

Finally, it should also be noted that the photolabilization <strong>of</strong> lig<strong>and</strong>s can<br />

be a pr<strong>of</strong>itable photochemical process: for example, a synthetic route to trisheteroleptic<br />

Ru complexes involves photosubstitution <strong>of</strong> lig<strong>and</strong>s [93] <strong>and</strong><br />

photochemical, reversible lig<strong>and</strong> exchange has been proposed to be used to<br />

photoswitch the complexation activity in a ruthenium complex containing<br />

a scorpionate terpyridine lig<strong>and</strong> [94].<br />

3.5<br />

Quenching <strong>of</strong> the 3 MLCT Excited State:<br />

Energy <strong>and</strong> Electron Transfer Processes<br />

The lowest 3 MLCT excited state <strong>of</strong> [Ru(bpy)3] 2+ lives long enough to encounter<br />

other solute molecules (even when these are present at relatively low<br />

concentration) <strong>and</strong> possesses suitable properties to play the role <strong>of</strong> energy


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 129<br />

Fig. 10 Molecular quantities <strong>of</strong> [Ru(bpy)3] 2+ relevant for energy <strong>and</strong> electron transfer<br />

processes. ∗∗ [Ru(bpy)3] 2+ indicates higher-energy spin-allowed excited states <strong>and</strong><br />

∗ [Ru(bpy)3] 2+ indicates the lowest spin-forbidden excited state ( 3 MLCT). Reported potentials<br />

are in aqueous solution vs SCE<br />

donor, electron donor, or electron acceptor. As is also shown in Fig. 10, the energy<br />

available to ∗ [Ru(bpy)3] 2+ for energy transfer processes is 2.12 eV <strong>and</strong> its<br />

reduction <strong>and</strong> oxidation potentials are + 0.84 <strong>and</strong> – 0.86 V(aqueoussolution,<br />

vs SCE). It follows that ∗ [Ru(bpy)3] 2+ is at the same time a good energy donor<br />

(Eq. 2), a good electron donor (Eq. 3), <strong>and</strong> a good electron acceptor (Eq. 4):<br />

∗ [Ru(bpy)3] 2+ +Q→ [Ru(bpy)3] 2+ + ∗ Q energy transfer (2)<br />

∗ [Ru(bpy)3] 2+ +Q→ [Ru(bpy)3] 3+ +Q –<br />

∗ [Ru(bpy)3] 2+ +Q→ [Ru(bpy)3] + +Q +<br />

oxidative quenching (3)<br />

reductive quenching . (4)<br />

The direct observation <strong>of</strong> redox products represents the strongest evidence to<br />

support the occurrence <strong>of</strong> oxidative <strong>and</strong> reductive quenching mechanisms.<br />

These observations can be performed in a few cases with continuous irradiation<br />

[95, 96] <strong>and</strong> more <strong>of</strong>ten in flash photolysis experiments, because<br />

usually the redox products rapidly decay either by back electron transfer reactions<br />

to re-form the starting materials or by secondary reactions to form<br />

other products. In practice, the possibility to observe transient absorptions<br />

is related to the changes in the optical density <strong>of</strong> the solution caused by the<br />

photoreaction. Bleaching <strong>and</strong> recovering <strong>of</strong> the [Ru(bpy)3] 2+ spectrum can<br />

be used for kinetic measurements. The absorption b<strong>and</strong> at 680 nm typical <strong>of</strong><br />

[Ru(bpy)3] 3+ is too weak to detect small [Ru(bpy)3] 3+ concentrations, so that<br />

in oxidative quenching processes one is forced to use the absorption spec-


130 S. Campagna et al.<br />

trum <strong>of</strong> Q – to monitor product formation. By contrast, [Ru(bpy)3] + exhibits<br />

a strong absorption b<strong>and</strong> at 510 nm, which is particularly useful to investigate<br />

reductive quenching reactions.<br />

Following some pioneering works [96–100], literally hundreds <strong>of</strong> bimolecular<br />

excited-state reactions <strong>of</strong> [Ru(bpy)3] 2+ <strong>and</strong> <strong>of</strong> its derivatives have been<br />

studied [101]. Here we only illustrate a few examples to show that these<br />

excited-state reactions can be used for mechanistic studies as well as for potential<br />

applications <strong>of</strong> the greatest interest.<br />

The early interest in [Ru(bpy)3] 2+ photochemistry arose from the possibility<br />

<strong>of</strong> using its long-lived excited state as energy donor in energy transfer<br />

processes. Although several sensitized reactions attributed to energy transfer<br />

processes (see, e.g., [102, 103]) were later shown to proceed via electron transfer<br />

[100], there are some very interesting cases in which energy transfer has<br />

been firmly demonstrated. A clear example is the quenching <strong>of</strong> ∗ [Ru(bpy)3] 2+<br />

by [Cr(CN)6] 3– , where sensitized phosphorescence <strong>of</strong> the chromium complex<br />

has been observed both in fluid solution [104–108] <strong>and</strong> in the solid<br />

state [109–111]:<br />

∗<br />

[Ru(bpy)3] 2+ + [Cr(CN)6] 3– → [Ru(bpy)3] 2+ +( 2 Eg)[Cr(CN)6] 3– (5)<br />

( 2 Eg)[Cr(CN)6] 3– → [Cr(CN)6] 3– + hν . (6)<br />

Energy transfer from ∗ [Ru(bpy)3] 2+ to [Cr(CN)6] 3– was also used to demonstrate<br />

that the photosolvation reaction observed upon direct excitation <strong>of</strong><br />

[Cr(CN)6] 3– does not originate from the luminescent 2 Eg state <strong>of</strong> the chromium<br />

complex [104, 112].<br />

It should be pointed out that both reductive <strong>and</strong> oxidative ∗ [Ru(bpy)3] 2+<br />

electron transfer quenchings by [Cr(CN)6] 3– are thermodynamically forbidden<br />

because it is very difficult to reduce or oxidize [Cr(CN)6] 3– [108].<br />

[Cr(bpy)3] 3+ , by contrast, can be very easily reduced <strong>and</strong> with this quencher<br />

oxidative electron transfer prevails over energy transfer<br />

∗ [Ru(bpy)3] 2+ + [Cr(bpy)3] 3+<br />

k=3.3×10 9 M –1 s –1<br />

–––––––––––––––––––→[Ru(bpy)3] 3+ + [Cr(bpy)3] 2+ , (7)<br />

as is shown by the appearance <strong>of</strong> the [Cr(bpy)3] 2+ absorption spectrum in<br />

flash photolysis experiments [113]. Equation 7 converts 71% <strong>of</strong>thespectroscopic<br />

energy (2.12 eV) <strong>of</strong> the excited-state reactant into chemical energy <strong>of</strong><br />

the products. As usually happens in these simple homogeneous systems, the<br />

converted energy cannot be stored but is immediately dissipated into heat by<br />

the back electron transfer reaction:<br />

[Ru(bpy)3] 3+ + [Cr(bpy)3] 2+<br />

k=2×10 9 M –1 s –1<br />

–––––––––––––––––→[Ru(bpy)3] 2+ + [Cr(bpy)3] 3+ . (8)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 131<br />

A carefully studied example <strong>of</strong> reductive electron transfer quenching (Eq. 9)<br />

is that involving Eu2+ aq as a quencher [114, 115]:<br />

∗ [Ru(bpy)3] 2+ +Euaq 2+ k=2.8×107 M –1 s –1<br />

–––––––––––––––––––→[Ru(bpy)3] + +Euaq 3+ . (9)<br />

The difference spectrum obtained by flash photolysis after a 30-ns light<br />

pulse shows a bleaching in the region around 430 nm due to depletion <strong>of</strong><br />

[Ru(bpy)3] 2+ <strong>and</strong> an increased absorption around 500 nm due to the forma-<br />

tion <strong>of</strong> [Ru(bpy)3] + (note that both Eu 2+<br />

aq<br />

<strong>and</strong> Eu3+ aq<br />

are transparent in this<br />

spectral region). Clear kinetic evidence for reductive quenching comes from<br />

the observation that the growth <strong>of</strong> the absorption at 500 nm occurs at a rate<br />

equal to the rate <strong>of</strong> decay <strong>of</strong> the luminescence emission <strong>of</strong> ∗ [Ru(bpy)3] 2+ .As<br />

it may happen in excited-state reactions, the products <strong>of</strong> Eq. 9 have a high<br />

energy content <strong>and</strong> thus they give rise to a back electron transfer reaction<br />

[Ru(bpy)3] + +Euaq 3+ k=2.7×107 M –1 s –1<br />

–––––––––––––––––––→[Ru(bpy)3] 2+ +Euaq 2+ , (10)<br />

which can be monitored (on a longer timescale) through the recovery <strong>of</strong> the<br />

430-nm absorption or the disappearance <strong>of</strong> the 500-nm absorption.<br />

In several cases direct evidence for energy transfer quenching (i.e., sensitized<br />

luminescence or absorption spectrum <strong>of</strong> the excited acceptor) or<br />

electron transfer quenching (i.e., absorption spectrum <strong>of</strong> redox products) is<br />

difficult or even impossible to obtain for bimolecular processes. In such cases,<br />

free energy correlations <strong>of</strong> rate constants are quite useful to elucidate the reaction<br />

mechanism [108, 116–118]. As we will see later, photoinduced energy<br />

<strong>and</strong> electron transfer processes can take place very easily in suitably organized<br />

supramolecular systems.<br />

3.6<br />

Chemiluminescence <strong>and</strong> Electrochemiluminescence Processes<br />

As mentioned in the introductory chapter (Balzani et al. 2007, in this volume)<br />

[119], excited states can be generated in very exergonic electron transfer<br />

reactions. Formation <strong>of</strong> excited states can be easily demonstrated when<br />

the excited states are luminescent species. Because <strong>of</strong> its stability in the reduced<br />

<strong>and</strong> oxidized forms <strong>and</strong> the strong luminescence <strong>of</strong> its excited state,<br />

[Ru(bpy)3] 2+ is an extremely versatile reactant for a variety <strong>of</strong> chemiluminescent<br />

processes [32, 120–124].<br />

In principle, there are two ways to generate the luminescent ∗ [Ru(bpy)3] 2+<br />

excited state in chemical reactions. One way (Eq. 11) is to oxidize [Ru(bpy)3] +<br />

with a species X having reduction potential E 0 (X/X – )morepositivethan<br />

0.84 V, <strong>and</strong> another way (Eq. 12) is to reduce [Ru(bpy)3] 3+ with a species Y –


132 S. Campagna et al.<br />

whose potential E 0 (Y/Y – )ismorenegativethan–0.86 V (see also Fig. 10).<br />

[Ru(bpy)3] + +X→ ∗ [Ru(bpy)3] 2+ +X – (11)<br />

[Ru(bpy)3] 3+ +Y – → ∗ [Ru(bpy)3] 2+ +Y (12)<br />

∗<br />

[Ru(bpy)3] 2+ → [Ru(bpy)3] 2+ + hν . (13)<br />

A variety <strong>of</strong> oxidants (e.g., S2O8 2– [125, 126]) <strong>and</strong> reductants (e.g., e – aq [127],<br />

hydrazine <strong>and</strong> hydroxyl anion [128], oxalate ion [129, 130]) have been used<br />

in these chemiluminescent processes. In some cases (e.g., with OH – ), the<br />

reaction mechanism cannot be a simple outer sphere electron transfer reaction<br />

<strong>and</strong> the emitting species could be a slightly modified (on the lig<strong>and</strong>s)<br />

complex. It should also be pointed out that minor amounts <strong>of</strong> oxidizing <strong>and</strong><br />

reducing impurities are sufficient to produce luminescence in chemiluminescence<br />

<strong>and</strong> electrochemiluminescence experiments [131].<br />

The most interesting way [132] to obtain chemiluminescence from<br />

[Ru(bpy)3] 2+ solutions is probably to produce the oxidized <strong>and</strong>/or reduced<br />

form <strong>of</strong> the complex “in situ” by electrochemical methods. Three classical<br />

experiments <strong>of</strong> this type can be performed:<br />

(a) To pulse the potential applied to a working electrode between the oxidation<br />

<strong>and</strong> reduction potentials <strong>of</strong> [Ru(bpy)3] 2+ in a suitable solvent [132,<br />

133]. In such a way the reduced <strong>and</strong> oxidized forms produced in the same<br />

region <strong>of</strong> space can undergo a comproportionation reaction where enough<br />

energy is available to produce an excited state <strong>and</strong> a ground state (see also<br />

Fig. 10):<br />

[Ru(bpy)3] 2+ +e – → [Ru(bpy)3] + (14)<br />

[Ru(bpy)3] 2+ –e – → [Ru(bpy)3] 3+ (15)<br />

[Ru(bpy)3] 3+ + [Ru(bpy)3] + → ∗ [Ru(bpy)3] 2+ + [Ru(bpy)3] 2+ . (16)<br />

(b) To reduce [Ru(bpy)3] 2+ in the presence <strong>of</strong> a strong oxidant (reductive<br />

oxidation). For example, luminescence is obtained upon continuous reduction<br />

<strong>of</strong> [Ru(bpy)3] 2+ at a working electrode in the presence <strong>of</strong> S2O8 2– [125,<br />

126]. This oxidant in a first one-electron oxidation reaction generates the very<br />

powerful oxidant SO4 – that can either oxidize [Ru(bpy)3] + to ∗ [Ru(bpy)3] 2+<br />

(Eq. 18) or [Ru(bpy)3] 2+ to [Ru(bpy)3] 3+ (Eq. 19), which then reacts with<br />

[Ru(bpy)3] + (Eq. 16) to yield the luminescent excited state:<br />

[Ru(bpy)3] 2+ +e – → [Ru(bpy)3] + (14)<br />

[Ru(bpy)3] + +S2O8 2– → [Ru(bpy)3] 2+ +SO4 – +SO4 2– (17)<br />

[Ru(bpy)3] + +SO4 – → ∗ [Ru(bpy)3] 2+ +SO4 2– (18)<br />

[Ru(bpy)3] 2+ +SO4 – → [Ru(bpy)3] 3+ +SO4 2– (19)<br />

[Ru(bpy)3] + + [Ru(bpy)3] 3+ → ∗ [Ru(bpy)3] 2+ + [Ru(bpy)3] 2+ . (16)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 133<br />

(c) To oxidize [Ru(bpy)3] 2+ in the presence <strong>of</strong> a strong reductant (oxidative<br />

reduction). For example, light is generated upon continuous oxidation<br />

<strong>of</strong> [Ru(bpy)3] 2+ at a working electrode in the presence <strong>of</strong> C2O4 2– [129, 130].<br />

This reductant in a first one-electron reaction generates the strongly reducing<br />

CO2 – radical that can reduce [Ru(bpy)3] 3+ to the excited ∗ [Ru(bpy)3] 2+<br />

[Ru(bpy)3] 2+ –e – → [Ru(bpy)3] 3+ (15)<br />

[Ru(bpy)3] 3+ +C2O4 2– → [Ru(bpy)3] 2+ +CO2 +CO2 – (20)<br />

[Ru(bpy)3] 3+ +CO2 – → ∗ [Ru(bpy)3] 2+ +CO2 . (21)<br />

These chemiluminescent electron transfer reactions are quite interesting<br />

from an applicative [134–136] as well as from a theoretical viewpoint. Actually,<br />

method a is at the basis <strong>of</strong> electroluminescent materials, such as organic<br />

light-emitting diodes (OLEDs) <strong>and</strong> similar devices, which are receiving increasing<br />

interest for practical applications [137–141].<br />

4<br />

Some Important Features <strong>of</strong> Ru(II) Polypyridine Complexes<br />

4.1<br />

Nonradiative Decay Rate Constants <strong>and</strong> Emission Spectral Pr<strong>of</strong>iles<br />

<strong>of</strong> Ru(II) Polypyridine Complexes<br />

Radiationless decay from MLCT states <strong>of</strong> metal polypyridine complexes occurs<br />

with energy release into medium-frequency (polypyridyl-based) modes<br />

<strong>and</strong>, to a lower degree, low-frequency modes <strong>and</strong> solvent [4, 142–149]. Averaging<br />

the medium-frequency modes which mainly promote the transition<br />

<strong>and</strong> combining low-frequency modes, including solvent, into a single mode,<br />

treated classically, the rate constant for radiationless decay knr is predicted to<br />

follow the so-called energy gap law [150–154]. Most <strong>of</strong> the work to define this<br />

topic has been made by using Ru(II) polypyridine complexes as models; however,<br />

the approach also applies to any MLCT emitter, as largely demonstrated<br />

for Os(II) [146, 147, 155] <strong>and</strong> Re(I) polypyridine [147, 149, 156] complexes.<br />

Actually, the energy gap law can be expressed by Eq. 22, where β0 includes the<br />

vibrationally induced electronic matrix element <strong>and</strong> F(calc) is the vibrational<br />

overlap factor (the quantity 1s in Eq. 22 is used to give unitless expression):<br />

ln(knr · 1s)=lnβ0 +ln[F(calc)] . (22)<br />

In a simplified version, F(calc) can be expressed as in Eq. 23 [157]:<br />

� �<br />

– γ E0<br />

F(calc) ∝<br />

�ω<br />

� �<br />

E0<br />

γ =ln –1.<br />

SM�ω<br />

(23)


134 S. Campagna et al.<br />

In Eq. 23, the energy gap E0 is related to the energy separation <strong>of</strong> the two<br />

coupled surfaces the �ω term describes, assuming a single configurational<br />

coordinate model, the average energy <strong>of</strong> the medium-frequency vibrational<br />

(accepting) modes that couple the MLCT <strong>and</strong> ground states, that is, the vibrational<br />

spacing <strong>of</strong> the ground state, <strong>and</strong> SM is the electron-vibrational coupling<br />

constant (Huang–Rhys factor). E0, �ω, <strong>and</strong>SM are expressed in cm –1 ,aswell<br />

as ∆ν1/2, the “classical” b<strong>and</strong>width that takes into account the low-frequency<br />

modes <strong>and</strong> is present in the detailed expression for F(calc) [146, 147], not<br />

shown here. If β0, the electronic term, remains roughly constant for a series<br />

<strong>of</strong> related complexes, Eq. 22 yields a straight line with intercept ln β0 [146].<br />

It is interesting to note that the parameters E0, �ω, SM, <strong>and</strong>∆ν1/2 also define<br />

the emission spectral pr<strong>of</strong>ile, so they can be obtained by a single-mode<br />

Franck–Condon analysis <strong>of</strong> the emission spectra, using Eq. 24:<br />

5�<br />

��E0 �3 �<br />

– x�ω Sx �<br />

M<br />

I(ν)=<br />

E0 x!<br />

x=0<br />

� � � � ���<br />

2<br />

ν – E0 + x�ω<br />

exp –4ln2<br />

.<br />

∆ν1/2<br />

(24)<br />

In Eq. 24, I(ν) is the relative emission intensity at energy ν (in cm –1 ), E0 is<br />

the energy <strong>of</strong> the zero–zero transition (i.e., the energy <strong>of</strong> the emitting 3 MLCT<br />

state), �ω is the average <strong>of</strong> medium-frequency acceptor modes coupled to the<br />

MLCT transition, x is the quantum number <strong>of</strong> such an averaged mediumfrequency<br />

mode which serves as the final vibronic state (note that x is usually<br />

limited to 5), ∆ν1/2 is the half-width <strong>of</strong> the individual vibronic b<strong>and</strong>s, <strong>and</strong> SM<br />

is the Huang–Rhys factor.<br />

Application <strong>of</strong> the above equations to Ru(II) polypyridine complexes allows<br />

important information to be obtained on the excited-state properties. It<br />

should be considered, however, that Eqs. 22–24 are based on several assumptions,<br />

the most important being the following. (1) For radiationless decay, the<br />

thermal population <strong>of</strong> higher-energy excited states is neglected; when such<br />

an activated route cannot be disregarded, Eq. 22 only gives a contribution to<br />

the observed radiationless rate constant, the one related to k0 <strong>of</strong> Eq. 1. (2) The<br />

Franck–Condon analysis <strong>of</strong> the emission spectral pr<strong>of</strong>ile here shown is based<br />

on a single coordinate; when more coordinates need to be considered, fitting<br />

<strong>of</strong> the spectral pr<strong>of</strong>ile following Eq. 24 can give uncorrected parameter values.<br />

A simple refinement <strong>of</strong> Eq. 24 requires inclusion <strong>of</strong> a second (low energy)<br />

frequency acceptor mode due to solvent contributions [158–161].<br />

More sophisticated theoretical methods to analyze the emission spectra<br />

<strong>of</strong> Ru(II) complexes have been introduced [162–166]. In particular, these<br />

methods allow for the detailed characterization <strong>of</strong> the high-frequency vibronic<br />

contributions to the emission spectra <strong>and</strong> the dependence <strong>of</strong> such<br />

contributions on various factors, such as the energy gap between ground <strong>and</strong><br />

excited states. Extensive discussion can be found in the cited references.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 135<br />

Emission spectral pr<strong>of</strong>iles calculated by equations such as Eq. 24 have also<br />

been used, along with theoretical quantum mechanical expressions <strong>and</strong> experimentally<br />

determined rate constant values, to estimate the electronic coupling<br />

matrix element for such intercomponent processes for photoinduced<br />

energy transfer in dinuclear Ru(II) polypyridine complexes [160].<br />

4.2<br />

Ultrafast Time-Resolved Spectroscopy<br />

<strong>and</strong> Localization/Delocalization Issues<br />

In the last 20 years, ultrafast pump–probe spectroscopy has become accessible<br />

to several research laboratories, including research groups interested<br />

in Ru photochemistry. The possibility <strong>of</strong> investigating the excited-state dynamics<br />

at very short time delays after the excitation pulse allowed clarification<br />

<strong>of</strong> several problems <strong>and</strong> shed light on many aspects <strong>of</strong> ruthenium<br />

photochemistry. New features, sometimes unexpected, have also been revealed<br />

<strong>and</strong> new questions <strong>and</strong> research topics have emerged. For example,<br />

as discussed in other parts <strong>of</strong> this chapter, it was found that singlet MLCT<br />

states can be involved in electron transfer <strong>and</strong> energy transfer processes,<br />

even before intersystem crossing <strong>and</strong>/or thermal relaxation. This is the case<br />

<strong>of</strong> photoinduced electron injection in semiconductors [167] <strong>and</strong> energy<br />

transfer/migration between Ru subunits <strong>of</strong> large, strongly coupled dendriticshaped<br />

systems [168–170]. Fluorescence from ruthenium complexes has also<br />

been detected [171].<br />

Powered by the availability <strong>of</strong> ultrafast techniques, the long-term issue <strong>of</strong><br />

localization/delocalization <strong>of</strong> MLCT states has also been revitalized. As previously<br />

stated, the general view is that the emissive state is localized on a single<br />

lig<strong>and</strong>, even for homoleptic species. However, open questions remain concerning<br />

the nature <strong>of</strong> the Franck–Condon state <strong>and</strong> the early-time dynamics<br />

which leads to the emissive state.<br />

As for the early-time dynamics, it is largely accepted that in [Ru(bpy)3] 2+<br />

<strong>and</strong> analogous homoleptic species light excitation in the MLCT singlet manifold<br />

initially produces a Franck–Condon state that is delocalized, which is<br />

where the promoted electron is shared by all the polypyridine lig<strong>and</strong>s. Then<br />

on the timescale <strong>of</strong> tens <strong>of</strong> femtoseconds, the promoted electron becomes<br />

localized on a single lig<strong>and</strong>, due to coupling with local solvent dipoles. Intersystem<br />

crossing then takes place in about 100 fs, producing a localized<br />

triplet state. The triplet MLCT state becomes “r<strong>and</strong>omized” by interlig<strong>and</strong><br />

hopping on the timescale <strong>of</strong> 10 ps, the same scale <strong>of</strong> thermal (including vibrational<br />

<strong>and</strong> solvent reorganization) relaxation <strong>of</strong> the 3 MLCT state. This<br />

general figure is schematized in Fig. 11, <strong>and</strong> is based on results dealing with<br />

many Ru(II) polypyridine complexes, taking advantage <strong>of</strong> various experimental<br />

techniques (transient absorption anisotropy, time-resolved resonance<br />

Raman, pump–probe femtosecond transient absorption spectroscopy).


136 S. Campagna et al.<br />

However, some experimental data contrast with the scheme shown in<br />

Fig. 11. For example, a time-resolved resonance Raman study indicates that<br />

in [Ru(bpy)3] 2+ even the initial excitation is localized [172]. Moreover, the recently<br />

reported femtosecond transient absorption spectrum <strong>of</strong> [Ru(bpy)3] 2+<br />

in the UV region suggests that complete relaxation within the emitting triplet<br />

MLCT state takes several tens <strong>of</strong> picoseconds, <strong>and</strong> that r<strong>and</strong>omization <strong>of</strong><br />

triplet MLCT is complete in less than 500 fs [173]. If this latter point is correct,<br />

interlig<strong>and</strong> hopping should largely occur from nonrelaxed states, probably<br />

even partly in the singlet state. In the relaxed triplet MLCT state, interlig<strong>and</strong><br />

hopping could be slower, but it would be difficult to measure since r<strong>and</strong>omization<br />

would already have happened.<br />

Fig. 11 Picture <strong>of</strong> the early-time dynamics <strong>of</strong> light excitation in the MLCT singlet <strong>of</strong><br />

[Ru(bpy)3] 2+ . A delocalizated Franck–Condon state is formed (a), which becomes localized<br />

on a single lig<strong>and</strong> (b) <strong>and</strong> then becomes “r<strong>and</strong>omized” by interlig<strong>and</strong> hopping (c)<br />

Related to the excited-state dynamics at short times after excitation, broadb<strong>and</strong><br />

femtosecond fluorescence spectroscopy <strong>of</strong> [Ru(bpy)3] 2+ has been recently<br />

reported, as already mentioned [171]. The authors get 15 ± 10 fs as the<br />

lifetime for the singlet emission, which is centered at about 520 nm.<br />

4.3<br />

Ru(II) Complexes Based on Tridentate Polypyridine Lig<strong>and</strong>s<br />

An important family <strong>of</strong> Ru(II) polypyridine complexes is that based on<br />

tridentate lig<strong>and</strong>s, with [Ru(terpy)2] 2+ as a prototype (terpy = 2,2 ′ :6 ′ ,2 ′′ -<br />

terpyridine). The absorption, emission, <strong>and</strong> redox properties <strong>of</strong> [Ru(terpy)2] 2+<br />

are similar to those <strong>of</strong> [Ru(bpy)3] 2+ ,exceptthat[Ru(terpy)2] 2+ is essentially<br />

nonluminescent at room temperature, with a lifetime <strong>of</strong> the 3 MLCT<br />

state in degassed acetonitrile at room temperature <strong>of</strong> about 250 ps (meas-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 137<br />

ured by transient absorption spectroscopy [3]), compared with a value <strong>of</strong><br />

about 1 µs exhibitedby[Ru(bpy)3] 2+ under the same conditions [1]. Such<br />

a short excited-state lifetime is very disappointing, as [Ru(terpy)2] 2+ has<br />

some advantage over [Ru(bpy)3] 2+ from a structural point <strong>of</strong> view. Whereas<br />

[Ru(bpy)3] 2+ can exist as a mixture <strong>of</strong> Λ <strong>and</strong> ∆ isomers, <strong>and</strong> the isomer<br />

problem can become even more complicated for polynuclear species based on<br />

“asymmetric” bidentate lig<strong>and</strong>s such as 2,3-bis(2 ′ -pyridyl)pyrazine (2,3-dpp),<br />

[Ru(bpy)3] 2+ is achiral. Moreover, by taking advantage <strong>of</strong> para substituents<br />

on the central pyridine <strong>of</strong> the terpy lig<strong>and</strong>, [Ru(terpy)2] 2+ can give rise to<br />

supramolecular architectures perfectly characterized from a structural viewpoint,<br />

in particular to multinuclear one-dimensional (“wire”-like) species.<br />

The reason for the poor photophysical properties <strong>of</strong> Ru(II) complexes with<br />

tridentate polypyridine lig<strong>and</strong>s at room temperature, compared to Ru(II)<br />

species with bidentate chelating polypyridine, stems from the bite angle <strong>of</strong><br />

the tridentate lig<strong>and</strong> that leads to a weaker lig<strong>and</strong> field strength <strong>and</strong> thus to<br />

lower-energy MC states as compared to Ru(II) complexes <strong>of</strong> bpy. The thermally<br />

activated process from the potentially emitting 3 MLCT state to the<br />

higher-lying 3 MC state is therefore more efficient in [Ru(terpy)2] 2+ <strong>and</strong> its<br />

derivatives <strong>and</strong> leads to fast deactivation <strong>of</strong> the excited state by nonradiative<br />

processes [1, 3, 4], although terpy-type Ru complexes are inherently more<br />

photostable than bpy-type ones because <strong>of</strong> a stronger chelating effect.<br />

Much effort has been devoted to the design <strong>and</strong> synthesis <strong>of</strong> tridentate<br />

polypyridine lig<strong>and</strong>s, leading to Ru(II) complexes with improved photophysical<br />

properties [3, 78, 92, 174–181]. For example, the use <strong>of</strong> lig<strong>and</strong>s containing<br />

electron-withdrawing <strong>and</strong> -donor substituents on tpy increases the<br />

gap between the 3 MLCT <strong>and</strong> the 3 MC states [174]. An increase in such<br />

an energy gap has also been obtained by the use <strong>of</strong> cyclometallating lig<strong>and</strong>s<br />

[177]. Unavoidably, the stabilization <strong>of</strong> 3 MLCT states causes an increase<br />

<strong>of</strong> the rate constant for radiationless decay to the ground state. This latter<br />

effect can be balanced by extension <strong>of</strong> the π ∗ orbital by appropriate substituents,<br />

which increases the delocalization <strong>of</strong> the acceptor lig<strong>and</strong> <strong>of</strong> the<br />

MLCT excited state leading to a smaller Franck–Condon factor for nonradiative<br />

decay [78, 175, 176, 178, 179, 182–188]. In this regard, species based on<br />

ethynyl-substituted terpy lig<strong>and</strong>s feature particularly interesting photophysical<br />

properties [175, 176, 182]. Various approaches to improve the photophysical<br />

properties <strong>of</strong> Ru(II) complexes with tridentate polypyridine lig<strong>and</strong>s have<br />

been reviewed [175, 182].<br />

The bis-tridentate Ru(II) polypyridine complex with the best photophysical<br />

properties reported up to now is probably the species 1, based on<br />

the 2,6-bis(8 ′ -quinolinyl)pyridine lig<strong>and</strong> [189]. This species exhibits 3 MLCT<br />

emission with a maximum at 700 nm, with a lifetime <strong>of</strong> 3.0 µs <strong>and</strong> a quantum<br />

yield <strong>of</strong> 0.02 in deoxygenated methanol–ethanol solution at room temperature.<br />

The emission maximum blue-shifts to 673 nm at 77 Kinthesame<br />

solvent mixture, exhibiting a luminescence lifetime <strong>of</strong> 8.5 µs <strong>and</strong> a quantum


138 S. Campagna et al.<br />

yield <strong>of</strong> 0.06. The authors attribute these excellent (particularly at room temperature)<br />

photophysical properties to the relief <strong>of</strong> structural distortion from<br />

the ideal octahedral geometry, due to lig<strong>and</strong> design. Actually, X-ray characterization<br />

<strong>of</strong> the compound reveals a quasi-ideal octahedral geometry around<br />

the metal center.<br />

4.4<br />

Interplay Between Multiple Low-Lying MLCT States<br />

Involving a Single Polypyridine Lig<strong>and</strong><br />

Usually, there is a linear relationship between redox data, namely first oxidation<br />

<strong>and</strong> reduction potentials <strong>of</strong> Ru complexes, <strong>and</strong> spectroscopic parameters<br />

such as MLCT absorption <strong>and</strong> emission b<strong>and</strong>s, provided that the considered<br />

compounds constitute a homogeneous series [1, 4, 190–192]. This relationship<br />

is based on the fact that the orbitals involved in metal-based oxidation<br />

<strong>and</strong> lig<strong>and</strong>-based reduction processes are the same (to a first approximation)<br />

as those involved in the MLCT absorption <strong>and</strong> emission transitions.<br />

Differences in solvent effects for redox <strong>and</strong> spectroscopic processes should<br />

be constant, so that the relationship is still linear, although the slops are not<br />

unitary [1, 191]. However, whereas until 20 years ago this relationship was followed<br />

by almost all the Ru complexes reported at that time, <strong>and</strong> exceptions<br />

were rare [193] <strong>and</strong> partly unexplained, Ru complexes which do not follow<br />

the rule, in particular as far as the absorption spectra are concerned, have<br />

become quite common in recent years. The availability <strong>of</strong> several examples allowed<br />

the development <strong>of</strong> a general interpretation <strong>of</strong> this behavior. In all the<br />

cases that do not obey the linear relationship, lig<strong>and</strong>s characterized by a large<br />

aromatic framework are present.<br />

It is now clear that the apparent mismatched relationship is linked to the<br />

presence <strong>of</strong> multiple low-energy MLCT transitions to a single polypyridine<br />

lig<strong>and</strong>, with one such transition being essentially almost invisible spectroscopically.<br />

The key feature here is that the “single” polypyridine lig<strong>and</strong> can<br />

actually be viewed as being made <strong>of</strong> two “separated” subunits (Fig. 12) with<br />

the LUMO centered on a part <strong>of</strong> the lig<strong>and</strong> framework which is not significantly<br />

coupled with the metal-based HOMOs (in other words, the LUMO does


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 139<br />

Fig. 12 Structural formula <strong>of</strong> two possible MLCT transitions in ruthenium complexes with<br />

large polypyridine lig<strong>and</strong>s<br />

not receive a significant contribution from the chelating nitrogens). In these<br />

systems, the lowest-energy MLCT transition (MLCT0)canhavevanishingoscillator<br />

strength <strong>and</strong> thus it does not significantly contribute to the absorption<br />

spectrum. On the contrary, a closely lying LUMO+1 (centered on a different<br />

moiety <strong>of</strong> the large lig<strong>and</strong>) receives a significant contribution from the<br />

chelating nitrogens, so it is largely coupled with the metal-based HOMO(s);<br />

as a consequence, its corresponding MLCT transition (the MLCT1 transition)<br />

dominates the absorption spectrum. Since reduction takes place in the<br />

LUMO, the linear relationship between absorption spectra <strong>and</strong> redox potential<br />

cannot be followed. This case will also be discussed in Sects. 5 <strong>and</strong> 6, for<br />

specific systems.<br />

As far as the relationship between emission spectra <strong>and</strong> redox potentials is<br />

concerned, whether it is followed or not depends on how fast the interconversion<br />

between the MLCT1 <strong>and</strong> MLCT0 states is, compared to the intrinsic decay<br />

<strong>of</strong> the MLCT1 excited state (here it is assumed that MLCT0 is lower in energy<br />

than MLCT1; otherwise, the relationship is always followed, except for very<br />

particular cases). Solvent, temperature, driving force, <strong>and</strong> medium effects are<br />

very important in this regard. For example, at room temperature in fluid


140 S. Campagna et al.<br />

solution the mononuclear complex [(phen)2Ru(tpphz)] 2+ (2, phen = 1,10phenanthroline;<br />

tpphz = tetrapyrido[3,2-a:2 ′ ,3 ′ -c:3 ′′ ,2 ′′ -h:2 ′′ ,3 ′′ -j]phenazine)<br />

exhibits emission from its MLCT1 level (λmax = 625 nm, τ = 1.25 ms, Φ =<br />

0.07), while the dinuclear species [(phen)2Ru(tpphz)Ru(phen)2] 4+ (3) emits<br />

from its MLCT0 level (λmax = 710 nm, τ = 0.100 ms, Φ = 0.005) [194]. The<br />

absorption spectra <strong>of</strong> both compounds in the visible region are very similar<br />

to one another (apart from the intensity), with the lowest-energy MLCT<br />

b<strong>and</strong> maximizing at about 440 nm in both cases. In a rigid matrix at 77 K,<br />

both the mononuclear <strong>and</strong> dinuclear metal complexes exhibit emission at<br />

about 585 nm (lifetime in the microsecond timescale), typical <strong>of</strong> the MLCT1<br />

level. Such results are interpreted on considering that the lig<strong>and</strong> tpphz has<br />

two empty orbitals close in energy: the LUMO is centered on the central<br />

pyrazine, with negligible contribution from the chelating nitrogen atoms, <strong>and</strong><br />

the LUMO+1 is essentially a bpy-type orbital. Reduction potential data <strong>of</strong><br />

the complexes indicate that LUMO+1 <strong>of</strong> tpphz is hardly affected on passing<br />

from mononuclear to dinuclear species, whereas the LUMO <strong>of</strong> tpphz is stabilized.<br />

For both [(phen)2Ru(tpphz)] 2+ <strong>and</strong> [(phen)2Ru(tpphz)Ru(phen)2] 4+ ,<br />

the absorption spectrum is dominated by Ru-to-tpphzLUMO+1 charge transfer<br />

(i.e., MLCT1) transition—almost coincident to the Ru-to-phen charge transfer<br />

transition—which occurs at roughly the same energy in mononuclear <strong>and</strong><br />

dinuclear species, with the Ru-to-tpphzLUMO charge transfer (i.e., MLCT0)<br />

transition not contributing to the absorption feature. Because <strong>of</strong> the different<br />

stabilization <strong>of</strong> MLCT1 <strong>and</strong> MLCT0 on passing from mononuclear to<br />

dinuclear species (see above), the driving force for the MLCT1-to-MLCT0 interconversion<br />

is more favorable, <strong>and</strong> therefore faster, in the dinuclear species.<br />

As a consequence, MLCT1-to-MLCT0 decay does not compete with the direct<br />

decay <strong>of</strong> MLCT1 to the ground state in the mononuclear species, whereas it is


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 141<br />

faster <strong>and</strong> efficient in the dinuclear species. However, at 77 KtheMLCT1-to-<br />

MLCT0 interconversion, which cannot occur without solvent reorganization,<br />

becomes inefficient in both systems. Essentially the same experimental behavior<br />

is featured by many other Ru(II) polypyridine complexes, <strong>and</strong> can be<br />

interpreted in a similar way [195–205].<br />

Time-resolved transient absorption spectroscopy [206, 207] confirmed<br />

that the MLCT1-to-MLCT0 excited-state conversion in [(phen)2Ru(tpphz)-<br />

Ru(phen)2] 4+ at room temperature is solvent dependent. Indeed, it was 120 ps<br />

in dichloromethane <strong>and</strong> faster than 40 ps in acetonitrile [206, 207]. The solvent<br />

dependence can be attributed to the difference in the MLCT1/MLCT0<br />

energy gap in the two solvents. It has to be considered, in fact, that the<br />

“charge separation” between donor <strong>and</strong> acceptor orbitals, <strong>and</strong>, as a consequence,<br />

the coulombic stabilization, is quite different in the two types <strong>of</strong><br />

MLCT states. A nonnegligible reorganization energy is therefore expected for<br />

the MLCT1-to-MLCT0 transition. A detailed study <strong>of</strong> the temperature dependence<br />

<strong>of</strong> the luminescence properties <strong>of</strong> species exhibiting this interesting<br />

behavior would be quite useful, but to our knowledge it has not yet been<br />

reported.<br />

5<br />

Ruthenium <strong>and</strong> Supramolecular <strong>Photochemistry</strong><br />

Supramolecular photochemistry has played a prominent role in chemical<br />

research since its definition in the late 1980s [208, 209]. The operational<br />

definition <strong>of</strong> supramolecular species is discussed (Balzani et al. 2007, in this<br />

volume) [119], so it will not be further commented on here. Since Ru(II)<br />

polypyridine complexes exhibit very interesting photochemical properties<br />

<strong>and</strong> can be prepared by relatively easy synthetic methods, even with madeto-order<br />

properties, the number <strong>of</strong> photoactive supramolecular species based<br />

on Ru(II) complexes has rapidly become extraordinarily large. Supramolecular<br />

systems in which donor <strong>and</strong> acceptor units are placed at designed<br />

distances can undergo photoinduced energy <strong>and</strong> electron transfer process<br />

(first-order kinetics) even in the case <strong>of</strong> short-lived excited states [208,<br />

209].<br />

Indeed, Ru(II) polypyridine compounds have been extensively used as<br />

photoactive units in supramolecular systems either exclusively made <strong>of</strong><br />

metal-based components, such as molecular racks, grids, <strong>and</strong> dendrimers, or<br />

in systems whose other active components <strong>of</strong> the assemblies are <strong>of</strong> an organic<br />

nature. In both cases, the final goals <strong>of</strong> the supramolecular systems are<br />

essentially two, reflecting the nature <strong>of</strong> the whole <strong>of</strong> photochemical science:<br />

(1) systems designed for the conversion <strong>of</strong> light energy into other forms <strong>of</strong><br />

energy, essentially chemical energy or electricity; <strong>and</strong> (2) systems focused on<br />

the elaboration <strong>of</strong> the information, including sensors. Quite <strong>of</strong>ten these two


142 S. Campagna et al.<br />

aspects are intertwined; for example, long-range photoinduced electron <strong>and</strong><br />

energy transfer processes are important both for the elaboration <strong>of</strong> optical<br />

information signals <strong>and</strong> for light-harvesting systems.<br />

5.1<br />

Photoinduced Electron/Energy Transfer Across Molecular Bridges<br />

in Dinuclear Metal Complexes<br />

Dinuclear metal complexes containing Ru(II) polypyridine subunits, where<br />

the metal centers are separated by molecular components (bridges), are particularly<br />

suited to investigating photoinduced electron <strong>and</strong> energy transfer<br />

processes, whose rate constants can give information on the electronic coupling<br />

mediated by the bridge. The latter topic has been recently reviewed <strong>and</strong><br />

deeply discussed [210]. The number <strong>of</strong> photoactive (usually, luminescent)<br />

dinuclear metal complexes based on Ru(II) subunits is very large. The last<br />

exhaustive review dealing with such species was published about 10 years<br />

ago [2]. Today it is impossible to be exhaustive even in this relatively narrow<br />

field. Therefore, we will only present a few examples. In most cases, Ru(II)<br />

subunits, which play the role <strong>of</strong> donors, are coupled to Os(II) units, which<br />

play the role <strong>of</strong> acceptors in photoinduced energy transfer processes. In all<br />

cases, the dinuclear homometallic Ru(II) species have also been investigated<br />

for comparison purposes. Their photophysical properties can be found in the<br />

original references.<br />

It is important to note that to have control <strong>of</strong> the distance between the<br />

metal centers, the bridges have to be rigid as much as possible. Therefore,<br />

it is not surprising that oligophenylenes have <strong>of</strong>ten been employed. In the<br />

series <strong>of</strong> dinuclear dyads 4 [211], having the general formula [Ru(bpy)3] 2+ -<br />

(ph)n-(R2ph)-(ph)n-[Os(bpy)3] 2+ (ph = 1,4-phenylene; n = 1, 2, 3), excitation<br />

<strong>of</strong> the [Ru(bpy)3] 2+ unit is followed by energy transfer to the [Os(bpy)3] 2+<br />

unit, as shown by the sensitized emission <strong>of</strong> the latter. For the compound with<br />

n = 3, with a total <strong>of</strong> seven phenylene spacers, the rate constant ken for energy<br />

transfer over the 4.2-nm metal-to-metal distance is 1.3 × 10 6 s –1 in acetonitrile<br />

solution at room temperature. This was probably the first example <strong>of</strong>


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 143<br />

a systematic study on the distance dependence <strong>of</strong> energy transfer rates for<br />

Ru(II)–Os(II) dyads. A Dexter-type mechanism for the Ru(II)–Os(II) energy<br />

transfer was proposed, <strong>and</strong> an attenuation factor β <strong>of</strong> 0.32 ˚A –1 for photoinduced<br />

energy transfer was obtained from the ln ken vs metal–metal distance<br />

plot.<br />

In the [Ru(bpy)3] 2+ -(ph)n-[Os(bpy)3] 3+ compounds, obtained by chemical<br />

oxidation <strong>of</strong> the Os-based moiety, photoexcitation <strong>of</strong> the [Ru(bpy)3] 2+<br />

unit causes the transfer <strong>of</strong> an electron to the Os-based one with a rate<br />

constant (kel) <strong>of</strong>3.4 × 10 7 s –1 for n = 3. Unless the electron added to the<br />

[Os(bpy)3] 3+ unit is rapidly removed, a back electron transfer reaction (rate<br />

constant 2.7 × 10 5 s –1 for n = 3) takes place from the [Os(bpy)3] 2+ unit to the<br />

[Ru(bpy)3] 3+ one [211]. The rate constants <strong>of</strong> all the transfer processes in<br />

the series <strong>of</strong> complexes decrease, as expected, with decreasing length <strong>of</strong> the<br />

oligophenylene spacer, whereas they were practically unaffected by temperature.<br />

Interestingly, a series <strong>of</strong> analogous dyads missing the central substituted<br />

phenylene (5) was successively prepared [212, 213] <strong>and</strong> the results <strong>of</strong> the<br />

two series have been compared: for the dyads containing bridges made <strong>of</strong><br />

a total <strong>of</strong> three <strong>and</strong> five phenylene spacers, the rate constants <strong>of</strong> photoinduced<br />

energy transfer are higher in the nonsubstituted phenyl series. This was attributed<br />

to effects <strong>of</strong> inter-phenylene twist angle on the electronic coupling<br />

between donor <strong>and</strong> acceptor subunits [213].<br />

Thepresence<strong>of</strong>meta substitution in oligophenylene bridges versus the<br />

all-para systems have been evidenced by the dyads 6 [210, 213]. In these<br />

species, the photoinduced energy transfer rate constants are lower than<br />

the rates for the respective compounds containing all-para phenylene units:<br />

for example, for spacers made <strong>of</strong> three <strong>and</strong> five phenylenes, respectively,<br />

ken is 1.32 × 10 9 <strong>and</strong> 6.67 × 10 7 s –1 for the meta series <strong>and</strong> 2.77 × 10 10 <strong>and</strong><br />

4.90 × 10 8 s –1 for the para series. A related result has been reported for the<br />

tetranuclear Ir(III)/Ru(II) mixed-metal species 7 (although not linear, this<br />

species is briefly discussed here for convenience reasons) [214]. In 7,bothIrbased<br />

emission (λ = 572 nm, τ = 2.9 µs) <strong>and</strong> Ru-based emission (λ = 682 nm,<br />

τ = 82 ns) are present, showing that photoinduced energy transfer from the


144 S. Campagna et al.<br />

Ir(III) chromophores to the Ru(II) units is inefficient at room temperature in<br />

fluid solution. This suggests that the Ir-to-Ru photoinduced energy transfer<br />

rate constant in this tetranuclear species is lower than the intrinsic rate constant<br />

for Ir decay (about 3.7 × 10 5 s –1 ), in spite <strong>of</strong> the nonnegligible driving<br />

force (about 0.3 eV, from emission data). At 77 K, energy transfer from Irbased<br />

to Ru-based chromophores is quantitative because <strong>of</strong> the much longer<br />

lifetime (205 µs) <strong>of</strong> the excited state <strong>of</strong> the Ir-based units. Indirectly, the<br />

room- <strong>and</strong> low-temperature results tend to suggest that Ir-to-Ru energy transfer<br />

in the tetranuclear mixed-metal species would occur with a rate constant<br />

<strong>of</strong> the order <strong>of</strong> 10 4 s –1 . The apparent discrepancy with the relatively fast energy<br />

transfer rate constant for the Ru–Os species with three phenylene unit<br />

bridges <strong>of</strong> the meta series discussed above (having a similar bridge to the<br />

Ir/Ru tetranuclear system here discussed) shows that the energy transfer


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 145<br />

rates are significantly affected by the partner properties, as expected for both<br />

coulombic <strong>and</strong> (superexchange-assisted) Dexter-type mechanisms.<br />

In the series <strong>of</strong> Ru–Os dyads with tridentate lig<strong>and</strong>s 8 [215–217], both<br />

the donor Ru-based <strong>and</strong> acceptor Os-based MLCT states taking part in the<br />

energy transfer process involve the bridging lig<strong>and</strong>, while in the analogous,<br />

cyclometallated species 9 [215, 218] the donor <strong>and</strong> acceptor MLCT states are<br />

based on the peripheral lig<strong>and</strong>s. In the larger systems, with two phenylene<br />

spacers, ken for energy transfer from the Ru-based chromophore to the Osbased<br />

one is 5 × 10 10 s –1 for the non-cyclometallated species [215–217] <strong>and</strong><br />

< 2 × 10 7 s –1 for the cyclometallated species [215, 218]. These different results<br />

are mainly attributed to the fact that the energy transfer pathway is longer for<br />

the cyclometallated system, although it is suggested that the different nature<br />

<strong>of</strong> the bridge could also play a role.<br />

Possible effects <strong>of</strong> excited-state localization on intramolecular energy<br />

transfer kinetics are also shown by the results obtained for the two isomeric<br />

dinuclear Ru(II) species 10 <strong>and</strong> 11 [219]. In both complexes, energy transfer<br />

from the non-cyclometallated Ru subunit to the cyclometallated Ru subunit<br />

takes place by a Dexter mechanism. Ultrafast spectroscopic measurements<br />

yield different energy transfer time constants for the two isomers, with that<br />

related to the bridge-cyclometallated complex (2.7 ps) being faster than that<br />

related to the terminal-cyclometallated one (8.0 ps). This difference is explained<br />

in terms <strong>of</strong> different electronic factors for Dexter energy transfer. The<br />

lowest MLCT excited state in the Ru cyclometallated unit <strong>of</strong> the dinuclear


146 S. Campagna et al.<br />

complexes (that is, the acceptor state <strong>of</strong> the energy transfer process) has the<br />

promoted electron on the non-cyclometallating lig<strong>and</strong>, i.e., on the bridging<br />

lig<strong>and</strong> for 10 <strong>and</strong> on the terminal lig<strong>and</strong> for 11. The lowest MLCT excited state<br />

<strong>of</strong> the non-cyclometallated Ru(II) subunit (that is, the donor state <strong>of</strong> the energy<br />

transfer process) has the promoted electron on the bridging lig<strong>and</strong> in<br />

both cases. The energy transfer process, in a Dexter mechanism, is equivalent<br />

to simultaneous electron–hole transfer between the molecular components.<br />

The hole transfer process is the same (metal-to-metal) for both isomers. The<br />

electron transfer, on the other h<strong>and</strong>, is different for the two isomers, taking<br />

place between the two halves <strong>of</strong> the bridging lig<strong>and</strong> for 10 <strong>and</strong> from the<br />

bridging lig<strong>and</strong> <strong>of</strong> one unit to the terminal lig<strong>and</strong> <strong>of</strong> the other unit in the<br />

case <strong>of</strong> 11. The exchange electronic coupling is clearly expected to be higher<br />

in the former than in the latter case. Interestingly, in both cases the energy<br />

transfer processes also have a slower component <strong>of</strong> about 40 ps, which was<br />

tentatively assigned to roughly isoenergetic electron hopping between terminal<br />

<strong>and</strong> bridging lig<strong>and</strong>s in the non-cyclometallated Ru chromophore, in<br />

agreement with reported rate constants for isoenergetic electron hopping in<br />

Ru(II) polypyridine complexes, as discussed in Sect. 4.2. This study clearly<br />

highlights the peculiar intricacies <strong>of</strong> intramolecular energy transfer in inorganic<br />

dyads involving MLCT excited states.<br />

Oligophenylene bridges have also been employed for studying photoinduced<br />

electron transfer from Ru(II) chromophores to Rh(III) subunits. In<br />

this type <strong>of</strong> multicomponent species, electron transfer from the Ru-based<br />

MLCT state to the Rh(III) component takes place, followed by charge recombination.<br />

The compounds 12–17 are an interesting series <strong>of</strong> homologous<br />

systems [220, 221]. The rate constants <strong>of</strong> the photoinduced electron transfer<br />

processes reported confirm the distance dependence <strong>of</strong> the process, as well as<br />

the effect <strong>of</strong> the twist angle between adjacent spacers [106, 213, 222, 224, 225]<br />

on the electronic coupling (<strong>and</strong>, therefore, on the electron transfer kinetics)<br />

across the spacer.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 147


148 S. Campagna et al.<br />

Another spacer subunit which allows for rigidity <strong>and</strong> controlled directionality<br />

is the alkynyl group. Dinuclear species incorporating an unsaturated<br />

polyacetylenic backbone in the spacer (18) have been extensively investigated<br />

[175, 226]. Energy transfer (electron exchange mechanism) from the<br />

Ru(II) chromophore to the Os(II) one takes place with a rate constant <strong>of</strong><br />

7.1 × 10 10 <strong>and</strong> 5.0 × 10 10 s –1 for n =1 <strong>and</strong> n = 2, respectively [175]. The β<br />

attenuation factor [227] for the polyacetylenic systems was calculated to<br />

be 0.17 ˚A –1 , indicating that the “electron conduction” for energy transfer<br />

through alkyne bridges is more efficient than that through oligophenylenic<br />

spacers [228].<br />

Several other differently connected multicomponent species based on<br />

Ru(II) chromophores as donors <strong>and</strong> incorporating polyacetylenic bridges<br />

have been studied. Interested readers can find information in [175, 176, 226,<br />

229]. A special comment is warranted by the species 19–21 [230]. These compounds<br />

indicate the effect <strong>of</strong> the incorporation <strong>of</strong> additional units into the


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 149<br />

polyacetylenic backbone. For the phenyl-containing bridge system, the triplet<br />

state <strong>of</strong> the bridge is higher in energy than both the Ru(II) <strong>and</strong> Os(II) MLCT<br />

levels. Energy transfer takes place directly from the Ru(II) chromophore<br />

to the Os(II) one via a superexchange-assisted Dexter mechanism. In the<br />

naphthyl-bridged Ru–Os species, the triplet state <strong>of</strong> the bridge is intermediate<br />

in energy between donor <strong>and</strong> acceptor levels: the energy transfer from the<br />

Ru(II) subunit to the Os(II) one occurs in a stepwise manner, first to the central<br />

bis(alkyl)naphthalene unit <strong>of</strong> the bridge <strong>and</strong> then to the Os(II) site. In<br />

the anthryl-bridged species, the bis(alkyl)anthracene triplet is lower in energy<br />

than both Ru- <strong>and</strong> Os-based MLCT states, <strong>and</strong> the bridge plays the role<br />

<strong>of</strong> an energy trap [230].<br />

The last Ru–Os compound discussed above has some similarity with the<br />

Ru–anthracene–Os species 22 [231, 232]. In this species, missing the ethynyl<br />

groups, the anthracene triplet lies in between the Ru donor <strong>and</strong> Os acceptor<br />

energy transfer subunits, so the behavior <strong>of</strong> the bridge is similar to that <strong>of</strong><br />

the naphthyl-bridged species mentioned above. However, in air-equilibrated<br />

solution the energy transfer rate constant significantly decreases with increasing<br />

irradiation time. This effect is due to the formation <strong>of</strong> singlet oxygen by<br />

bimolecular energy transfer from the Os(II) excited state. The singlet oxygen<br />

reacts with the anthracene unit to give a peroxide species which cannot behave<br />

as the intermediate “station” for energy transfer, so that the overall process<br />

is significantly slowed down. This complex was called a “self-poisoning”<br />

species [231, 232].


150 S. Campagna et al.<br />

Another example showing an “active” role <strong>of</strong> the bridge in mediating intercomponent<br />

transfer processes involving Ru(II) species is evidenced by<br />

23 [233]. In this species, there are two close-lying MLCT states per metal<br />

center involving the bridging lig<strong>and</strong> (leaving aside the MLCT state involving<br />

the peripheral lig<strong>and</strong>s), because <strong>of</strong> the particular nature <strong>of</strong> the bridge<br />

(see Sect. 5.9). The higher energy <strong>of</strong> such MLCT states (MLCT1) involves<br />

a bridging lig<strong>and</strong> orbital mainly centered in the bpy-like coordinating site<br />

(LUMO+1), <strong>and</strong> the lower energy one (MLCT0) islocalizedonthecentral<br />

phenazine-like site (LUMO). Light excitation <strong>of</strong> the Ru-based chromophore<br />

populates the singlet MLCT1 state, which rapidly decays to its triplet counterpart.<br />

Direct light excitation into the singlet MLCT0 level (<strong>and</strong> successive<br />

population <strong>of</strong> its triplet) is inefficient because <strong>of</strong> the negligible oscillator<br />

strength <strong>of</strong> the transition. For Ru-to-Os energy transfer, two possible pathways<br />

are possible: (1) Ru-to-Os energy transfer at the 3 MLCT1 level (EnT),<br />

followed by 3 MLCT1-to- 3 MLCT0 relaxation within the Os(II) chromophore<br />

(a sort <strong>of</strong> intralig<strong>and</strong> electron transfer, ILET, within the Os(II) subunit); <strong>and</strong><br />

(2) 3 MLCT1-to- 3 MLCT0 relaxation within the Ru(II) chromophore (ILET in<br />

Ru(II) subunit), followed by Ru-to-Os energy transfer at the 3 MLCT0 level.<br />

The situation is schematized in Fig. 13 [210, 233].<br />

Interestingly, ultrafast spectroscopy shows that pathway 1 is followed in<br />

dichloromethane <strong>and</strong> pathway 2 prevails in the more polar acetonitrile solvent.<br />

Oligophenyl bridges are reported to play “active” roles in the dinuclear<br />

Ir(III)–Ru(II) species 24–27 [234]. In this series <strong>of</strong> complexes, the Ru-based<br />

component is the energy transfer acceptor subunit. Indeed, Ru-based emission<br />

takes place in all the species at about 625 nm (lifetime about 200 ns) in<br />

aerated acetonitrile at room temperature <strong>and</strong> at about 590 nm (lifetime about<br />

6 µs) in butyronitrile at 77 K, whereas the high-energy Ir-based chromophore<br />

has a very short excited-state lifetime, determined by time-resolved emission<br />

<strong>and</strong> subpicosecond transient absorption spectroscopy, slightly dependent on<br />

the bridge. The energy transfer rate constant is very weakly slowed down by<br />

increasing the bridge length, passing from 8.3 × 10 11 s –1 for the species with<br />

two phenyls as spacer to 3.3 × 10 11 s –1 for the species with five interposed<br />

phenyls. The apparent attenuation parameter β for energy transfer rate con-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 151<br />

Fig. 13 Energy transfer pathways in a dinuclear Ru(II)–Os(II) species containing a πextended<br />

bridge<br />

stant would be 0.07 ˚A –1 . However, increasing the bridge length changes the<br />

excited-state energy level <strong>of</strong> the bridge itself, <strong>and</strong> it is proposed that the MLCT<br />

excited state <strong>of</strong> the Ir center assumes an increasing LC character involving<br />

the bridge orbitals as the number <strong>of</strong> phenyls increases. As a consequence,<br />

metal–metal separation does not reflect the effective donor–acceptor separation<br />

for the energy transfer process; with the donor excited state largely<br />

involving the oligophenyl spacer, the energy transfer takes place by an in-


152 S. Campagna et al.<br />

coherent hopping mechanism, <strong>and</strong> cannot be considered a through-bond,<br />

superexchange-assisted energy transfer. Interestingly, the use <strong>of</strong> the same<br />

spacers has already been mentioned for the couple Ru/Os (see above), where<br />

a normal through-bond superexchange mechanism was proposed to be operative:<br />

the reason for the difference between Ru/Os <strong>and</strong> Ir/Ru series is the<br />

energy level <strong>of</strong> the donor excited state. In Ir(III) complexes, such a state is<br />

much higher in energy <strong>and</strong> can interact significantly with oligophenyl-based<br />

excited states.<br />

An interesting series <strong>of</strong> papers dealing with photoinduced electron transfer<br />

in a series <strong>of</strong> dinuclear Ru(II)–Co(III) species allowed the accumulation<br />

<strong>of</strong> further information on the role <strong>of</strong> the bridging lig<strong>and</strong> [207, 235, 236]. Typical<br />

studied complexes were [(terpy)Ru(terpy-terpy)Co(terpy)] 5+ , [(terpy)Ru<br />

(terpy-ph-terpy)Co(terpy)] 5+ , <strong>and</strong> [(bpy)2Ru(tpphz)Co(bpy)2] 5+ (terpy-terpy<br />

=6,6 ′ -bis(2-pyridyl)-2,2 ′ :4 ′ ,4 ′′ :2 ′′ ,2 ′′′ -quarterpyridine, that is, the bridge with<br />

n =0in5; terpy-ph-terpy is the bridge with n =1in5; tpphzisthebridgein<br />

3). In these studies, the quantum yields <strong>of</strong> the thermally equilibrated product<br />

<strong>of</strong> the photoinduced electron transfer from the Ru-based MLCT state to<br />

the Co(III) subunit were carefully measured [207]. Interestingly, ∗ Ru(II)-to-<br />

Co(III) electron transfer takes place in less than 10 ps in all three abovementioned<br />

species; however, the quantum yields <strong>of</strong> the thermally equilibrated<br />

electron transfer product were quite different in the series, with the tpphzbridged<br />

dinuclear species exhibiting a quantum yield <strong>of</strong> about 0.8 <strong>and</strong> the<br />

other two compounds featuring significantly smaller values (0.53 <strong>and</strong> 0.41 for<br />

the terpy-terpy <strong>and</strong> the terpy-ph-terpy species, respectively) in butyronitrile<br />

at 298 K [207]. The authors proposed that the relatively low yield <strong>of</strong> photoinduced<br />

electron transfer in the two terpy-based complexes is due to formation<br />

<strong>of</strong> the d–d (MC) excited state <strong>of</strong> the [Co(terpy)2] 3+ moiety during solvent


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 153<br />

relaxation within the electron transfer excited state. In fact, fast tunneling<br />

transition <strong>of</strong> the nonrelaxed electron transfer product to the lowest d–d excited<br />

states <strong>of</strong> the [Co(terpy)2] 3+ moiety can take place via hole transfer from<br />

[Ru(terpy)2] 3+ to the [Co(terpy)2] 2+ , generating a dπ 6 dσ ∗ configuration.<br />

Strong through-lig<strong>and</strong> electronic coupling <strong>of</strong> dπ(Ru)–dπ(Co), as estimated<br />

from the strong intensity <strong>of</strong> the intervalence b<strong>and</strong> <strong>of</strong> [(terpy)Ru(terpyterpy)Ru(terpy)]<br />

5+ , can effectively mediate the fast hole transfer process.<br />

For the tpphz-bridged system, through-lig<strong>and</strong> electronic coupling between<br />

dπ(Ru III ) <strong>and</strong> dπ(Co II ) orbitals is much smaller, as suggested by the absence<br />

<strong>of</strong> any sizeable intervalence b<strong>and</strong> in [(bpy)2Ru(tpphz)Ru(bpy)2] 5+ [207]. It<br />

turns out that the weak tpphz superexchange interaction between dπ(Ru III )<br />

<strong>and</strong> dπ(Co II ) orbitals may be unable to open the channel <strong>of</strong> hole transfer<br />

during the relaxation <strong>of</strong> the electron transfer product, leading to a higher<br />

quantum yield <strong>of</strong> the charge-separated, thermally equilibrated product. However,<br />

the charge recombination rate constant was fast in all cases: in butyronitrile<br />

at room temperature it was 2.1 × 10 7 s –1 for the tpphz species <strong>and</strong><br />

biphasic <strong>and</strong> faster for the other two compounds (81 × 10 9 <strong>and</strong> 5 × 10 9 s –1<br />

for the terpy-ph-terpy containing species <strong>and</strong> 52 × 10 10 <strong>and</strong> 3 × 10 10 s –1 for<br />

the terpy-terpy species). Even in the charge recombination (back electron<br />

transfer) process, the different coupling <strong>of</strong>fered by the bridging lig<strong>and</strong>s could<br />

explain the results.<br />

5.2<br />

Photoactive Multinuclear Ruthenium Species<br />

Exhibiting Particular Topologies<br />

5.2.1<br />

Racks <strong>and</strong> Grids<br />

Rack-type metal complexes are linearly arranged species [237], but differ<br />

from the species discussed in the former section since they are made <strong>of</strong><br />

several repeating, roughly identical, metal-based subunits orthogonally appended<br />

to a roughly linear <strong>and</strong> rigid polytopic molecular str<strong>and</strong>. The metal<br />

centers are never aligned along the main axis <strong>of</strong> the bridging lig<strong>and</strong>.<br />

The first rack-type Ru(II) polypyridine complex investigated from a photochemical<br />

viewpoint is 28 [238]. In this species, the anthryl group has only<br />

the function <strong>of</strong> absorbing additional light energy; in fact, its triplet state is<br />

higher in energy than the MLCT triplet state(s) <strong>of</strong> the Ru(II) subunits (here,<br />

the lowest-lying MLCT states involve the bridging lig<strong>and</strong>, which is very easily<br />

reduced). For 28, near-IR emission occurs (λem = 845 nm) with a relatively<br />

long lifetime (60 ns). Such an emission is totally quenched in the somewhat<br />

related tetranuclear Ru – Fe grid 29 [239], where energy transfer from the Rubased<br />

MLCT state to the Fe-based MC levels is likely to occur. Kinetic data for<br />

the quenching processes were not reported.


154 S. Campagna et al.<br />

Ru(II) racks based on different molecular str<strong>and</strong>s (30, 31) havealsobeen<br />

recently studied [240] (S Campagna, unpublished results). The pyrimidinecontaining<br />

Ru complex exhibits a 3 MLCT emission with maximum at 758 nm


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 155<br />

in acetonitrile at room temperature (τ = 30 ns), which moves to 740 nm<br />

in nitrile glass at 77 K(τ = 335 ns) [240]. The pyrazine-containing species<br />

exhibits very similar emission properties (room temperature, acetonitrile:<br />

λmax = 765 nm; τ = 60 ns; 77 K, nitrile glass: 750 nm; τ = 400 ns) (S Campagna,<br />

unpublished results). For these species, the lowest (emitting) MLCT<br />

state(s) involve(s) the bridging lig<strong>and</strong>s, as for the former rack-type complex<br />

discussed above.<br />

5.2.2<br />

Dendrimers<br />

Luminescent Ru(II) dendrimers have been deeply investigated <strong>and</strong> the field<br />

has been reviewed recently [2, 241–245]. We will only mention some examples.<br />

5.2.2.1<br />

Dendrimers Containing Only One Metal Center Unit<br />

The ruthenium compound 32 is a classical example <strong>of</strong> a dendrimer containing<br />

a luminescent ruthenium complex core surrounded by organic wedges.<br />

In this dendrimer, the 2,2 ′ -bipyridine (bpy) lig<strong>and</strong>s <strong>of</strong> the {Ru(bpy)3} 2+ -type<br />

core carry branches containing 1,3-dimethoxybenzene- <strong>and</strong> 2-naphthyl-type<br />

chromophoric units [246]. All three types <strong>of</strong> chromophoric groups present in<br />

the dendrimer, namely, {Ru(bpy)3} 2+ , dimethoxybenzene, <strong>and</strong> naphthalene,<br />

are potentially luminescent species. In 32, however, the fluorescence <strong>of</strong> the<br />

dimethoxybenzene- <strong>and</strong> naphthyl-type units is almost completely quenched<br />

in acetonitrile solution, with concomitant sensitization <strong>of</strong> the {Ru(bpy)3} 2+<br />

core luminescence. These results show that very efficient energy transfer processes<br />

take place, converting the very short-lived (nanosecond timescale) UV<br />

fluorescence <strong>of</strong> the aromatic units <strong>of</strong> the wedges to the relatively long-lived<br />

(microsecond timescale) orange luminescence <strong>of</strong> the metal-based dendritic<br />

core. This dendrimer is therefore an excellent example <strong>of</strong> a light-harvesting<br />

antenna system as well as <strong>of</strong> a species capable <strong>of</strong> acting as a frequency converter.<br />

It should also be noted that in aerated solution the phosphorescence<br />

intensity <strong>of</strong> the dendritic core is more than twice as intense as that <strong>of</strong> the


156 S. Campagna et al.<br />

[Ru(bpy)3] 2+ parent compound, because the dendrimer branches protect the<br />

Ru–bpy-based core from dioxygen quenching [247].<br />

More recently, dendrimers based on {Ru(phen)3} 2+ or {Ru(bpy)2(phen)} 2+<br />

(phen = 1,10-phenanthroline) cores with appended carbazole chromophoric<br />

units have been investigated (see, e.g., 33) [248]. In these compounds, energy<br />

transfer from the peripheral carbazole units to the metal-based core occurs<br />

with ca. 100% efficiency, with sensitization <strong>of</strong> the 3 MLCT luminescence at ca.<br />

630 nm.<br />

The {Ru(bpy)3} 2+ core was used to build a first-generation dendrimer<br />

containing 12 coumarin 450 units in the periphery (34). Inacetonitrilesolution,<br />

excitation <strong>of</strong> the coumarin 450 chromophores resulted in luminescence<br />

emission at 625 nm, typical <strong>of</strong> the 3 MLCT excited state <strong>of</strong> the {Ru(bpy)3} 2+<br />

core, with only a minor residual fluorescence <strong>of</strong> the initially excited chromophores.<br />

An antenna effect is thus operative, with an estimated intramolecular<br />

energy transfer efficiency close to unity [249]. Using the same dendrimer<br />

<strong>and</strong> the [Ru(dmb)3] 2+ (dmb = 4,4 ′ -methyl-2,2 ′ -bipyridine) complex, which<br />

served as a model for the “naked” dendrimer core, it was possible to investigate<br />

the shielding effect <strong>of</strong> the dendritic structure on the intermolecular energy<br />

<strong>and</strong> electron transfer processes [250]. Bimolecular quenching constants<br />

were measured in acetonitrile solution for dioxygen, 9-methylanthracene<br />

(MA), phenothiazine (PTZ), <strong>and</strong> methyl viologen (MV 2+ ), using three dif-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 157<br />

ferent techniques: intensity <strong>and</strong> lifetime quenching, <strong>and</strong> transient absorption<br />

spectroscopy. These quenchers were chosen to explore different quenching<br />

mechanisms. Whereas in the case <strong>of</strong> dioxygen the quenching mechanism is<br />

still an object <strong>of</strong> controversy, <strong>and</strong> should involve both energy <strong>and</strong> electron<br />

transfer, it is known that the [Ru(bpy)3] 2+ excited state is quenched by MA,<br />

PTZ, <strong>and</strong> MV 2+ via energy transfer, reductive <strong>and</strong> oxidative electron transfer<br />

mechanisms, respectively. The results obtained with the quenchers MA, PTZ,


158 S. Campagna et al.<br />

<strong>and</strong> MV 2+ indicated that the first-generation dendritic structure <strong>of</strong> this class<br />

<strong>of</strong> dendrimers is unable to shield the Ru-based core from bimolecular energy<br />

<strong>and</strong> electron transfer reactions. On the contrary, quenching by dioxygen was<br />

attenuated in going from [Ru(dmb)3] 2+ to larger dendrimers, suggesting an<br />

effect <strong>of</strong> the dendritic structure. As this effect is not observed with MA, PTZ,<br />

<strong>and</strong> MV 2+ , which are certainly much bulkier species, the shielding effect observedinthecase<strong>of</strong>dioxygenwasattributedtolowerO2<br />

solubility within the<br />

dendritic structure [250].<br />

5.2.2.2<br />

Multimetallic Dendrimers<br />

For the class <strong>of</strong> dendrimers where metal complexes are the branching centers,<br />

a key role is played by polytopic chelating lig<strong>and</strong>s (bridging lig<strong>and</strong>s), which<br />

can control the shape <strong>of</strong> the polynuclear array <strong>and</strong> the electronic interaction<br />

between metal chromophores.<br />

The largest family <strong>of</strong> these dendrimers is based on the 2,3-bis(2 ′ -<br />

pyridyl)pyrazine (dpp) bridging lig<strong>and</strong>. Within such a series, the largest<br />

species contain 22 metal centers [251–253]. The decanuclear compound 35 is


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 159<br />

a second-generation dendrimer <strong>of</strong> this family (in the sketch <strong>of</strong> the compound,<br />

the peripheral lig<strong>and</strong>s, schematized as NˆN, st<strong>and</strong> for 2,2 ′ -bipyridine) [254].<br />

For the dendrimers <strong>of</strong> this series, the modular synthetic strategy [252] allows<br />

a high degree <strong>of</strong> synthetic control in terms <strong>of</strong> the nature <strong>and</strong> position <strong>of</strong> metal<br />

centers, bridging lig<strong>and</strong>s, <strong>and</strong> terminal lig<strong>and</strong>s. Since the excited-state level<br />

<strong>of</strong> each metal center in the dendrimer depends on the nature <strong>of</strong> the metal, <strong>of</strong><br />

its coordination sphere (which in its turn depends on the metal position, inner<br />

or outer, within the dendritic array) <strong>and</strong> on the lig<strong>and</strong>s, each metal-based<br />

subunit is characterized by specific excited-state properties, which because <strong>of</strong><br />

the symmetry <strong>of</strong> the dendritic structure are usually identical for each metalbased<br />

subunit belonging to the same dendritic layer. Therefore, the synthetic<br />

control translates into control <strong>of</strong> specific properties, such as the direction<br />

<strong>of</strong> electronic energy flow within the dendritic array (antenna effect) [245].<br />

For example, in 35 the lowest excited-state level involves the peripheral subunit(s),<br />

<strong>and</strong> the emission <strong>of</strong> the species (acetonitrile, room temperature:<br />

λmax = 780 nm; τ = 60 ns; Φ = 3 × 10 –3 ) is assigned to a (bpy)2Ru→ µ-dpp<br />

MLCT triplet state [254]. Excitation spectroscopy indicates that quantitative<br />

energy transfer takes place from inner subunits to the peripheral ones [254].<br />

Because <strong>of</strong> the energy gradient between the dendritic layers, the energy transfer<br />

is ultrafast (see later), occurring in the femtosecond timescale. On the<br />

basis <strong>of</strong> the above discussion, it is not surprising that all the homometallic<br />

dendrimers <strong>of</strong> the same family, independent <strong>of</strong> the number <strong>of</strong> Ru subunits<br />

(i.e, tetranuclear [255, 256], decanuclear [253, 254], <strong>and</strong> docosanuclear [251–<br />

253], as well as hexanuclear [257, 258], heptanuclear [259], <strong>and</strong> tridecanuclear<br />

species [260], which have particular connections/geometries), exhibit<br />

practically identical photophysical properties, since the lowest-energy subunit(s)<br />

is in all cases the identical peripheral (bpy)2Ru(µ-dpp) MLCT state(s).<br />

A nonanuclear species has also been prepared [261], but its photophysical<br />

properties have not yet been reported.<br />

On increasing nuclearity, a unidirectional gradient (center-to-periphery<br />

or vice versa) for energy transfer is hardly obtained with only two types <strong>of</strong><br />

metals (commonly, Ru(II) <strong>and</strong> Os(II)) <strong>and</strong> lig<strong>and</strong>s (bpy <strong>and</strong> 2,3-dpp). In fact,<br />

by using a divergent synthetic approach starting from a metal-based core<br />

it becomes unavoidable that metal-based building blocks with high-energy<br />

excited states (high-energy subunits) are interposed between donor <strong>and</strong> acceptor<br />

subunits <strong>of</strong> the energy transfer processes [245]. For example, while in<br />

the tetranuclear [Os{(µ-dpp)Ru(bpy)2}3] 8+ (OsRu3) species, in which a central<br />

{Os(µ-dpp)3} 2+ subunit is surrounded by three {Ru(bpy)2} 2+ subunits,<br />

only the osmium-based core emission is obtained (acetonitrile, room temperature:<br />

λmax = 860 nm; τ = 18 ns; Φ = 1 × 10 –3 ) [262], indicating quantitative<br />

energy transfer from the peripheral Ru-based chromophore to the central<br />

Os-based site; for the larger systems the peripheral Ru-based emission is not<br />

quenched [245, 253]. This result highlights that although downhill or even<br />

isoergonic energy transfer between nearby building blocks in the dendrimers


160 S. Campagna et al.<br />

based on the 2,3-dpp bridging lig<strong>and</strong> is fast <strong>and</strong> efficient, direct downhill energy<br />

transfer between partners separated by high-energy subunits is much<br />

slower <strong>and</strong> can be highly inefficient. This problem has been overcome (1) by<br />

using a third type <strong>of</strong> metal center, namely a Pt(II) one, to prepare decanuclear<br />

species (second-generation dendrimers) having different metal centers<br />

in each “generation” layer (schematically, OsRu3Pt6 species) [263] or, more<br />

recently, (2) in a heptanuclear dendron where the barrier made <strong>of</strong> highenergy<br />

subunits is bypassed via the occurrence <strong>of</strong> consecutive electron transfer<br />

steps [264]. Quite interestingly, this latter study suggests that long-range<br />

photoinduced electron transfer processes do not appear to be dramatically<br />

slowed down by interposed high-energy subunits in this class <strong>of</strong> dendrimers.<br />

The efficiency <strong>of</strong> energy migration in 2,3-dpp-based dendrimers has attracted<br />

a large interest for the potential use <strong>of</strong> these species as synthetic<br />

antennae in artificial photosynthesis processes, <strong>and</strong> this has stimulated detailed<br />

kinetic investigations by means <strong>of</strong> ultrafast techniques. Studies on dinuclear<br />

model compounds have shown that esoergonic <strong>and</strong> isoergonic energy<br />

transfer between nearby units occurs within 200 fs, probably from nonthermalized<br />

excited states [168]. A direct consequence <strong>of</strong> such results is that<br />

energy transfer involving singlet states can compete with intersystem crossing.<br />

This conclusion is supported by the fact that the energy transfer from<br />

the peripheral Ru(II) subunits to the central Os(II) core in a tetranuclear<br />

OsRu3 dendrimer takes place both by a singlet–singlet pathway, with a lifetime<br />

<strong>of</strong> less than 60 fs, <strong>and</strong> by triplet–triplet energy transfer, with a lifetime<br />

<strong>of</strong> 600 fs [169, 170]. The finding <strong>of</strong> singlet–singlet energy transfer is a particularly<br />

important result, since it indicates that the idea that any excited-state<br />

process involving metal polypyridine complexes had to be ascribed only to<br />

triplet states should be taken with caution when a significant electronic coupling<br />

between donor <strong>and</strong> acceptor is present. In some way, this finding also<br />

parallels the results obtained for photoinduced injection <strong>of</strong> electrons into<br />

semiconductors [167, 265–271].<br />

An extension <strong>of</strong> this kind <strong>of</strong> antenna is a first-generation heterometallic<br />

dendrimer with appended organic chromophores like pyrenyl units<br />

(36) [272]. In this species, consisting <strong>of</strong> an Os(II)-based core surrounded by<br />

three Ru(II)-based moieties <strong>and</strong> six pyrenyl units in the periphery, 100% efficient<br />

energy transfer to the Os(II) core is observed, regardless <strong>of</strong> the light<br />

absorbing unit. A detailed investigation <strong>of</strong> the excited-state dynamics occurring<br />

in this multicomponent species on exciting in the UV region (267 nm)<br />

has also been performed [273]. Transient absorption spectra (in the range<br />

420–700 nm) for the various intermediates have been reported by the acquisition<br />

<strong>of</strong> evolution-associated difference spectra.<br />

Energy transfer processes from the nonrelaxed <strong>and</strong> relaxed S1 state <strong>of</strong> the<br />

peripheral pyrenyl chromophores to the lowest-lying Os-based MLCT triplet<br />

excited state occur with lifetimes <strong>of</strong> about 6 <strong>and</strong> 45 ps, respectively [273]. Subpicosecond<br />

energy transfer from the excited Ru manifold to the Os-based


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 161<br />

chromophore <strong>and</strong> interconversion between the initially prepared S3 state<br />

<strong>and</strong> the low-lying S1 level within the pyrenyl subunits have also been evidenced.<br />

The rate constant <strong>of</strong> the energy transfer from the pyrenyl groups<br />

to the Ru/Os excited state manifold is in good agreement with the Förster<br />

mechanism when the relaxed S1 pyrene state is taken into account. Energy<br />

transfer from the nonrelaxed state most likely involves folded conformations<br />

in which the pyrenyl subunits are strongly interacting with inner subunits <strong>of</strong><br />

the tetranuclear core. Such interactions were also suggested by the groundstate<br />

absorption spectrum <strong>of</strong> the compound [272].<br />

Because octahedral metal complexes can exist in two chiral forms, Λ <strong>and</strong><br />

∆, it could be expected that the photophysical properties <strong>of</strong> dendrimers<br />

containing metal complexes as branching centers could be different for the<br />

various isomers (the situation can be even more complicated in the case <strong>of</strong><br />

geometrical isomers). However, the investigation <strong>of</strong> optically pure isomers <strong>of</strong><br />

dinuclear <strong>and</strong> dendritic-shaped tetranuclear species (37 is the general structural<br />

formula <strong>of</strong> the tetranuclear systems: optical geometry is not evidenced)<br />

has shown that stereochemical isomerism does not cause any sizeable difference,<br />

at least for the studied compounds [194].<br />

In 37 the emissive state, which involves the peripheral Ru(II) centers, does<br />

not have a sizeable absorption counterpart, since it is a special type <strong>of</strong> chargeseparated<br />

state, with the formal “hole” localized on a peripheral Ru(II) center<br />

<strong>and</strong> the “electron” localized on an orbital mainly centered on the pyrazine<br />

moiety <strong>of</strong> the bridging lig<strong>and</strong>: the absorption related to such a state has negligible<br />

oscillator strength, due to the poor overlap <strong>of</strong> the orbitals involved


162 S. Campagna et al.<br />

(for similar systems, see Sect. 4.4) [194]. A recently investigated, closely related<br />

tetranuclear dendritic-shaped compound, where the bridging lig<strong>and</strong><br />

is the asymmetric PHEHAT lig<strong>and</strong> (PHEHAT = 1,10-phenanthrolino[5,6b]-1,4,5,8,9,12-hexaazatriphenylene),<br />

displays similar photophysical properties<br />

[274].<br />

Mixed-metal Os(II)–Ru(II) dendrimers whose bridging lig<strong>and</strong>s contain<br />

ether linkages have also been reported: compounds 38 <strong>and</strong> 39 are two examples<br />

[275, 276]. The tetranuclear species in 38 exhibits quantitative energy<br />

transfer from the Ru(II) chromophores to the Os(II) core [275], whereas<br />

for 39 the efficiency <strong>of</strong> the energy transfer process is highly temperature<br />

dependent; in fact the Ru-to-Os energy transfer is highly efficient at low temperature,<br />

whereas at room temperature it does not compete with the intrinsic<br />

decay <strong>of</strong> the Ru(II)-based subunits [276].<br />

In most <strong>of</strong> the mixed-metal dendrimers featuring energy transfer <strong>and</strong><br />

containing Ru(II) subunits, the Ru(II) centers play the role <strong>of</strong> the energy<br />

donor components (<strong>and</strong> usually Os(II) centers are the acceptors). Examples<br />

<strong>of</strong> photoactive dendrimers in which Ru(II) subunits behave as acceptor components<br />

are the tetranuclear compound 40 [277] <strong>and</strong> its higher-generation,<br />

Y-shaped octanuclear species, in which four additional Ir(III) chromophores<br />

have been connected at the two peripheral Ir(III) subunits <strong>of</strong> the compound<br />

40 (JAG Williams, personal communication). Here, efficient energy transfer<br />

takes place for the Ir(III) cyclometallated subunits toward the single Ru(II)<br />

polypyridine chromophore, which acts as the energy trap <strong>of</strong> the assemblies.<br />

The Ru(II) subunit then emits with relatively high quantum yield (0.12 in<br />

degassed acetonitrile at room temperature) <strong>and</strong> long lifetime (1.6 µs) [277].


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 163<br />

The fluoro substituents which are present on the peripheral Ir chromophores<br />

have the role <strong>of</strong> increasing the excited-state energy levels <strong>of</strong> the outer Ir chromophores<br />

with regard to the inner ones, so generating the correct energy<br />

gradient.<br />

Multi-ruthenium dendrimers <strong>of</strong> the 2,3-dpp family discussed above have<br />

also been functionalized with organic electron donor subunits, to yield integrated<br />

light-harvesting antennae/electron donor systems for charge separation.<br />

In particular, a triruthenium dendron has been coupled with a tetrathiafulvalene<br />

(TTF) derivative [278] <strong>and</strong> the tetranuclear OsRu3 has been functionalized<br />

at the peripheral bpy lig<strong>and</strong>s with up to six phenothiazine subunits<br />

[279]. In both cases, the light-harvesting antenna emission was totally<br />

quenched by reductive electron transfer from the electron donor subunit.<br />

Interestingly, the electron donor quenchers were not directly linked to the energy<br />

trap <strong>of</strong> the antenna; however, the electron transfer process was fast <strong>and</strong>


164 S. Campagna et al.<br />

efficient, which confirms that in this type <strong>of</strong> artificial antenna metallodendrimer,<br />

long-range electron transfer can be quite effective, as in the case <strong>of</strong><br />

the heptanuclear complex mentioned above [264], <strong>and</strong> could suggest interesting<br />

options to build up integrated donor–antenna–acceptor systems. This<br />

discussion leads us directly to the next section.<br />

5.3<br />

Donor–Chromophore–Acceptor Triads<br />

Triad systems (Fig. 14) are key components <strong>of</strong> the early events in artificial<br />

photosynthesis: the light energy collected by the chromophore (P) is transformed<br />

into chemical (redox) energy by a sequence <strong>of</strong> electron transfer steps<br />

involving electron donor (D) <strong>and</strong> electron acceptor (A) units, ultimately leading<br />

to charge separation [208, 209, 228]. Charge separation is probably the<br />

most important photoinduced process on Earth, so it is not surprising that<br />

many triads based on Ru(II) complexes have been prepared <strong>and</strong> studied in<br />

the last 20 years [228, 280]. It should be noted that there are literally dozens<br />

<strong>of</strong> dyads based on Ru polypyridine complexes [228, 281]. Only some examples<br />

<strong>of</strong> triads (that is, species where Ru(II) chromophores are simultaneously<br />

coupled to electron donor <strong>and</strong> acceptor units) are discussed here.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 165<br />

Fig. 14 Schematization <strong>of</strong> a triad for photoinduced charge separation. P, chromophore; D,<br />

donor; A, acceptor; cs, primary charge separation; cr, primary charge recombination; cs ′ ,<br />

secondary charge separation; cr ′ , final charge recombination<br />

Structurally speaking, Ru complexes with tridentate lig<strong>and</strong>s like terpy are<br />

ideal systems for molecular triads: indeed, substitution at the 4 position <strong>of</strong><br />

the central pyridine ring <strong>of</strong> terpy-like lig<strong>and</strong>s allows a linear arrangement<br />

<strong>of</strong> subunits, with control <strong>of</strong> geometry <strong>and</strong> distance. Some <strong>of</strong> the first ruthenium<br />

triads based on such an arrangement are shown in Fig. 15 [282].<br />

Whereas fully developed charge separation, with formation <strong>of</strong> the D + -P-A –<br />

Fig. 15 Structural formulae <strong>of</strong> Ru(II) terpyridine triads containing acceptor (A) <strong>and</strong><br />

donor (D) components


166 S. Campagna et al.<br />

charge-separated state, did not take place in the triad with D = PTZ (PTZ =<br />

phenothiazine subunit), the formation <strong>of</strong> such a charge-separated species was<br />

inferred for the species with D = DPPA (DPPA = diphenylamino moiety) at<br />

150 K, although it could not be evidenced spectroscopically because it did not<br />

accumulate as a consequence <strong>of</strong> a fast recombination rate [282].<br />

In spite <strong>of</strong> less control <strong>of</strong> distance <strong>and</strong> orientation between donor <strong>and</strong> acceptor,<br />

better results have been reported for a series <strong>of</strong> systems exemplified<br />

by 42 [283–285]. The components <strong>of</strong> this series <strong>of</strong> triads are a tris-bipyridine<br />

Ru(II) chromophore covalently linked to one or two phenothiazine electron<br />

donors <strong>and</strong> to quaternized bipyridinium electron acceptors. The saturated<br />

alkyl chains bridging the molecular components are electrically insulating<br />

<strong>and</strong> flexible. This latter point is apparently a drawback since it does not allow<br />

for control <strong>of</strong> geometry. Moreover, even the octahedral arrangement <strong>of</strong><br />

the bpy subunits adds some difficulties in defining the real structure: for<br />

example, geometrical isomers can also exist, since each bpy <strong>of</strong> 42 is nonsymmetric.<br />

The compound 42 exhibited formation <strong>of</strong> the D + –P–A – chargeseparated<br />

state in dichloromethane at room temperature with an initially<br />

reported efficiency <strong>of</strong> about 26%. Such a value was later corrected to about<br />

86% by using a slightly different solvent (1,2-dichloroethane) [283, 286, 287].<br />

Once formed, the charge-separated state decayed with a relatively fast rate<br />

(kCR = 6.3 × 10 6 s –1 ), corresponding to a lifetime <strong>of</strong> about 160 ns. On the basis<br />

<strong>of</strong> redox data, the charge-separated state stored about 1.3 eV.<br />

Similar complexes were prepared that differed from one another by the<br />

length <strong>of</strong> the alkyl chains connecting the subunits <strong>and</strong>/or by changing


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 167<br />

the methylene chains connecting the quaternary nitrogens (<strong>and</strong>, as a consequence,<br />

the reduction potential <strong>of</strong> the electron acceptor) [283–287]. In<br />

a homogeneous series <strong>of</strong> experiments performed on such species, however,<br />

the efficiency <strong>of</strong> charge separation does not change appreciably, remaining<br />

larger than 0.80, although the driving forces <strong>and</strong> the rate constants <strong>of</strong> the various<br />

electron transfer steps, as obtained by independent studies performed on<br />

isolated dyads <strong>of</strong> the type D–P or P–A, were different.<br />

In D–P ∗ –A systems, the fully developed charge-separated state can be obtained,<br />

in principle, by two different routes (excluding direct electron transfer<br />

from D to A): (1) a route initiated by oxidative quenching, that is, the series<br />

<strong>of</strong> events described by the sequence D–P ∗ –A, D–P + –A – ,D + –P–A – ;<strong>and</strong><br />

(2) a route initiated by reductive quenching, described by the sequence D–<br />

P ∗ –A, D + –P – –A, D + –P–A – . Both routes can also take place simultaneously.<br />

The comparison between the photophysical properties <strong>of</strong> the various triads<br />

<strong>and</strong> the corresponding isolated dyads <strong>of</strong> this family <strong>of</strong> compounds indicated<br />

that the emission decay rates <strong>of</strong> any D–P–A triad never differed by more<br />

than a factor <strong>of</strong> two from those <strong>of</strong> the P–A dyads, although the absolute decay<br />

rate values changed by over a factor <strong>of</strong> 10 3 (over the whole collection<br />

<strong>of</strong> compounds). This prompted the authors to attribute the initial quenching<br />

event in all the D–P–A triads <strong>of</strong> this family to oxidative electron transfer, with<br />

formation <strong>of</strong> the D–P + –A – intermediate, with the route initiated by reductive<br />

quenching playing a negligible role [288]. However, in all the P–A dyad<br />

systems, it was always impossible to detect the A – radical anion [284], indicating<br />

that back electron transfer in the P–A dyads was faster than the<br />

forward, oxidative electron transfer. This posed some problems in justifying<br />

the efficiency <strong>of</strong> formation <strong>of</strong> the fully developed charge-separated state,<br />

where apparently reduction <strong>of</strong> P + from D in D–P + –A – species efficiently<br />

competes with back electron transfer in the intermediate. In fact, this looks<br />

somewhat puzzling because the reductive electron transfer in D–P ∗ dyads is<br />

reported to be <strong>of</strong> the order <strong>of</strong> 10 6 s –1 [289], while oxidative electron transfer<br />

in P ∗ –A dyads ranges from 10 10 to 10 7 s –1 [284, 285, 290–293] <strong>and</strong>, based<br />

on the circumstances mentioned above, back electron transfer in P + –A – (<strong>and</strong><br />

for extension in D–P + –A – ), opposing the formation <strong>of</strong> the fully developed<br />

charge-separated state, could be even faster. To justify the experimental data,<br />

electron transfer from D to P + in D–P + –A – should be about 1 × 10 10 s –1 or<br />

faster. Therefore, the exceptional properties <strong>of</strong> these compounds as far as the<br />

efficiency <strong>of</strong> charge separation is concerned remained largely unexplained.<br />

A recent paper has shed light on the photophysical behavior <strong>of</strong> these<br />

triads [288]. A series <strong>of</strong> new experiments, including transient absorption<br />

measurements, emission decay, <strong>and</strong> a careful examination <strong>of</strong> the ground-state<br />

absorption spectra <strong>of</strong> the triads <strong>and</strong> <strong>of</strong> various separated dyad components,<br />

suggested that in the D–P–A triads <strong>of</strong> this family an association between the<br />

tethered phenothiazine electron donor subunit <strong>and</strong> the Ru(II) chromophore<br />

takes place, in a folded conformation. The association is already present in the


168 S. Campagna et al.<br />

ground state, before excitation <strong>and</strong> any electron transfer process. The triad<br />

would then be better defined as a D/P–A system, <strong>and</strong> electron transfer between<br />

D <strong>and</strong> P + subunits in the intermediate state D/P + –A – can be largely<br />

faster than that reported for the D–P ∗ dyad, <strong>and</strong> even faster than charge<br />

recombination between C + <strong>and</strong> A – in the assembly, justifying the high efficiency<br />

<strong>of</strong> formation <strong>of</strong> the fully developed charge separation. Ground-state<br />

association <strong>of</strong> tethered aromatic lig<strong>and</strong>s (with flexible linkages) with Ru(II)<br />

chromophores was also known in a tetranuclear dendrimer [272, 273], further<br />

supporting these conclusions.<br />

The triad 43 involves components similar to those used in the former<br />

discussed systems, but the connecting scheme is different [294]. Here, the<br />

chromophore <strong>and</strong> the electron donor <strong>and</strong> acceptor subunits are all linked to<br />

a lysine moiety. The charge-separated state <strong>of</strong> this triad stores 1.17 eV <strong>and</strong><br />

lives for 108 ns (rate constant <strong>of</strong> charge recombination kCR = 9.26 × 10 6 s –1 )<br />

in acetonitrile, as observed by transient spectroscopy. The efficiency <strong>of</strong> formation<br />

<strong>of</strong> the charge-separated state is about 34%. On the basis <strong>of</strong> detailed<br />

photophysical studies <strong>of</strong> isolated dyads [294], the authors believe that the only<br />

route leading to fully developed charge separation is reductive quenching <strong>of</strong><br />

D–P ∗ –A leading to D + –P – –A, followed by a second electron transfer leading<br />

to D + –P–A – , whereas the route passing through the population <strong>of</strong> the D–P + –<br />

A – intermediate does not lead to the full charge-separated species, but leads<br />

directly to the ground state. Therefore, the occurrence <strong>of</strong> oxidative electron<br />

transfer <strong>of</strong> P* would be the main reason for the efficiency loss in the whole<br />

process. Better results were obtained by modifying the acceptor site <strong>of</strong> the<br />

triad, as shown in 44 [295]. In this species, the energy stored by the chargeseparated<br />

species is higher (1.54 eV), since the anthraquinone has a less negative<br />

reduction potential than the methyl viologen subunit, <strong>and</strong> the lifetime<br />

<strong>of</strong> the charge-separated state is longer (174 ns, kCR = 5.7 × 10 6 s –1 ), but the


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 169<br />

quantum yield is slightly smaller (26%). The longer lifetime <strong>of</strong> the chargeseparated<br />

species <strong>of</strong> the anthraquinone-containing triad was attributed to the<br />

charge-recombination process being further in the inverted region (a similar<br />

reorganization energy λ,about0.80 eV, is assumed in both cases). Even in<br />

this case, the main losses in the efficiency <strong>of</strong> formation <strong>of</strong> the fully developed<br />

charge separation were attributed to the inefficiency <strong>of</strong> the route in which the<br />

initial electron transfer is oxidative to produce the final D + –C–A – state.<br />

Within the field <strong>of</strong> molecular triads, a peculiar example is 45 [296]. In this<br />

species, the Ru(II) subunit playing the role <strong>of</strong> the chromophore is mechanically<br />

linked with a cyclobis(paraquat-p-phenylene) unit (BV 4+ ,acceptor)<strong>and</strong><br />

covalently linked with a protoheme unit (the donor), located in a myoglobin<br />

pocket (not shown in figure). The overall system is therefore a reconstituted<br />

protein bearing a molecular triad. Excitation <strong>of</strong> the ruthenium chromophore<br />

is followed by a series <strong>of</strong> photoinduced electron transfer steps (the first<br />

one being electron transfer to the electron acceptor), leading to the chargeseparated<br />

species containing the porphyrin radical cation <strong>and</strong> the paraquatbased<br />

radical anion, Mb(Fe III OH2) + -Ru 2+ -BV 3+ . This species successively undergoes<br />

a series <strong>of</strong> deprotonation processes ultimately leading to another<br />

charge-separated species, identified as Mb(Fe IV =O)-Ru 2+ -BV 3+ ,withanap-


170 S. Campagna et al.<br />

parent first-order rate constant <strong>of</strong> 6.6 × 10 3 s –1 . This species is produced with<br />

alowquantumyield(0.5%), stores about 1.3 eV, <strong>and</strong> recombines to the<br />

ground state with a lifetime exceeding 2 ms. In spite <strong>of</strong> the low efficiency<br />

<strong>of</strong> the charge-separation process, there are several interesting points: (1) in<br />

the absence <strong>of</strong> myoglobin, charge separation is not obtained at all; (2) the<br />

complete process is pH dependent; (3) the final charge-separated state lifetime<br />

is comparable with that <strong>of</strong> natural photosynthetic reaction centers; <strong>and</strong><br />

(4) back electron transfer is regulated by protonation/deprotonation <strong>of</strong> distal<br />

histidine moieties, which appears to be needed to reduce the Mb(Fe IV =O)<br />

subunit. The low efficiency <strong>of</strong> the overall process is mainly attributed to<br />

charge recombination within the Mb(Fe III OH2) + -Ru 2+ -BV 3+ state, which efficiently<br />

competes with the deprotonation processes. This study highlights the<br />

potential <strong>of</strong> mixed synthetic–natural systems for obtaining long-lived charge<br />

separation.<br />

5.4<br />

Polyads Based on Oligoproline Assemblies<br />

The interesting results obtained by organizing D–P–A triads on the structure<br />

<strong>of</strong> the amino acid lysine (Sect. 5.3) prompted the preparation <strong>of</strong> D–P–A systems<br />

assembled on oligoproline scaffolds, by means <strong>of</strong> solid-state peptide<br />

synthesis [294, 297–299]. One such system is 46. In this species, a phenothiazine<br />

(PTZ) group acts as the electron donor <strong>and</strong> an anthraquinone (Anq)<br />

subunit plays the role <strong>of</strong> the electron acceptor. Oligoprolines were selected,<br />

since it is known that oligoproline chains <strong>of</strong> nine or more proline units<br />

fold into stable helices even with large functional groups on the proline<br />

units. The terminal segments allow the helix to begin <strong>and</strong> end with capped<br />

Pro3 turns which prevent unwinding <strong>of</strong> the helix. For 46, a fully developed<br />

charge-separated state is gained in acetonitrile solution with good efficiency<br />

(53%). The charge-separated state stores 1.65 eV relative to the ground<br />

state <strong>and</strong> returns to the ground state with a rate constant <strong>of</strong> 5.7 × 10 6 s –1<br />

(τ = 175 ns) [299, 300]. Quenching <strong>of</strong> the Ru-based excited state is domi-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 171<br />

nated by reductive electron transfer involving the PTZ electron donor, in<br />

a process which is largely solvent dependent, as is the driving force <strong>of</strong> the<br />

process. Then, there is a fast electron transfer from the reduced metal chromophore<br />

to the Anq subunit, which yields the charge-separated state. There<br />

is a strong solvent-dependent competition between such a second electron<br />

transfer, which allows for fully developed charge separation <strong>and</strong> back electron<br />

transfer in the D + –P – –A intermediate. The consequence <strong>of</strong> this solvent<br />

dependence is that going from 1,2-dichloroethane to dimethylacetamide, the<br />

efficiency <strong>of</strong> charge separation changes from 33 to 96%. Also, the charge recombination<br />

is solvent dependent, <strong>and</strong> the electronic coupling between PTZ +<br />

<strong>and</strong> Anq – was calulated to be about 0.13 cm –1 [300].<br />

A more elaborated polyad based on the formerly described systems is<br />

the D–P–P–A tetrad 47 [301]. This is the evolution <strong>of</strong> a system quite related<br />

to 46, where substituents on the terminal bpy lig<strong>and</strong>s <strong>of</strong> the metal<br />

chromophore are used to favor the thermodynamics <strong>of</strong> the (reductive) first<br />

electron transfer step. This modification led to an efficiency <strong>of</strong> charge separation<br />

<strong>of</strong> 90% in acetonitrile for the corresponding triad. In the tetrad, 13<br />

proline spacers are present between PTZ <strong>and</strong> Anq. The efficiency <strong>of</strong> formation<br />

<strong>of</strong> the charge-separated state in the tetrad is 60% <strong>and</strong> its lifetime is 2 µs<br />

(kCR = 5.0 × 10 5 s –1 ). Excitation can occur in both the Ru chromophores, but<br />

apparently the result is not identical. Excitation <strong>of</strong> the Ru(II) complex adjacent<br />

to the PTZ electron donor gives the D + –P – –P–A system. To produce the<br />

fully developed D + –P–P–A + species, it is proposed that a stepwise mechanism<br />

occurs, with the species D + –P–P – –A as an intermediate. Efficient, isoergonic<br />

electron transfer between the two chromophores is therefore foreseen. Excitation<br />

<strong>of</strong> the Ru(II) complex adjacent to the electron acceptor Anq subunit<br />

would be unproductive in a direct sense, since oxidative electron transfer by<br />

Anq is unfavorable thermodynamically <strong>and</strong> direct quenching from the PTZ<br />

unit is unlikely because <strong>of</strong> the large distance. However, even excitation <strong>of</strong> this<br />

Ru(II) chromophore can become productive, provided that isoergonic energy<br />

transfer to the Ru(II) chromophore adjacent to the PTZ unit takes place. Since<br />

the quantum efficiency <strong>of</strong> formation <strong>of</strong> the charge-separated state is 60%,<br />

<strong>and</strong> considering that excitation <strong>of</strong> the two identical Ru(II) chromophores


172 S. Campagna et al.<br />

is equivalent (that is, 50% each), clearly interchromophoric energy transfer<br />

takes place, although not with a high efficiency.<br />

As far as the mechanism <strong>of</strong> the intrastr<strong>and</strong> energy <strong>and</strong> electron transfer<br />

in such oligoprolines is concerned, through-bond <strong>and</strong> through-space pathways<br />

can be considered. An important step to elucidate the mechanism was<br />

a systematic investigation <strong>of</strong> the reductive electron transfer in oligoprolines<br />

containing only PTZ <strong>and</strong> one Ru(II) chromophore [302]. From such a study,<br />

when the number <strong>of</strong> interposed prolines varies from two to five, it was<br />

demonstrated that a through-space mechanism is operative, with an apparent<br />

β value <strong>of</strong> 0.41 ˚A –1 . Superexchange coupling with the solvent <strong>and</strong> oligoproline<br />

scaffold are proposed to play important roles in promoting electronic<br />

coupling [302].<br />

To complete the overview <strong>of</strong> these oligoproline-based systems, it has to<br />

be mentioned that also pentads <strong>of</strong> type D–P–P–P–A have been prepared <strong>and</strong><br />

studied, where each functional subunit is separated from the adjacent ones by<br />

two prolines [297, 303]. The results obtained indicate that interchromophoric<br />

energy transfer is relatively slow across five proline units, but rapid across two<br />

prolines [297].<br />

5.5<br />

Multi-ruthenium Assemblies Based on Derivatized Polystyrene<br />

A suitable choice to assemble a large number <strong>of</strong> chromophores is polymer<br />

derivatization: within this class <strong>of</strong> compounds, probably the largest<br />

series deals with polystyrene polymers. Indeed, up to 30 Ru–bpy-type chromophores<br />

have been linked to soluble styrene polymers [297, 304–308, 310,<br />

318]. An early synthetic strategy [309] involved attachment <strong>of</strong> the chromophores<br />

to polystyrene by an ether linkage (48). Successively [310], the


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 173<br />

same authors developed an alternative route based on amide linkage (49),<br />

finding a dramatic enhancement in the ability <strong>of</strong> the polymeric arrays to<br />

promote intrastr<strong>and</strong> energy transfer. In the (amide-functionalized) mixed<br />

polymers containing a 3 : 13 ratio between the lower-energy Os-based chromophores<br />

<strong>and</strong> the higher-energy Ru-based ones, triplet–triplet energy transfer<br />

was found to occur with an efficiency higher than 0.90 in acetonitrile<br />

solution.<br />

It was pointed out [310] that energy transfer from the excited Ru-based<br />

moieties to the ground-state Os-based moieties requires two processes: energy<br />

migration among the Ru-based units (site-to-site energy hopping) <strong>and</strong><br />

a final energy transfer from a Ru-based to a nearby Os-based unit. For both<br />

processes the rate constants exceeded 2 × 10 8 s –1 for the amide-linked polymer,<br />

whereas in the ether-linked polymer the rate constant for intrastr<strong>and</strong><br />

energy migration from ∗ Ru to Ru was orders <strong>of</strong> magnitude slower. The<br />

rate constants for the amide-linked species, coupled with the relatively long<br />

excited-state lifetime <strong>of</strong> the Ru-based chromophores (910 ns), account for the<br />

ability <strong>of</strong> the polymer arrays containing the amide-linked chromophores to<br />

act as efficient antennae. The reason for the different behavior <strong>of</strong> the etherlinked<br />

<strong>and</strong> amide-linked arrays lies in the direction <strong>of</strong> the excited-state MLCT<br />

dipole <strong>of</strong> the chromophores involved in the energy transfer processes. This<br />

dipole is directed toward the polymer backbone in the most effective amidelinked<br />

antenna systems, whereas it is out from the polymer backbone in the<br />

less efficient ether-linked arrays. This difference affects the electronic coupling<br />

between donor <strong>and</strong> acceptor sites <strong>of</strong> the energy migration processes.<br />

Ru(II) chromophores with appended electron acceptor (a methyl viologen<br />

species, A) <strong>and</strong> donor (a phenothiazine group, D) subunits have been incorporated<br />

within the antenna polystyrene system (50) [311], to play the role<br />

<strong>of</strong> “reaction center” (RC) units. In such an integrated antenna–RC polymer,


174 S. Campagna et al.<br />

containing 17 normal Ru chromophores <strong>and</strong> three RCs, the D + –A + chargeseparated<br />

state is formed, as shown by transient absorption spectroscopy.<br />

Emission at the Ru(II) chromophores was quenched by about 34% compared<br />

to the homopolymer containing 20 “normal” Ru chromophores. Energy<br />

transfer from the normal Ru chromophores to the RC sites was favored by<br />

– 0.1 eV. It was shown that about 50% <strong>of</strong> the charge-separated state was<br />

formed during the 7-ns laser pulse, indicating intrastr<strong>and</strong> sensitization, with<br />

the charge-separated state formed by direct excitation at the RC complex <strong>and</strong><br />

by excitation to nonadjacent Ru chromophores followed by energy migration<br />

to the RC sites. In this system, 1.15 eV is stored in the charge-separated state<br />

<strong>and</strong> the efficiency <strong>of</strong> the process varies from 12 to 18% dependingonlaser<br />

irradiance, indicating excited-state annihilation at high irradiance. Charge<br />

recombination is similar to that <strong>of</strong> the “isolated” RC, but an additional longlivedtransient(formedinlowefficiency,about0.5%)<br />

was observed, which<br />

decayed by second-order kinetics with k = 48 M –1 s –1 .Thislong-livedtransient<br />

was attributed to polymers in which D + <strong>and</strong> A – were formed on different<br />

RC units, by invoking mechanisms in which electron transfer quenching by<br />

oxidative or reductive electron transfer in a RC site is followed by intrastr<strong>and</strong><br />

hole or electron transfer to a second RC site [311].<br />

5.6<br />

Photoinduced Collection <strong>of</strong> Electrons<br />

into a Single Site <strong>of</strong> a Metal Complex<br />

An essential property <strong>of</strong> natural photosynthesis is the collection <strong>of</strong> multiredox<br />

equivalents at specific sites. Indeed, all the important light energy storage


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 175<br />

processes require more than one electron to operate: for example, reduction<br />

<strong>of</strong> H + to H2 is bielectronic, <strong>and</strong> oxidation <strong>of</strong> oxygen in water to produce O2 is<br />

a four-electron process. Reduction <strong>of</strong> CO2 to the high-energy content glucose<br />

species is also a multielectron process. Whereas artificial systems capable <strong>of</strong><br />

performing photoinduced charge separation have been reported, species able<br />

to collect, by successive photoinduced processes, more than one single electron<br />

(or hole) in one specific site <strong>of</strong> their structure are very rare. These species<br />

differ from polymers or dendritic species, which are also able to reversibly<br />

store more than one single electron (or hole) in their structure (in several,<br />

roughly identical, but spatially separated sites), since the accumulated charges<br />

should be located in a single subunit <strong>and</strong>, at least in principle, could be more<br />

easily delivered simultaneously to a unique substrate.<br />

A breakthrough in this field was the study <strong>of</strong> the two dinuclear Ru complexes<br />

51 <strong>and</strong> 52 [312]. These complexes are indeed able to collect two electrons<br />

(<strong>and</strong> two protons) <strong>and</strong> four electrons (<strong>and</strong> four protons), respectively,<br />

within their bridging lig<strong>and</strong> moieties upon successive light excitation <strong>and</strong><br />

in the presence <strong>of</strong> sacrificial donor species. In a typical (schematized) sequence<br />

<strong>of</strong> events involving 51: (1) light excitation produces a MLCT state<br />

involving the bpy-like subunit <strong>of</strong> the bridge; (2) a charge shift takes place<br />

from the bpy-like bridge moiety to the inner, phenazine-like portion <strong>of</strong> the<br />

bridge, so producing a sort <strong>of</strong> charge-separated state; (3) the sacrificial donor,<br />

a triethylamino (TEA) species, reduces the Ru(III) center, so restoring the<br />

chromophore; (4) the reduced central moiety <strong>of</strong> the bridge adds a proton<br />

(originated from irreversible TEA oxidation), so reaching charge neutrality;<br />

<strong>and</strong> (5) the sequence <strong>of</strong> events 1–4 is repeated <strong>and</strong> two electrons <strong>and</strong><br />

two protons are collected [312]. However, a recent refinement <strong>of</strong> the ultrafast<br />

spectroscopic results has evidenced that the product <strong>of</strong> step 2, initially<br />

identified as a sort <strong>of</strong> charge-separated state [313], receives a significant contribution<br />

also from a bridge-centered triplet state [314]. The overall process<br />

is perfectly reversible, <strong>and</strong> 51 is fully restored on leaving molecular oxygen<br />

reaching the complex [312]. For 52, the formerly described sequence <strong>of</strong> events<br />

is repeated four times, thanks to the presence <strong>of</strong> the quinone subunits responsible<br />

for the addition <strong>of</strong> two extra electron/proton couples [312]. All the<br />

various steps <strong>of</strong> the multielectron processes occurring in 51 have also been<br />

characterized by UV/Vis spectroscopy <strong>and</strong> each intermediate has a unique


176 S. Campagna et al.<br />

signature. This characterization was made by extensive work (including spectroelectrochemistry<br />

in various solvents, as well as at different pH values in<br />

aqueous solution) to identify the intermediates <strong>and</strong> to clarify the effect <strong>of</strong><br />

pH on the processes [315–317]. Indeed, it has been demonstrated that in<br />

some solvents, processes 2 <strong>and</strong> 4 take place simultaneously, so it would be<br />

more correct to talk <strong>of</strong> proton-coupled electron transfer instead <strong>of</strong> successive<br />

electron/proton transfer events. At pH 4, only a fully reversible, coupled twoelectron/two-proton<br />

transfer process is observed for 51. This is a rare example<br />

<strong>of</strong> a proton-coupled multielectron transfer reaction [317] (FM MacDonnell,<br />

personal communication).<br />

Another interesting example <strong>of</strong> photoinduced multielectron collection,<br />

this time at a metal center rather than at a lig<strong>and</strong> site, has been reported for<br />

a series <strong>of</strong> trimetallic mixed-metal species [318–320]. The most recent example<br />

<strong>of</strong> the series is 53 [320]. In this species, two Ru(II) chromophores (the<br />

light absorber units, LA) are linked to one Rh(III) center, which represents the<br />

electron collection (EC) core, through polyazine bridging lig<strong>and</strong>s (BLs). The<br />

absorption spectrum <strong>of</strong> this species is dominated by the absorption <strong>of</strong> the Ru<br />

LA subunits, while the reduction properties identify the Rh(dσ ∗ ) orbital as<br />

the LUMO [319]. Upon excitation <strong>of</strong> the peripheral Ru(II) chromophore, an<br />

oxidative electron transfer to the Rh(III) center takes place (k = 1.2 × 10 8 s –1 ),<br />

populating a triplet Ru(II)-to-Rh(III) charge transfer state. In the presence <strong>of</strong><br />

a sacrificial donor, dimethylaniline, which restores the Ru(II) center(s), the<br />

starting compound undergoes a net photoreduction process with formation<br />

<strong>of</strong> [{(bpy)2Ru(dpp)}2Rh I ] 5+ , as also demonstrated by spectroelectrochemistry,<br />

in which two chlorides have been lost (probably by irreversible chlo-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 177<br />

ride oxidation), <strong>and</strong> a two-electron reduced unit Rh(I) is formed [320]. The<br />

Rh center should simultaneously undergo a structural reorganization (most<br />

likely, octahedral/square planar). Interestingly, the photoreduced species is<br />

coordinatively unsaturated <strong>and</strong> therefore could be available to interact with<br />

substrates. It warrants mentioning that photoinitiated electron collection is<br />

obtained in a trimetallic species having an Ir(III) redox-active center instead<br />

<strong>of</strong> a Rh(III) one <strong>and</strong> similar Ru-based LA units [318].<br />

5.7<br />

Photoinduced Multihole Storage: Mixed Ru–Mn Complexes<br />

Complementary to the topic discussed in the previous section (that is, to accumulate<br />

multiple electrons in a single site <strong>of</strong> a (supra)molecular species)<br />

is the development <strong>of</strong> systems capable <strong>of</strong> accumulating holes, as happens in<br />

the oxygen evolving systems <strong>of</strong> natural photosynthesis. The source <strong>of</strong> inspiration<br />

is the photosystem II [321], where the excited primary chlorophyll donor,<br />

∗ P680, one <strong>of</strong> the most effective photooxidants <strong>of</strong> natural systems, is able to extract<br />

up to four electrons in consecutive steps from the so-called manganese<br />

cluster, whose structure—at least for a specific natural system—has recently<br />

been revealed [322–324]. The four-times oxidized manganese cluster successively<br />

produces molecular oxygen, thus returning to its initial state, ready for<br />

another photoinduced catalytic cycle.<br />

Since the inspiration is photosystem II, it is not surprising that the largest<br />

family <strong>of</strong> complexes made to photochemically accumulate “holes” are Ru(II)<br />

polypyridine complexes coupled to manganese species. The field has been<br />

recently reviewed [325].<br />

Several Ru–Mn dyads were initially studied to investigate some specific<br />

parameters for electron transfer (see for example 54–57 [326]). In a typical<br />

experiment, the excited Ru(II) chromophore is quenched via a bimolecular<br />

oxidative electron transfer by a sacrificial acceptor (usually methyl viologen),<br />

<strong>and</strong> the oxidized Ru(III) species oxidizes the attached Mn(II) subunit to<br />

Mn(III). Time constants <strong>of</strong> these latter processes spanned a large range, from


178 S. Campagna et al.<br />

< 50 ns to 10 µs, depending on the complex [326]. This study demonstrated<br />

that the reorganization energy for the Mn(II)-to-Ru(III) electron transfer was<br />

quite large (1.4–2.0 eV), suggesting significant inner reorganization <strong>of</strong> the<br />

manganese moiety during the process [327]. As an obvious consequence,<br />

very fast Mn(II)-to-Ru(III) electron transfer could not be expected. A further<br />

complication was that the manganese subunit could directly quench the<br />

excited ruthenium chromophore by Dexter energy transfer, competing with<br />

electron transfer quenching by the sacrificial acceptor for short Ru–Mn distances.<br />

These arguments led to the preparation <strong>of</strong> more elaborated systems<br />

in which intermediate donor species were interposed between ruthenium <strong>and</strong>


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 179<br />

manganese subunits [328–330], with phenolate <strong>and</strong> tyrosine moieties playing<br />

the role <strong>of</strong> an intermediate donor. In most cases, proton-coupled electron<br />

transfer processes took place.<br />

To achieve multielectron catalysts, more than one manganese ion was included<br />

in the systems. Figure 16 shows some examples [329, 331, 332] <strong>of</strong><br />

Ru(II) species covalently linked to Mn dimers or trimers via phenolate lig-<br />

<strong>and</strong>s. In particular, for the Ru–Mn II,II<br />

2<br />

complex 58 reported in the figure,<br />

repeated flashes in the presence <strong>of</strong> a Co(III) sacrificial electron acceptor allowed<br />

three successive one-electron oxidations <strong>of</strong> the manganese moiety by<br />

the photooxidized Ru(III) subunit [333], as evidenced by the disappearance<br />

<strong>of</strong> the characteristic Mn2 II,II signals <strong>and</strong> the appearance <strong>of</strong> the characteristic<br />

Mn2 III,IV signals in EPR experiments. Manganese oxidation was suggested to<br />

involve a lig<strong>and</strong> exchange, in which acetate is released <strong>and</strong> water molecules<br />

are bound to form a di-µ-oxo bridge (see the reaction scheme in Fig. 16). According<br />

to the authors, this was the first example <strong>of</strong> a light-driven, multiple<br />

oxidation <strong>of</strong> a manganese complex attached to a photosensitizer. The lig<strong>and</strong><br />

exchange at the manganese sites (presumably occurring in the Mn2 III,III<br />

state, that is, after second electron release from the initial Mn2 II,II center)<br />

is functional to the overall process, as it allows introduction <strong>of</strong> negative<br />

Fig. 16 Electron transfer from the manganese moiety to the photooxidized Ru(III) in a<br />

aRu– Mn2 II,II complex <strong>and</strong> b aRu– Mn3 II,II,II complex


180 S. Campagna et al.<br />

charges to the complex thus making feasible oxidation to the Mn2 III,IV state,<br />

which would have otherwise been impossible on thermodynamic grounds.<br />

The lig<strong>and</strong> exchange <strong>and</strong> reorganization <strong>of</strong> the manganese ion coordination<br />

sphere, which has a nonneutral effect from the viewpoint <strong>of</strong> the charge, would<br />

be a charge compensating process, analogous to proton release occurring<br />

in most <strong>of</strong> the oxidation states <strong>of</strong> the “manganese cluster” in natural systems<br />

[321, 325]. It can be inferred that charge compensating processes are<br />

needed requirements to maintain the oxidation potential <strong>of</strong> the redox-active<br />

catalytic site roughly constant when moving along the various steps <strong>of</strong> the<br />

overall hole accumulating process, an aspect that should be well taken into<br />

account in designing new systems.<br />

Recently a mixed Ru–Mn2 species featuring a photoinduced chargeseparation<br />

state with an impressive lifetime (0.6 ms at room temperature<br />

<strong>and</strong> 0.1–1 sat140 K, comparable to many <strong>of</strong> the naturally occurring chargeseparated<br />

states in photosynthetic systems) has been reported [334]. The<br />

slow charge-recombination rate obtained for such a species has been mainly<br />

attributed to the large reorganization energy connected with the inner reorganization<br />

<strong>of</strong> the manganese subunit already mentioned (about 2 eV for the<br />

compound in [334]). This would suggest that there is no needed to look for<br />

charge-recombination processes occurring in the Marcus inverted region to<br />

obtain long-lived separated states, since the large inner reorganization energy<br />

typical <strong>of</strong> the manganese systems could lead to the same (or better) result.<br />

5.8<br />

Photocatalytic Processes Operated by Supramolecular Species<br />

5.8.1<br />

Photogeneration <strong>of</strong> Hydrogen<br />

Since the early papers appeared in the 1970s [335–339], Ru(II) polypyridine<br />

complexes have been extensively used to produce hydrogen in heterogeneous<br />

cycles under light irradiation, by using sacrificial donor species (most commonly<br />

amines), electron acceptor relays (usually methyl viologens), <strong>and</strong> colloidal<br />

metal catalysts (Pt, Rh, etc.). This aspect <strong>of</strong> Ru(II) photochemistry has<br />

been extensively reviewed [1, 340] <strong>and</strong> will not be discussed in detail here. We<br />

will discuss some recent papers in which a (supramolecular) multicomponent<br />

approach is used.<br />

One <strong>of</strong> the important steps in designing a multicomponent hydrogen<br />

evolving system would be to assemble in the same (supramolecular) system<br />

as many key components as possible. Key components would be (based on the<br />

systems operating in heterogeneous schemes): (a) light-harvesting units (antennae);<br />

(b) a charge-separation unit made <strong>of</strong> a photosensitizer (the energy<br />

trap <strong>of</strong> the antenna, if an antenna is present), an electron acceptor, <strong>and</strong> an<br />

electron donor; <strong>and</strong> (c) a catalyst [280, 297, 341–345]. The compound 60 [346]


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 181<br />

is a quite interesting example toward the preparation <strong>of</strong> functional supramolecular<br />

species <strong>of</strong> this type, where all the key components would be integrated<br />

into a single multicomponent system. Under visible irradiation, 60 evolves<br />

molecular hydrogen from water (in the presence <strong>of</strong> EDTA as sacrificial donor)<br />

with an efficiency <strong>of</strong> about 1%. A relatively low turnover number (4.8) is estimated<br />

on the basis <strong>of</strong> the total amount <strong>of</strong> H2 evolved after 10 h(2.4 µmol) <strong>and</strong><br />

the amount <strong>of</strong> the complex used (1 µmol). The hydrogen formation is immediate<br />

<strong>and</strong> is rather higher in its rate when UV light is eliminated by suitable<br />

filters. This latter observation suggests some photoinstability <strong>of</strong> the system<br />

upon UV irradiation.<br />

The wavelength dependence in the visible region <strong>of</strong> the rate <strong>of</strong> H2 formation<br />

agrees with the absorption spectrum <strong>of</strong> the complex. Moreover, the<br />

rate <strong>of</strong> H2 formation increases linearly with the photon flux, indicating that<br />

one-photon excitation <strong>of</strong> the molecule operates. This multicomponent system<br />

integrates in its structure the light harvester/photosensitizer (the Ru(II) chromophore)<br />

<strong>and</strong> the electron acceptor, which also plays the role <strong>of</strong> the catalyst<br />

(the Pt(II) subunit). Compared to the formerly investigated systems [1, 335–<br />

337, 339, 456], neither the electron relay (usually a methyl viologen species) or<br />

the solid-state catalyst (usually colloidal platinum or rhodium) are required.<br />

Mechanistic aspects have not been discussed yet.<br />

Another multicomponent system (61) has been recently reported [347].<br />

When excited at 470 nm in acetonitrile <strong>and</strong> in the presence <strong>of</strong> triethylamine<br />

(TEA, 2.08 mol L –1 ), such a species evolves H2 from the solution with a turnover<br />

number <strong>of</strong> about 50 (mol <strong>of</strong> H2 per mol <strong>of</strong> compound). No H2 evolution is ob-


182 S. Campagna et al.<br />

tained in the dark or when one <strong>of</strong> the key components (the Ru(II) chromophore,<br />

the Pd catalyst, or the particular bridging lig<strong>and</strong>) is missing. Interestingly,<br />

the same compound where the bridging lig<strong>and</strong> between the two metal sites is<br />

a bipyrimidine unit does not evolve molecular hydrogen. It is also reported<br />

that the amount <strong>of</strong> photocatalytically formed hydrogen depends strongly on<br />

the TEA concentration <strong>and</strong> the exposure time, <strong>and</strong> chloride ions inhibit the<br />

reaction. The amount <strong>of</strong> hydrogen produced increases steadily <strong>and</strong> levels <strong>of</strong>f<br />

after 1200 min. After about 1800 min no more hydrogen is produced. The<br />

rate <strong>of</strong> hydrogen formation increases with increasing TEA concentration for<br />

low TEA concentration, but becomes independent at a TEA concentration<br />

> 0.86 mol L –1 , where it is about 1600 nmol min –1 . Although detailed mechanistic<br />

data are not available, the authors suggest as a first step <strong>of</strong> the process<br />

a tw<strong>of</strong>old photoinduced reduction <strong>of</strong> the compound by TEA, analogously to<br />

what was reported for the related photoinduced electron collection system 51.<br />

Probably reduction is concomitant with proton extraction from TEA oxidation<br />

products: TEA should therefore be the proton source. The successive step<br />

should be reduction <strong>of</strong> the protons at the nearby Pd center. This latter step probably<br />

passes through a temporary chloride loss. The same paper also reports the<br />

photocatalyzed selective hydrogenation <strong>of</strong> tolane to cis-stilbene, accomplished<br />

by the same compound. Analogously to 60, the Ru(II) chromophore <strong>of</strong> 61 acts<br />

as the light harvester/photosensitizer (which also contains in its structure the<br />

electron acceptor subunit, that is, the phenazine moiety <strong>of</strong> the bridging lig<strong>and</strong>),<br />

while the role <strong>of</strong> the catalytic unit is here played by the Pd(II) center.<br />

Molecular hydrogen evolution under visible light irradiation has also been<br />

reported for trimetallic species like 53 [348]. Mechanistic details are not available.<br />

5.8.2<br />

Other Photocatalytic Systems<br />

The catalytic potential <strong>of</strong> heterometallic species containing Ru(II) <strong>and</strong> Re(I)<br />

chromophores for the conversion <strong>of</strong> CO2 to CO has been recently shown [349].<br />

Compound 62 is one <strong>of</strong> the species in this regard. This investigation highlighted<br />

the fact that the photocatalytic activity is deeply influenced by the nature <strong>of</strong><br />

both the bridging lig<strong>and</strong> <strong>and</strong> the peripheral lig<strong>and</strong>s at the light-harvesting<br />

Ru(II) chromophore. The proposed mechanism is that upon light irradiation<br />

(λ > 480 nm) in DMF/triethanolamine (TEOA, acting as base) with 1-benzyl-<br />

1,4-dihydronicotinamide (BNAH) as sacrificial donor, the initially produced<br />

Ru-based MLCT state is reduced by BNAH. Then intrabridging lig<strong>and</strong> electron<br />

transfer occurs, with formation <strong>of</strong> the reduced rhenium subunit. This latter<br />

speciesisknowntoreactwithCO2 upon Cl lig<strong>and</strong> loss [350, 351]. The reduction<br />

<strong>of</strong> CO2 is bielectronic, so it is assumed that the second electron transfer follows<br />

a similar route. The most efficient species <strong>of</strong> this series <strong>of</strong> compounds, which is<br />

exactly 62, exhibits a turnover number <strong>of</strong> 170.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 183<br />

Two other dinuclear species (63 <strong>and</strong> 64) have been reported to show photocatalytic<br />

activity for specific reactions. The Ru–Pd dimer 63 is active for the<br />

photocatalytic dimerization <strong>of</strong> 1-methylstyrene [352], <strong>and</strong> the turnover numbers<br />

<strong>of</strong> the photocatalyzed reaction (> 90 within 4 h) <strong>and</strong> the high selectivity<br />

compete well with thermal catalytic systems. Compound 64 is active for the<br />

conversion <strong>of</strong> trans-4-cyanostilbene to its cis form [353]. Other aspects <strong>of</strong> the<br />

last mentioned works <strong>and</strong> <strong>of</strong> similar systems are also commented on in a very<br />

recent paper [354]. Photocatalytic processes based on photoelectrochemical<br />

cells in which the Ru chromophores are physically interfaced to electrodes or<br />

other solid systems are reported later.<br />

5.9<br />

Photoactive Molecular Machines Able to Perform Nuclear Motions<br />

In the last 10 years there has been great interest in designing molecular<br />

machines [280]. As machines <strong>of</strong> the macroscopic world, even molecular machines<br />

need energy to operate, <strong>and</strong> a suitable form <strong>of</strong> energy to power


184 S. Campagna et al.<br />

molecular machines is light. It is therefore almost obvious that several photoactive<br />

molecular machines containing Ru(II) polypyridine complexes as<br />

the photoactive subunits have been prepared <strong>and</strong> studied. For an exhaustive<br />

discussion, the reader should consult [355–358]. A recent outst<strong>and</strong>ing<br />

example [359] is discussed Balzani et al. 2007, in this volume [119].<br />

We mention here the case <strong>of</strong> photoinduced ring motion in the catenane<br />

65 [360], illustrated in Fig. 17. Visible light excitation <strong>of</strong> this compound in<br />

acetonitrile leads to population <strong>of</strong> the MLCT triplet state <strong>and</strong> subsequent<br />

formation (via thermal activation) <strong>of</strong> the MC state which causes the decoordination<br />

<strong>of</strong> the sterically hindered bpy-type lig<strong>and</strong>. As a result, swinging <strong>of</strong><br />

the bpy-containing ring occurs, <strong>and</strong> the catenane structure made <strong>of</strong> discon-<br />

Fig. 17 Structural formula <strong>and</strong> photochemically <strong>and</strong> thermally induced motions <strong>of</strong><br />

a Ru(II) catenane complex


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 185<br />

nected rings is obtained (66). Heating regenerates the starting catenane. The<br />

process can be repeated at will, since both the reactions (coordination <strong>and</strong><br />

decoordination) are quantitative. It can be noted that in this system the photosubstitution<br />

process, usually considered as an undesired property for Ru(II)<br />

complexes, is instead used to perform the desired function. More recently,<br />

light-induced motion on a rotaxane system has also been obtained by using<br />

the same strategy [361].<br />

6<br />

Ruthenium Complexes <strong>and</strong> Biological Systems<br />

The effects <strong>of</strong> the interaction <strong>of</strong> photoactive ruthenium complexes with biological<br />

structures have been extensively studied. Because <strong>of</strong> the outst<strong>and</strong>ing<br />

excited-state properties <strong>of</strong> Ru(II) polypyridine complexes, these systems<br />

have been employed as probes <strong>of</strong> biological sites, as well as photocleavage<br />

agents <strong>and</strong>, in recent times, as inhibitors <strong>of</strong> biological functions [362–369].<br />

Among the species used as luminescent probes, one <strong>of</strong> the most studied compounds<br />

in the last 15 years is 67 [200–204, 362–379]. This complex is very<br />

weakly emissive in aqueous solution, but becomes strongly emitting in the<br />

presence <strong>of</strong> DNA, giving rise to the so-called light-switch effect [200, 201].<br />

The reason for such a behavior lies in the electronic properties <strong>of</strong> this specific<br />

chromophore <strong>and</strong> in the intercalation ability <strong>of</strong> the dipyrido[3,2-a:2 ′ ,3 ′ -<br />

c]phenazine (dppz) lig<strong>and</strong>. In this species, there are several triplet excited<br />

states quite close in energy: (1) a MLCT state directly populated by light excitation,<br />

in which the excited electron resides in the LUMO+1 centered on<br />

the “bpy-like” portion <strong>of</strong> dppz lig<strong>and</strong>; (2) a MLCT state in which the excited<br />

electron is located in the LUMO centered on the “phenazine-like” portion <strong>of</strong><br />

the dppz lig<strong>and</strong>; <strong>and</strong> (3) a lig<strong>and</strong>-centered (dppz-based) excited state. This<br />

compound represents another example <strong>of</strong> interplay between multiple MLCT<br />

states, discussed in more detail in Sect. 4.4. The energy gap <strong>and</strong> order <strong>of</strong> the<br />

three low-lying excited states mentioned above, as well as their dynamics,<br />

can be modulated by various parameters, including solvent dielectrics, protic<br />

ability <strong>of</strong> the solvent, <strong>and</strong> hydrophobic interactions. In a simplified schema-


186 S. Campagna et al.<br />

tization, in aqueous solution the dominant (lowest energy) excited state is<br />

the MLCT involving the phenazine-like dppz subunit (populated by a sort<br />

<strong>of</strong> “charge shift” decay from the directly excited MLCT level), which deactivates<br />

largely by nonradiative processes. When the complex interacts with<br />

DNA, the MLCT state involving the bpy-like dppz subunit becomes dominant,<br />

<strong>and</strong> since such a state has better luminescent properties, the luminescence<br />

<strong>of</strong> the complex is switched on. Looking in more detail, the interplay<br />

among the various states, in this <strong>and</strong> related species <strong>and</strong> in the absence<br />

<strong>and</strong> presence <strong>of</strong> DNA, is more complicated, as demonstrated by various theoretical<br />

[379] <strong>and</strong> experimental techniques, including transient absorption<br />

femtosecond spectroscopy [370, 371, 378, 380, 381] <strong>and</strong> time-resolved resonance<br />

Raman spectroscopy [372, 373, 382]. Details can be found in the original<br />

references.<br />

Besides 67, many other Ru(II) polypyridine complexes have been reported<br />

to exhibit luminescence enhancement in the presence <strong>of</strong> DNA. In most cases,<br />

the luminescence enhancement is moderate <strong>and</strong> can be assigned to the protection<br />

<strong>of</strong>fered toward oxygen quenching by DNA structures to the surfaceattached<br />

Ru(II) complex (which can bind, essentially for electrostatic reasons,<br />

to the major or minor grooves). If the interaction is limited to surface binding,<br />

the luminescence enhancement is usually within 20–40% in the presence<br />

<strong>of</strong> oxygen, whereas it is negligible in deoxygenated samples. Several compounds,<br />

however, exhibit noticeable luminescence enhancement (one order <strong>of</strong><br />

magnitude or higher): in most <strong>of</strong> these cases, the compounds quite <strong>of</strong>ten have<br />

a lig<strong>and</strong> with a large, flat framework <strong>and</strong> intercalation takes place. The compound<br />

68 [383], which is nonemissive in water solution <strong>and</strong> strongly emissive<br />

in organic solvents (λ = 610 nm; τ = 1.1 µs; Φ = 0.12) is an exception. In water,<br />

the presence <strong>of</strong> DNA switches the luminescence on. The authors suggest<br />

that in water solvent-specific interactions with the amido moiety promote<br />

radiationless decay <strong>of</strong> the (potentially) emitting MLCT state, <strong>and</strong> that the protection<br />

<strong>of</strong>fered by DNA versus solvent interaction restores MLCT emission.<br />

Ru(II) complexes whose luminescence is significantly quenched in the<br />

presence <strong>of</strong> DNA have also been reported. Usually, the excited state <strong>of</strong> these<br />

species is a very good oxidant, <strong>and</strong> photoinduced reductive electron transfer<br />

involving guanine residues is responsible for the luminescence quenching<br />

[362, 369, 384]. In some cases, it has been proposed that the quenching


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 187<br />

process can occur via photoinduced proton-coupled electron transfer with<br />

guanosine-5 ′ -monophosphate.<br />

Besides being used as luminescent probes, ruthenium complexes have<br />

been reported to form photoadducts with DNA <strong>and</strong> other species <strong>of</strong> biological<br />

relevance. The most studied photoadducts are probably the ones formed<br />

by Ru(II) complexes containing 1,4,5,8-tetraazaphenanthrene (TAP) as lig<strong>and</strong><br />

<strong>and</strong> guanine residues on DNA str<strong>and</strong>s (see for example 69) [386]. The<br />

mechanism <strong>of</strong> photoadduct formation has been extensively investigated. The<br />

initially formed MLCT state undergoes reductive electron transfer from guanine.<br />

This process is followed by fast formation <strong>of</strong> a covalent bond between<br />

the electron donor <strong>and</strong> acceptor, which leads to an adduct between the metallic<br />

complex <strong>and</strong> the nucleobase [386]. Such photoadduct formation has also<br />

been used to induce photocrosslinks between two nucleotide str<strong>and</strong>s when<br />

one <strong>of</strong> the str<strong>and</strong>s was chemically derivatized by the photoreactive metal<br />

complex <strong>and</strong> the complementary str<strong>and</strong> contained a guanine base in the<br />

proximity <strong>of</strong> the tethered complex [387]. The necessary requirement for this<br />

photoreaction to occur is a MLCT excited state which is a very strong oxidant,<br />

as guaranteed by the TAP lig<strong>and</strong>.<br />

More recently, a photoadduct between similar Ru complexes <strong>and</strong> the<br />

amino acid tryptophan have also been reported [388]. The authors mention<br />

that this photoreaction appears very promising for a wide range <strong>of</strong> applications<br />

to peptides <strong>and</strong> proteins.<br />

Ru(II) complexes have also been inserted into synthetic oligonucleotides to<br />

obtain specific information on the properties <strong>of</strong> DNA str<strong>and</strong>s <strong>and</strong>/or to prepare<br />

particular (super)structures [389–393]. For example, Ru(II)-derivatized<br />

oligonucleotides have been used to investigate the distance dependence <strong>of</strong> the<br />

quenching <strong>of</strong> suitable Ru luminescence by guanine residues [393]. Oligonucleotide<br />

conjugates containing Ru(II) polypyridine units as photosensitizers<br />

have also been reported to induce photodamage on single-str<strong>and</strong>ed DNA<br />

sites [394].<br />

The potential <strong>of</strong> Ru(II)-derivatized oligonucleotides has been explored<br />

to synthesize novel, interesting, <strong>and</strong> beautiful nanometer-sized luminescent<br />

structures in which the DNA str<strong>and</strong>s act as templates <strong>and</strong> the Ru complexes<br />

act as both template <strong>and</strong> photoactive units [395–397], giving rise to


188 S. Campagna et al.<br />

3D-networked structures. In these systems, the Ru(II) polypyridine subunits<br />

carry their own photoluminescence properties in the networked assemblies.<br />

Finally, it has been demonstrated that the photoexcitation <strong>of</strong> suitable<br />

Ru(II) complexes can inhibit biological functions; for example, in Ru(II)labeled<br />

oligonucleotides DNA polymerase is inhibited by a photocrosslinking<br />

process [387]. On the basis <strong>of</strong> these <strong>and</strong> similar results, which indicate strong<br />

<strong>and</strong> even complete inhibition <strong>of</strong> gene transcription by photoexcited Ru(II)<br />

complexes [398], it has been proposed that properly designed compounds<br />

can be ideal c<strong>and</strong>idates for a phototherapy with implemented fiber-optic light<br />

source [398–400]. It should be noted that the photoactivity <strong>of</strong> Ru(II) complexes<br />

for phototherapy does not depend on the presence <strong>of</strong> oxygen: this<br />

could represent a real advantage as compared to other dyes used in photodynamic<br />

therapy [398].<br />

7<br />

Dye-Sensitized Photoelectrochemical Solar Cells<br />

One <strong>of</strong> the most important developments involving Ru(II) polypyridine complexes<br />

in the last two decades is related to the design <strong>of</strong> dye-sensitized photoelectrochemical<br />

solar cells, which have outst<strong>and</strong>ing properties for application<br />

in the field <strong>of</strong> solar energy conversion, in particular photovoltaics. Since their<br />

appearance in the early 1990s [401, 402], dye-sensitized photoelectrochemical<br />

solar cells based on the principle <strong>of</strong> sensitization <strong>of</strong> wide-b<strong>and</strong>gap mesoporous<br />

semiconductors have indeed attracted the interest <strong>of</strong> the scientific<br />

community, due to their performances which started the vision <strong>of</strong> a promising<br />

alternative to conventional junction-based photovoltaic devices. For the<br />

first time a solar energy device operating on a molecular level showed the stability<br />

<strong>and</strong> the efficiency required for potential practical applications. In the<br />

last few years, several excellent review articles have been published in this<br />

field [403–405]. These articles indicated that research on dye-sensitized solar<br />

cells is strongly multidisciplinary, involving areas such as nanotechnology,<br />

materials science, interfacial electron transfer, <strong>and</strong> supramolecular photochemistry<br />

<strong>and</strong> electrochemistry [405]. Here we mention the main basic aspects<br />

<strong>and</strong> describe a few <strong>of</strong> recent Ru(II) photosensitizers which exhibit quite<br />

interesting performances.<br />

7.1<br />

General Concepts<br />

The principle <strong>of</strong> dye sensitization <strong>of</strong> semiconductors can be traced back to<br />

the end <strong>of</strong> the 1960s [406]. However, the practical use <strong>of</strong> these systems was<br />

limited for a long time because the efficiencies obtained with single-crystal<br />

substrates were too low due to the poor light absorption <strong>of</strong> the adsorbed


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 189<br />

monolayer <strong>of</strong> dye molecules. The breakthrough in the field was brought about<br />

by the introduction <strong>of</strong> mesoscopic films made <strong>of</strong> sintered nanoparticles <strong>of</strong><br />

a semiconductor metal oxide with a large surface area, which allowed the<br />

adsorption, at monolayer coverage, <strong>of</strong> a much larger number <strong>of</strong> sensitizer<br />

molecules leading to absorbance values, <strong>of</strong> thin films <strong>of</strong> a few microns, well<br />

above unity [407].<br />

Wide-b<strong>and</strong>gap semiconductor materials such as TiO2 show a separation<br />

between the energy levels <strong>of</strong> the valence <strong>and</strong> conduction b<strong>and</strong>s <strong>of</strong> the order<br />

<strong>of</strong> 3 eV, which means that the electron–hole pair needs to be produced by irradiation<br />

with light having a wavelength shorter than 400 nm, i.e., UV light.<br />

In order to use sunlight, mainly visible <strong>and</strong> near-IR, two general approaches<br />

have been developed: doping <strong>and</strong> molecular sensitization. The doping approach<br />

is the preferred choice for conventional photovoltaic devices [405].<br />

Dye-sensitized photelectrochemical cells rely on photosensitization. In this<br />

process, a photoexcited species, the sensitizer (S), is capable <strong>of</strong> injecting an<br />

electron into the conduction b<strong>and</strong> (CB) or a hole into the valence b<strong>and</strong> (VB)<br />

<strong>of</strong> the semiconductor (Fig. 18).<br />

Fig. 18 Schemes for sensitized charge injection in the photoelectrochemical solar cells:<br />

a electron injection, b hole injection<br />

In fact, when the excited-state energy level <strong>of</strong> the sensitizer is higher with<br />

respect to the bottom <strong>of</strong> the conduction b<strong>and</strong>, an electron can be injected<br />

with no thermal activation barrier in the semiconductor, leaving the sensitizer<br />

in its one-electron oxidized form (Fig. 18a). When the excited state<br />

is lower in energy with respect to the top <strong>of</strong> the valence b<strong>and</strong>, an electron<br />

transfer (formally a hole transfer) between the semiconductor <strong>and</strong> the sensitizer<br />

can take place, leaving the molecule in its one-electron reduced form<br />

(Fig. 18b) [408].<br />

The operation <strong>of</strong> a dye-sensitized solar cell is schematized in Fig. 19 [403,<br />

405]. The system is comprised <strong>of</strong> two facing electrodes, a photoanode <strong>and</strong><br />

a counter electrode, with an electrolyte in between. The transparent conductive<br />

photoanode is covered with a thin film (7–10 µm) <strong>of</strong> a mesoporous


190 S. Campagna et al.<br />

Fig. 19 Schematic operation principle <strong>of</strong> a dye-sensitized solar cell<br />

semiconductor oxide obtained via a sol–gel procedure. Dye coverage <strong>of</strong> semiconductor<br />

nanoparticles is generally obtained from alcoholic solutions <strong>of</strong> the<br />

sensitizer, in which the sintered film is left immersed for a few hours. Sensitizers<br />

are usually designed to have functional groups such as – COOH, – PO3H2,<br />

or – B(OH)2 for stable adsorption onto the semiconductor substrate. The dyecovered<br />

film is in intimate contact with an electrolytic solution containing<br />

a redox couple dissolved in a suitable solvent. The electron donor member <strong>of</strong><br />

the redox couple must reduce quickly <strong>and</strong> quantitatively the oxidized sensitizer,<br />

so closing the circuit. A variety <strong>of</strong> solvents with different viscosity <strong>and</strong> <strong>of</strong><br />

redox mediators have been the object <strong>of</strong> intense studies, the most commonly<br />

used being the couple I – 3 /I– in acetonitrile or methoxypropionitrile solution.<br />

The counter electrode is a conductive glass covered with a few clusters <strong>of</strong><br />

metallic platinum, which has a catalytic effect in the reduction process <strong>of</strong> the<br />

electron mediator. Further details on the cells <strong>and</strong> on their preparation can be<br />

found in the literature [403–405].<br />

The complete photoelectrochemical cycle <strong>of</strong> the device can be outlined as<br />

follows. The adsorbed sensitizer molecules (S) are brought into their excited<br />

state (S∗ ) by photon absorption <strong>and</strong> inject one electron into the empty conduction<br />

b<strong>and</strong> <strong>of</strong> the semiconductor in a timescale <strong>of</strong> femtoseconds. Injected<br />

electrons percolate through the nanoparticle network <strong>and</strong> are collected by the<br />

conductive layer <strong>of</strong> the photoanode electrode, while the oxidized sensitizer<br />

(S + ) in its ground state is rapidly reduced by I – ions in solution. Photoinjected<br />

electrons flow in the external circuit where useful electric work is produced<br />

<strong>and</strong> are available at the counter electrode for the reduction <strong>of</strong> the electron<br />

mediator acceptor I – 3 . The entire cycle consists in the quantum conversion <strong>of</strong><br />

photons to electrons.<br />

S+hν → S∗ photoexcitation (25)<br />

S∗ +TiO2→S + +(e – ,TiO2) electron injection (26)<br />

2S + +3I – → 2S+I3 – sensitizer regeneration (27)<br />

I3 – +2e – → 3I – electron donor regeneration . (28)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 191<br />

Photoinjected electrons should escape from any recombination process in<br />

order to have a unit charge collection efficiency at the photoelectrode back<br />

contact. The two major waste processes in a dye-sensitized solar cell are<br />

due to (1) back electron transfer, at the semiconductor/electrolyte interface,<br />

between electrons in the conduction b<strong>and</strong> <strong>and</strong> the oxidized dye molecules<br />

(Eq. 29), <strong>and</strong> (2) reduction <strong>of</strong> the electron relay (I – 3 , in this case) at the semiconductor<br />

nanoparticle surface (Eq. 30).<br />

S + +(e – ,TiO2) → S back electron transfer (29)<br />

I3 – +2(e – ,TiO2) → 3I – electron capture from mediator . (30)<br />

A detailed knowledge <strong>of</strong> all the kinetic mechanisms occurring in a photoelectrochemical<br />

cell under irradiation is an essential feature toward optimization<br />

<strong>of</strong> the process.<br />

7.2<br />

Ruthenium-Sensitized Photoelectrochemical Solar Cells<br />

A major breakthrough in the field relied on the performance <strong>of</strong> dye-sensitized<br />

solar cells employing Ru(II) complexes as sensitizers [401–405, 409–411].<br />

Several reasons are at the basis <strong>of</strong> the success <strong>of</strong> Ru(II) polypyridine complexes<br />

in playing this leading role:<br />

1. Strong absorption throughout all the visible region, which can also extend<br />

to the near-IR. This result is obtained by means <strong>of</strong> intense MLCT b<strong>and</strong>s<br />

due to a judicious choice <strong>and</strong> combination <strong>of</strong> lig<strong>and</strong>s [1].<br />

2. Strong electronic coupling between the MLCT excited state <strong>of</strong> the chromophore<br />

<strong>and</strong> the semiconductor conduction b<strong>and</strong>. To fulfill this requirement,<br />

it has to be noted that the polypyridine lig<strong>and</strong> connected to the<br />

semiconductor via suitable functionalization <strong>of</strong> the lig<strong>and</strong> (usually carboxylated<br />

lig<strong>and</strong>s) must be that involved in the lowest-lying MLCT state.<br />

3. Tunability <strong>of</strong> the excited-state redox properties. This allows the preparation<br />

<strong>of</strong> compounds whose excited-state oxidation potential can ensure an<br />

efficient electron injection in the semiconductor conduction b<strong>and</strong>. In this<br />

regard, it should be considered that to estimate a “reduction potential”<br />

(Ecb) for the semiconductor conduction b<strong>and</strong> is not an easy task [404, 412–<br />

414], <strong>and</strong> in nonaqueous solvents adsorption <strong>of</strong> cations, which are present<br />

as electrolytes, also has a significant effect on Ecb values. For example,<br />

Ecb for nanostructured TiO2 has been reported to be – 1.0 VvsSCEin<br />

0.1M LiClO4/acetonitrile <strong>and</strong> about – 2.0 VwhenLi + cations are replaced<br />

by tetrabutylammonium [413, 414].<br />

4. Stability <strong>of</strong> the Ru(II) polypyridine complexes, in the ground state as well<br />

as in the excited <strong>and</strong> redox states. However, it is useful to note that photostability<br />

is not a strict requisite here, since the excited state is rapidly<br />

deactivated by electron injection. The same applies to chromophores hav-


192 S. Campagna et al.<br />

ing intrinsic short excited-state lifetimes, provided that their intrinsic<br />

lifetime (that is, the reverse <strong>of</strong> the summation <strong>of</strong> the radiative <strong>and</strong> radiationless<br />

decay <strong>of</strong> the “isolated” chromophores) does not compete with the<br />

timescale for photoinjection.<br />

The electronic coupling between MLCT states <strong>and</strong> TiO2 conduction b<strong>and</strong><br />

is so efficient that electron injection takes place in the ultrafast regime.<br />

In particular, for the complex [Ru(dcbpy)2(NCS)2] (dcbpy=4,4 ′ -carboxylbipyridine),<br />

also called N3 or “red dye” (70), biphasic kinetics was reported<br />

for photoinduced electron injection [266, 267, 270]: the first ultrafast component<br />

was estimated to have a risetime <strong>of</strong> 28 fs <strong>and</strong> the slower component was<br />

reported to occur within the 1–50 ps time range [270, 271]. This behavior was<br />

initially interpreted on the basis <strong>of</strong> a two-state mechanism, the fast <strong>and</strong> slow<br />

components being attributed to injection from the MLCT singlet <strong>and</strong> triplet<br />

states, respectively. Also, the singlet-state injection was from nonthermalized<br />

vibronic states. Whereas injection from the singlet state was later confirmed<br />

(although with risetime slightly different, shorter than 20 fs), the origin <strong>of</strong> the<br />

picosecond component has been questioned: it has been recently proposed<br />

that such a “slow” component <strong>of</strong> electron injection arises from sensitizer<br />

molecules which are loosely attached to the semiconductor or are present in<br />

aggregated forms [271].<br />

Dozens <strong>of</strong> Ru(II) complexes have been explored as sensitizers in dyesensitized<br />

solar cells <strong>of</strong> this type. We mention here some <strong>of</strong> the species<br />

exhibiting the best performances. The above mentioned N3 complex has<br />

very interesting properties: it shows a photoaction spectrum dominating almost<br />

the entire visible region, with incident photon-to-current conversion<br />

efficiency (IPCE) <strong>of</strong> the order <strong>of</strong> 90% between 500–600 nm. Short-circuit photocurrents<br />

exceeding 17 mA/cm 2 in simulated A.M. 1.5 sunlight <strong>and</strong> opencircuit<br />

photovoltages <strong>of</strong> the order <strong>of</strong> 0.7 V were obtained by using the couple<br />

I – 3 /I– as redox electrolyte [415]. For the first time a photoelectrochemical device<br />

was found to give an overall conversion efficiency <strong>of</strong> 10%. These performances,<br />

in part expected for the high reducing ability <strong>of</strong> the 3 MLCT state (ca.<br />

– 1 eV vs SCE) <strong>and</strong> the positive ground-state oxidation potential (+ 0.85 eV


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 193<br />

vs SCE), contrast with the lower IPCE observed for other sensitizers having<br />

comparable ground- <strong>and</strong> excited-state properties [416]. It has been suggested<br />

that a peculiar molecular level property <strong>of</strong> the cis-[Ru(dcbpy)2(NCS)2] complex<br />

could affect one <strong>of</strong> the key processes <strong>of</strong> the cell mechanism. This view is<br />

consistent with the results <strong>of</strong> photoelectron spectroscopy <strong>and</strong> INDO/S calculations<br />

indicating that the Ru d orbitals interact strongly with the π orbitals<br />

<strong>of</strong> NCS, resulting in MOs <strong>of</strong> mixed nature [417]. In particular, the calculations<br />

show that the sulfur 3p orbitals give a considerable contribution to the<br />

HOMO <strong>of</strong> the complex. Hole delocalization across the NCS lig<strong>and</strong>s can thus<br />

be responsible for an increased electronic matrix element for the electron<br />

transfer reaction involving TiO2/Ru III NCS <strong>and</strong> I – . This would lead to an increase<br />

<strong>of</strong> the rate constant <strong>of</strong> the reductive process, <strong>and</strong> as a consequence<br />

<strong>of</strong> IPCE.<br />

Even better photoelectrochemical performances, compared to those given<br />

by the complex [Ru(dcbpy)2(NCS)2], are featured by a species based on<br />

the terpyridine lig<strong>and</strong> [418]: TiO2 electrodes covered with the complex<br />

[Ru(L)(NCS)3] (L=4,4 ′ ,4 ′′ -tricarboxy-2,2 ′ :6 ′ ,2 ′′ -terpyridine) display very efficient<br />

panchromatic sensitization covering the whole visible spectrum <strong>and</strong><br />

extend the spectral response at the near-IR region up to 920 nm, with maximum<br />

IPCE values comparable to that obtained with the dithiocyanate complex.<br />

Another species based on a substituted terpyridine is the mixed lig<strong>and</strong><br />

complex [Ru(HP-terpy)(dmb)(NCS)], where P-terpy = 4-phosphonato-<br />

2,2 ′ :6,2 ′′ -terpyridine <strong>and</strong> dmb = 4,4 ′ -dimethyl-2,2 ′ -bipyridine [419]. A quantitative<br />

study <strong>of</strong> dye adsorption on TiO2 has shown that complexes containing<br />

the phosphonated terpyridine lig<strong>and</strong> adsorb more efficiently <strong>and</strong> strongly,<br />

giving an adsorption constant about 80 times larger than that for the dicarboxy<br />

bipyridine compounds. Since one <strong>of</strong> the problems encountered with the<br />

carboxy polypyridine class <strong>of</strong> sensitizers is the desorption from the semiconductor<br />

surface in the presence <strong>of</strong> water, the search for new anchoring<br />

groups is advisable. Along this line <strong>of</strong> research, complexes based on the<br />

derivatization <strong>of</strong> 2,2 ′ -bipyridine with a phenylboronic functionality were prepared.<br />

The photoaction spectra <strong>of</strong> TiO2 electrodes sensitized with the [Ru(4phenylboronic-2,2<br />

′ -bipyridine)2(CN)2] complex showed IPCE values comparable<br />

to those observed for [Ru(dcbH2)2(CN)2], indicating that the new type<br />

<strong>of</strong> linkage does not reduce the electronic coupling between sensitizer <strong>and</strong><br />

semiconductor [405].<br />

7.3<br />

Supramolecular Sensitizers<br />

Besides mononuclear Ru(II) complexes, multinuclear (supramolecular) compounds,<br />

as well as chromophore–acceptor or chromophore–donor dyads<br />

made <strong>of</strong> Ru(II) species <strong>and</strong> organic quenchers, have been used as sensitizers.<br />

There are several aims that inspired the design <strong>of</strong> such systems:


194 S. Campagna et al.<br />

1. To increase the absorption properties <strong>of</strong> the (multicomponent) sensitizer,<br />

by using systems featuring the antenna effect, with the energy trap <strong>of</strong> the<br />

antenna being the Ru(II) unit directly connected to the semiconductor.<br />

One example is the trinuclear Ru(II) species 71 showninFig.20;in71,the<br />

light absorbed by the peripheral Ru(II) chromophores is transferred quantitatively<br />

to the central Ru(II) chromophore, from which electron injection<br />

takes place [420]. Experiments on this complex adsorbed on polycrystalline<br />

TiO2 gave an overall conversion efficiency <strong>of</strong> ca. 7% withturnover<br />

numbers <strong>of</strong> at least 1 × 10 6 . The antenna effect is expected to be <strong>of</strong> relevance<br />

for applications requiring very thin TiO2 layers.<br />

2. To spatially separate the injected electron <strong>and</strong> the hole on the sensitizer,<br />

so decreasing losses due to charge recombination. A system designed for<br />

this aim (72) is shown in Fig. 21, where the possible electron transfer steps<br />

are indicated [421]. In 72, the MLCT state <strong>of</strong> the Ru(II) unit is rapidly<br />

quenched by the Rh(III) species (step k1 in Fig. 21), followed by injection<br />

<strong>of</strong> the electron onto the semiconductor conductance b<strong>and</strong> from the reduced<br />

Rh unit (step k2, in competition with step k4). The recombination<br />

process (k5) is slow because <strong>of</strong> a very weak electronic coupling. To reach<br />

thesameaim,thespecies73 shown in Fig. 22 has been designed [422].<br />

Here, the first step is the electron injection from the Ru unit. Then, electron<br />

transfer from the phenothiazine unit to the oxidized Ru unit takes<br />

place, resulting in a charge-separated species which decays to the ground<br />

state with a rate <strong>of</strong> 3.6 × 10 3 s –1 (lifetime, 0.3 ms). The dyad <strong>and</strong> model<br />

molecules were also tested in solar cells, with iodide as an electron donor.<br />

While the observed IPCE was <strong>of</strong> the order <strong>of</strong> 45% for both systems, the<br />

open circuit photovoltage was higher for the dyad by 100 mV. The effect<br />

was more pronounced in the absence <strong>of</strong> iodide with Voc = 180 mV. Applying<br />

the measured interfacial electron transfer rates to the diode equation<br />

Fig. 20 Structural formula <strong>of</strong> a branched antenna system <strong>and</strong> schematization <strong>of</strong> energy<br />

transfer <strong>and</strong> charge injection in TiO2. Complex charge is omitted for clarity


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 195<br />

Fig. 21 Structural formula <strong>of</strong> a linear antenna system <strong>and</strong> schematization <strong>of</strong> energy transfer<br />

<strong>and</strong> charge injection in TiO2. Complex charge is omitted for clarity<br />

Fig. 22 Structural formula <strong>of</strong> TiO2/Ru-PTZ heterotriad system <strong>and</strong> schematization <strong>of</strong> the<br />

electron transfer steps. Complex charge is omitted for clarity


196 S. Campagna et al.<br />

gave the predicted increase <strong>of</strong> Voc <strong>of</strong> 200 mV, which was in agreement<br />

with the obtained value (180 mV) [422]. It is interesting that an increase<br />

<strong>of</strong> the lifetime <strong>of</strong> the interfacial charge-separated state TiO2(e – )–Ru(II)–<br />

PTZ + has a direct influence on the overall efficiency <strong>of</strong> the cell. A similar<br />

approach inspired the design <strong>of</strong> the supramolecular species 74, basedon<br />

the “red dye” N3 sensitizer. Optical excitation <strong>of</strong> a nanocrystalline TiO2<br />

film dye coated with such a species showed a long-lived charge-separated<br />

state [423].<br />

8<br />

Miscellanea<br />

The fields that have been recently powered by Ru photochemistry are much<br />

more than those reported in some detail in this article. A few <strong>of</strong> those that are<br />

not discussed above will be briefly mentioned.<br />

Ru(II)-based chromophores have been linked to a plethora <strong>of</strong> receptor<br />

species, like calixarenes, crowns, <strong>and</strong> azacrowns, essentially for sensing<br />

purposes [280, 424, 425]. Ru(II) chromophores have also been embedded in<br />

oxygen-permeating polymers to yield luminescent sensors for molecular oxygen<br />

determination in atmosphere [426–429]. New systems have been designed<br />

<strong>and</strong> studied for obtaining OLED materials. In this regard, a dinuclear<br />

Ru complex has been used in conjunction with an organic luminophore to<br />

generate two-color electroluminescence [430].<br />

Multichromophoric species made <strong>of</strong> Ru(II) chromophores interfaced with<br />

organic aromatics having suitable triplet-state levels have been studied to extend<br />

the lifetime <strong>of</strong> the MLCT excited state by a sort <strong>of</strong> delayed luminescence<br />

involving intercomponent energy transfer, with the organic triplet states used<br />

as excited-state energy storage systems [431–440, 442]. A few such species are<br />

compounds 75 (which is the first reported example <strong>of</strong> such a behavior [431]),<br />

76 [437], <strong>and</strong> 77 [435]. Compound 77 is one <strong>of</strong> the species featuring the most<br />

outst<strong>and</strong>ing behavior: its emission in fluid solution at room temperature<br />

(with maximum at about 600 nm) has lifetimes ranging from 43 µs (acetone<br />

solution) to 61 µs (acetonitrile)to115 µs (DMSO solution) [435]. With<br />

the aim <strong>of</strong> increasing the excited-state lifetime as well as the luminescence


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 197<br />

quantum yield, a goal which is missed by the formerly mentioned multichromophoric<br />

species, several Ru(II) chromophores having polypyridine lig<strong>and</strong>s<br />

able to feature extended electron delocalization in their structure have been<br />

prepared [441]. In this case, the improved photophysical properties are due to<br />

reduced Franck–Condon factors for radiationless decay, a consequence <strong>of</strong> extended<br />

delocalization <strong>of</strong> the emitting MLCT state [78, 157, 443–446]. A typical<br />

example is the compound 78 shown in Fig. 23 [445]. In 78, in spite <strong>of</strong> the quite<br />

low emission energy at room temperature (820 nm), a relatively long luminescence<br />

lifetime <strong>and</strong> high quantum yield are found (420 ns <strong>and</strong> about 0.01,<br />

respectively).<br />

Hydrogen-bonded or, generally, noncovalently linked supramolecular<br />

species exhibiting photoinduced electron <strong>and</strong>/or energy transfer have also<br />

been prepared to mimic natural systems (see, for example, 79–81) [447–452].<br />

Efficient intercomponent energy transfer through the noncovalently linked<br />

frameworks is usually obtained. However, in some cases the interest in the<br />

potential application <strong>of</strong> these systems is reduced by the small value <strong>of</strong> the<br />

association constants.


198 S. Campagna et al.<br />

Fig. 23 Thermal ellipsoid views <strong>and</strong> structural formula <strong>of</strong> a dinuclear Ru(II) complex<br />

Multiple emissions from Ru(II) polypyridine complexes have been reported.<br />

Besides multiple emission connected to supramolecular species featuring<br />

nonquantitative interchromophoric energy transfer (a relatively common<br />

<strong>and</strong> in some way an expected behavior), in some cases multiple emission<br />

from the same Ru(II) subunit has also been proposed at room temperature<br />

[453]. This behavior has not been fully explained yet.<br />

Many efforts have aimed to take advantage <strong>of</strong> the photophysical properties<br />

<strong>of</strong> Ru(II) chromophores in electropolymerized thin film structures <strong>and</strong>


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 199<br />

in SiO2-based sols–gels [342]. In some cases, these solid-interfaced systems<br />

allowed interesting photocatalytic results to be obtained, such as (1) the dehydrogenation<br />

<strong>of</strong> 2-propanol to acetone <strong>and</strong> molecular hydrogen, obtained<br />

in a photoelectrochemical cell in which a dinuclear Ru complex is adsorbed


200 S. Campagna et al.<br />

to TiO2 [454], <strong>and</strong> (2) the dehydrogenation <strong>of</strong> hydroquinone to quinone, by<br />

means <strong>of</strong> two Ru chromophores coadsorbed on TiO2 at the anode <strong>of</strong> a photoelectrochemical<br />

cell, with molecular hydrogen produced at a platinizedplatinum<br />

cathode [455]. These fields have been recently reviewed <strong>and</strong> several<br />

details can be found in [342]. Quite recently photoinduced charge transfer between<br />

CdSe nanocrystal quantum dots <strong>and</strong> Ru(II) complexes has also been<br />

reported [456].<br />

Interesting results, from the photocatalysis point <strong>of</strong> view, have been reported<br />

for other heterogeneous systems. For example, the sensitization <strong>of</strong> platinized<br />

layered metal oxide semiconductors with Ru(II) polypyridine dyes enabled<br />

photolysis <strong>of</strong> aqueous hydrogen iodide to molecular hydrogen <strong>and</strong> triiodide<br />

using visible light [457–459]. Photocatalytic water oxidation has been<br />

accomplished with good efficiency in a system based on [Ru(bpy)3] 2+ as the<br />

photosensitizer <strong>and</strong> using a colloidal solution <strong>of</strong> IrO2 as a catalyst [460, 461].<br />

Ru(II) polypyridine complexes containing lig<strong>and</strong>s which can be protonated/deprotonated<br />

have been extensively studied. Interesting effects <strong>of</strong> electronic<br />

coupling between metal centers in dinuclear systems with protonable/deprotonable<br />

bridging lig<strong>and</strong>s have been reported [462–464], as well<br />

improved emissive properties in Ru(terpy)2-like complexes [188]. Quite interesting<br />

light-controlled electronic coupling between the Ru(II) centers <strong>of</strong><br />

dinuclear systems has also been obtained by inserting photoisomerizable<br />

subunits with the bridging lig<strong>and</strong>s [465, 466].<br />

Finally, the recently published X-ray time-resolved spectra <strong>of</strong> [Ru(bpy)3] 2+ ,<br />

obtained by different techniques [467–469], warrants recalling. These studies<br />

give information on the structural changes induced by excitation. Interestingly,<br />

they confirm the very small distortion <strong>of</strong> the 3 MLCT state in comparison<br />

with the ground state, as formerly predicted on the basis <strong>of</strong> other<br />

experimental results <strong>and</strong> theoretical arguments.<br />

Acknowledgements We acknowledge MIUR (PRIN projects nos. 2006034123 <strong>and</strong> 2006-<br />

030320), the University <strong>of</strong> Bologna, <strong>and</strong> the University <strong>of</strong> Messina for financial support.<br />

We also wish to thank our colleagues F. Barigelletti, L. Hammarström, <strong>and</strong> C.A. Bignozzi<br />

for useful discussions.<br />

References<br />

1. Juris A, Balzani V, Barigelletti F, Campagna S, Belser P, von Zelewsky A (1988) Coord<br />

Chem Rev 84:85<br />

2. Balzani V, Juris A, Venturi M, Campagna S, Serroni S (1996) Chem Rev 96:759<br />

3. Sauvage JP, Collin JP, Chambron JC, Guillerez S, Coudret C, Balzani V, Barigelletti F,<br />

De Cola L, Flamigni L (1994) Chem Rev 94:993<br />

4. Meyer TJ (1986) Pure Appl Chem 58:1193<br />

5. Crosby GA, Watts RJ, Carstens DHW (1970) Science 170:1195<br />

6. Crosby GA (1975) Acc Chem Res 8:231


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 201<br />

7. Demas JN (1983) J Chem Educ 60:803<br />

8. Chen Y, Meyer TJ (1998) Chem Rev 98:1439<br />

9. Crosby GA (1976) Adv Chem Ser 150:149<br />

10. Kober EM, Meyer TJ (1982) Inorg Chem 21:3967<br />

11. M<strong>and</strong>al K, Pearson TLD, Krug WP, Demas JN (1983) J Am Chem Soc 105:701<br />

12. Rillema DP, Jones DS, Levy HA (1979) J Chem Soc Chem Commun, p 849<br />

13. Orgel LE (1961) J Chem Soc 3683<br />

14. Kalyanasundaram K (1982) Coord Chem Rev 46:159<br />

15. Seddon EA, Seddon KR (1984) The chemistry <strong>of</strong> ruthenium, chap 15. Elsevier, Amsterdam<br />

16. Watts RJ (1983) J Chem Educ 60:834<br />

17. Balzani V, Carassiti V (1970) <strong>Photochemistry</strong> <strong>of</strong> coordination compounds. Academic,<br />

London<br />

18. Adamson AW (1983) J Chem Educ 60:797<br />

19. Fleischauer PD, Fleischauer P (1970) Chem Rev 70:199<br />

20. DeArmond MK, Carlin CM (1981) Coord Chem Rev 36:325<br />

21. DeArmond MK, Hanck KW, Wertz DW (1985) Coord Chem Rev 64:65<br />

22. Ohsawa Y, DeArmond MK, Hanck KW, Morris DE (1983) J Am Chem Soc 105:6522<br />

23. Ohsawa Y, Whangho MH, Hanck KW, DeArmond MK (1984) Inorg Chem 23:3426<br />

24. Morris DE, Hanck KW, DeArmond MK (1983) J Am Chem Soc 105:3032<br />

25. Elliott CM (1980) J Chem Soc Chem Commun, p 261<br />

26. Heath GA, Yellowlees LJ, Braterman PS (1982) Chem Phys Lett 92:646<br />

27. Ohsawa Y, DeArmond MK, Hanck KW, Morel<strong>and</strong> CG (1985) J Am Chem Soc<br />

107:5383<br />

28. Angel SM, DeArmond MK, Donohoe RJ, Hanck KW, Wertz DH (1984) J Am Chem<br />

Soc 106:3688<br />

29. Dallinger RF, Woodruff WH (1979) J Am Chem Soc 101:4391<br />

30. Omberg KM, Schoonover JR, Treadway JA, Leasure LM, Dyer RB, Meyer TJ (1997)<br />

J Am Chem Soc 119:7013<br />

31. Balzani V, Bolletta F, G<strong>and</strong>olfi MT, Maestri M (1978) Top Curr Chem 75:1<br />

32. Lytle FE, Hercules DM (1969) J Am Chem Soc 91:253<br />

33. Klassen DM, Crosby GA (1968) J Chem Phys 48:1853<br />

34. Xie P, Chen YJ, Uddin J, Endicott JF (2005) J Phys Chem A 109:4671<br />

35. Kober EM, Sullivan BP, Meyer TJ (1984) Inorg Chem 23:2098<br />

36. Bensasson R, Salet C, Balzani V (1976) J Am Chem Soc 98:3722<br />

37. Bensasson R, Salet C, Balzani V (1979) CR Acad Sci Paris 289B:41<br />

38. Lachish U, Infelta PP, Grätzel M (1979) Chem Phys Lett 62:317<br />

39. Creutz C, Chou M, Netzel TL, Okumura M, Sutin N (1980) J Am Chem Soc 102:1309<br />

40. Bradley PC, Kress N, Hornberger BA, Dallinger RF, Woodruff WH (1981) J Am Chem<br />

Soc 103:7441<br />

41. Brauenstein CH, Baker AD, Strekas TC, Gafney HD (1984) Inorg Chem 23:857<br />

42. Demas JN, Crosby GA (1971) J Am Chem Soc 93:2841<br />

43. Demas JN, Taylor DG (1979) Inorg Chem 18:3177<br />

44. Bolletta F, Juris A, Maestri M, S<strong>and</strong>rini D (1980) Inorg Chim Acta 44:L175<br />

45. Harrigan RW, Crosby GA (1973) J Phys Chem 59:3468<br />

46. Hager GD, Crosby GA (1975) J Am Chem Soc 97:7031<br />

47. Hager GD, Watts RJ, Crosby GA (1975) J Am Chem Soc 97:7037<br />

48. Hipps KW, Crosby GA (1975) J Am Chem Soc 97:7042<br />

49. Crosby GA, Elfring WH Jr (1976) J Phys Chem 80:2206


202 S. Campagna et al.<br />

50. Sprouse S, King KA, Spellane PJ, Watts RJ (1984) J Am Chem Soc 106:6647<br />

51. Maestri M, S<strong>and</strong>rini D, Balzani V, Chassot L, Jolliet P, von Zelewsky A (1985) Chem<br />

Phys Lett 122:375<br />

52. Roundhill DM (1994) <strong>Photochemistry</strong> <strong>and</strong> photophysics <strong>of</strong> metal complexes.<br />

Plenum, New York<br />

53. Maestri M, Balzani V, Deuschel-Cornioley C, von Zelewsky A (1991) Adv Photochem<br />

17:1<br />

54. King KA, Spellane PJ, Watts RJ (1985) J Am Chem Soc 107:1431<br />

55. Hillis JE, De Armond MK (1971) J Lumin 4:273<br />

56. Shaw JR, Schmehl RH (1991) J Am Chem Soc 113:389<br />

57. Guglielmo G, Ricevuto V, Giannetto A, Campagna S (1989) Gazz Chim Ital 119:457<br />

58. Juris A, Campagna S, Bidd I, Lehn JM, Ziessel R (1988) Inorg Chem 27:4007<br />

59. Ohsawa Y, Sprouse S, King KA, DeArmond MK, Hanck KW, Watts RJ (1987) J Phys<br />

Chem 91:1047<br />

60. Maestri M, S<strong>and</strong>rini D, Balzani V, von Zelewsky A, Jolliet P (1988) Helv Chim Acta<br />

71:134<br />

61. Van Houten J, Watts RJ (1976) J Am Chem Soc 98:4853<br />

62. Van Houten J, Watts RJ (1978) Inorg Chem 17:3381<br />

63. Elfring WH Jr, Crosby GA (1981) J Am Chem Soc 103:2683<br />

64. Durham B, Caspar JV, Nagle JK, Meyer TJ (1982) J Am Chem Soc 104:4803<br />

65. Caspar JV, Sullivan BP, Kober EM, Meyer TJ (1982) Chem Phys Lett 91:91<br />

66. Calvert JM, Caspar JV, Binstead RA, Westmorel<strong>and</strong> TD, Meyer TJ (1982) J Am Chem<br />

Soc 104:6620<br />

67. Agnew SF, Stone ML, Crosby GA (1982) Chem Phys Lett 85:57<br />

68. Caspar JV, Meyer TJ (1983) J Am Chem Soc 105:5583<br />

69. Barigelletti F, Juris A, Balzani V, Belser P, von Zelewsky A (1983) Inorg Chem 22:3335<br />

70. Juris A, Barigelletti F, Balzani V, Belser P, von Zelewsky A (1985) Inorg Chem 24:1758<br />

71. Barigelletti F, Belser P, von Zelewsky A, Juris A, Balzani V (1985) J Phys Chem<br />

89:3680<br />

72. Wacholtz WF, Auerbach RA, Schmehl RH (1986) Inorg Chem 25:227<br />

73. Anderson S, Constable EC, Seddon KR, Turp JE, Baggott JE, Pilling MJ (1985) J Chem<br />

Soc Dalton Trans, p 2247<br />

74. Barigelletti F, Juris A, Balzani V, Belser P, von Zelewsky A (1986) J Phys Chem<br />

90:5190<br />

75. Barigelletti F, Juris A, Balzani V, Belser P, von Zelewsky A (1987) J Phys Chem<br />

91:1095<br />

76. Kitamura N, Sato M, Kim HB, Obata R, Tazuke S (1988) Inorg Chem 27:651<br />

77. Hiraga T, Kitamura N, Kim HB, Tazuke S, Mori N (1989) J Phys Chem 93:2940<br />

78. Hammarström L, Barigelletti F, Flamigni L, Indelli MT, Armaroli N, Calogero G,<br />

Guardigli M, Sour A, Collin JP, Sauvage JP (1997) J Phys Chem A 101:9061<br />

79.AbrahamsonM,WolpherH,JohanssonO,LarssonJ,KritikosM,ErikssonL,Norrby<br />

PO, Bergquist J, Sun L, Akermark B, Hammarström L (2005) Inorg Chem 44:3215<br />

80. Cherry WR, Henderson LJ Jr (1984) Inorg Chem 23:983<br />

81. Van Houten J, Watts RJ (1975) J Am Chem Soc 97:3843<br />

82. Englman R, Jortner J (1970) Mol Phys 18:145<br />

83. Gelbart WM, Freed KF, Rice SA (1970) J Chem Phys 52:2460<br />

84. Barbara P, Meyer TJ, Ratner MA (1996) J Phys Chem 100:13148<br />

85. Gleria M, Minto F, Beggiato G, Bortolus P (1978) J Chem Soc Chem Commun, p 285<br />

86. Porter GB, Sparks RH (1980) J Photochem 13:123


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 203<br />

87. Fetterolf ML, Offen HW (1985) J Phys Chem 89:3320<br />

88. Fetterolf ML, Offen HW (1986) J Phys Chem 90:1828<br />

89. De Cola L, Barigelletti F, Balzani V, Belser P, von Zelewsky A, Vögtle F, Ebmeyer F,<br />

Grammenudi S (1988) J Am Chem Soc 110:7210<br />

90. Barigelletti F, De Cola L, Balzani V, Belser P, von Zelewsky A, Vögtle F, Ebmeyer F<br />

(1989) J Am Chem Soc 111:4662<br />

91. Treadway JA, Strouse GF, Anderson PA, Keene RF, Meyer TJ (2002) J Chem Soc<br />

Dalton Trans, p 3820<br />

92. Medlycott EA, Hanan GS (2005) Chem Soc Rev 34:133<br />

93. Juris A, Campagna S, Balzani V, Gremaud G, von Zelewsky A (1988) Inorg Chem<br />

27:3652<br />

94. Sch<strong>of</strong>ield ER, Collin JP, Gruber N, Sauvage JP (2003) Chem Commun, p 188<br />

95. Bolletta F, Juris A, Maestri M, S<strong>and</strong>rini D (1980) Inorg Chim Acta 44:L175<br />

96. Laurence GS, Balzani V (1974) Inorg Chem 13:2976<br />

97. Gafney HD, Adamson AW (1972) J Am Chem Soc 94:8238<br />

98. Demas JN, Adamson AW (1973) J Am Chem Soc 95:5159<br />

99. Bock CR, Meyer TJ, Whitten DG (1974) J Am Chem Soc 96:4710<br />

100. Navon G, Sutin N (1974) Inorg Chem 13:2159<br />

101. H<strong>of</strong>fman MZ, Moggi L, Bolletta F, Hugh GL (1989) J Phys Chem Ref Data 18:219<br />

102. Natarajan P, Endicott JF (1973) J Phys Chem 77:971<br />

103. Natarajan P, Endicott JF (1973) J Phys Chem 77:1823<br />

104. Sabbatini N, Sc<strong>and</strong>ola MA, Carassiti V (1973) J Phys Chem 77:1307<br />

105. Sabbatini N, Sc<strong>and</strong>ola MA, Balzani V (1974) J Phys Chem 78:541<br />

106. Juris A, G<strong>and</strong>olfi MT, Manfrin MF, Balzani V (1976) J Am Chem Soc 98:1047<br />

107. Ballardini R, Varani G, Sc<strong>and</strong>ola F, Balzani V (1976) J Am Chem Soc 98:7432<br />

108. Juris A, Manfrin MF, Maestri M, Serpone N (1978) Inorg Chem 17:2258<br />

109. Fujita I, Kobayashi H (1970) J Chem Phys 52:4904<br />

110. Fujita I, Kobayashi H (1972) J Chem Phys 59:2902<br />

111. Fujita I, Kobayashi H (1972) Ber Bunsenges Phys Chem 76:115<br />

112. Sabbatini N, Balzani V (1972) J Am Chem Soc 94:7857<br />

113. Bolletta F, Maestri M, Moggi L, Balzani V (1975) J Chem Soc Chem Commun, p 901<br />

114. Creutz C, Sutin N (1976) J Am Chem Soc 98:6384<br />

115. Creutz C (1978) Inorg Chem 17:1046<br />

116. Sutin N, Creutz C (1978) Adv Chem Ser 168:1<br />

117. Brunschwig B, Sutin N (1978) J Am Chem Soc 100:7568<br />

118. Hoselton MA, Lin C-T, Schwarz HA, Sutin N (1978) J Am Chem Soc 100:2383<br />

119. Balzani V, Bergamini G, Campagna S, Puntoriero F (2007) <strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong><br />

<strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Overview <strong>and</strong> General Concepts (in this<br />

volume)<br />

120. Balzani V, Bolletta F, Ciano M, Maestri M (1983) J Chem Educ 60:447<br />

121. Balzani V, Bolletta F (1983) Comments Inorg Chem 2:211<br />

122. Gafney HD, Adamson AW (1975) J Chem Educ 52:480<br />

123. Bolletta F, Bonafede S (1986) Pure Appl Chem 58:1229<br />

124. Liu DK, Brunschwig BS, Creutz C, Sutin N (1986) J Am Chem Soc 108:1749<br />

125. Bolletta F, Ciano M, Balzani V, Serpone N (1982) Inorg Chim Acta 62:207<br />

126. White HS, Bard AJ (1982) J Am Chem Soc 104:6891<br />

127. Jonah CD, Matheson MS, Meisel D (1978) J Am Chem Soc 100:1449<br />

128. Hercules LM, Lytle FE (1966) J Am Chem Soc 88:4795<br />

129. Rubinstein I, Bard AJ (1981) J Am Chem Soc 103:512


204 S. Campagna et al.<br />

130. Balzani V, Bolletta F (1981) J Photochem 17:179<br />

131. Luttmer JD, Bard AJ (1978) J Electrochem Soc 125:1423<br />

132. Tokel-Takvoryan NE, Hemingway RE, Bard AJ (1973) J Am Chem Soc 95:6582<br />

133. Faulkner LR, Bard AJ (1977) In: Bard AD (ed) Electroanalytical chemistry, vol 10,<br />

chap 1. Dekker, New York<br />

134. Bard AJ, Laser D (1975) J Electrochem Soc 122:632<br />

135. Brilmyer GH, Bard AJ (1980) J Electrochem Soc 127:104<br />

136. Ege D, Becker WG, Bard AJ (1984) Anal Chem 56:2413<br />

137. Tung YL, Chen LS, Chi Y, Chou PT, Cheng YM, Li EY, Lee GH, Shu CF, Wu FI,<br />

Carty AJ (2006) Adv Funct Mater 16:1615<br />

138. Kalyuzhny G, Buda M, McNeill J, Barbara P, Bard AJ (2003) J Am Chem Soc 125:6272<br />

139. Slinker JD, Gorodetsky AA, Lowry MS, Wang S, Parker S, Rohl P, Bernhard S,<br />

Malliaras GG (2004) J Am Chem Soc 126:2763<br />

140. Yang J, Gordon KC (2004) Chem Phys Lett 385:481<br />

141. Chou PT, Chi Y (2006) Eur J Inorg Chem 3319<br />

142. Caspar JV, Westmorel<strong>and</strong> TD, Allen GH, Bradley PJ, Meyer TJ (1984) J Am Chem Soc<br />

106:3492<br />

143. McClanahan SF, Dallinger R, Holler FJ, Kincaid JR (1985) J Am Chem Soc 197:4860<br />

144. Mabrouk PA, Wrighton MS (1986) Inorg Chem 25:3004<br />

145. Worl LA, Duesing R, Chen P, Della Ciana L, Meyer TJ (1991) J Chem Soc Dalton<br />

Trans, p 849<br />

146. Kober EM, Caspar JV, Lumpkin RS, Meyer TJ (1986) J Phys Chem 90:3722<br />

147. Claude JP, Meyer TJ (1995) J Phys Chem 99:51<br />

148. Caspar JV, Meyer TJ (1983) Inorg Chem 22:2444<br />

149. Caspar JV, Meyer TJ (1983) J Am Chem Soc 105:5583<br />

150. Freed KF, Jortner J (1970) J Chem Phys 52:6272<br />

151. Englman R, Jortner J (1970) Mol Phys 18:145<br />

152. Bixon M, Jortner J (1968) J Chem Phys 48:715<br />

153. Siebr<strong>and</strong> W, Williams DF (1967) J Chem Phys 46:403<br />

154. Avouris P, Gelbart WM, El-Sayed MA (1977) Chem Rev 77:793<br />

155. Caspar JV, Kober EM, Sullivan BP, Meyer TJ (1982) J Am Chem Soc 104:630<br />

156. Y Wang, KS Schanze (1994) Inorg Chem 33:1354<br />

157. Damraurer NH, Boussie TR, Devenney M, McCusker JK (1997) J Am Chem Soc<br />

119:8253<br />

158. Murtaza Z, Zipp AP, Worl LA, Graff DK, Jones WE Jr, Bates WD, Meyer TJ (1991)<br />

J Am Chem Soc 113:5113<br />

159. Murtaza Z, Graff DK, Zipp AP, Worl LA, Jones WE Jr, Bates WD, Meyer TJ (1994)<br />

J Phys Chem 98:10504<br />

160. Liang YY, Baba AI, Kim WY, Atherton SJ, Schmehl RH (1996) J Phys Chem 100:<br />

18408<br />

161. Bergman SD, Gut D, Kol M, Sabatini C, Barbieri A, Barigelletti F (2005) Inorg Chem<br />

44:7943<br />

162. Seneviratne DS, Uddin MJ, Swayambunathan V, Schlegel HB, Endicott JF (2002)<br />

Inorg Chem 41:1502<br />

163. Xie P, Chen YJ, Endicott JF, Uddin MJ, Senerivatne D, McNamara PG (2003) Inorg<br />

Chem 42:5040<br />

164. Xie P, Chen YJ, Uddin MJ, Endicott JF (2005) J Phys Chem A 109:4671<br />

165. Endicott JF, Chen YJ, Xie P (2005) Coord Chem Rev 249:343<br />

166. Endicott JF, Chen YJ (2007) Coord Chem Rev 251:328


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 205<br />

167. Moser JE, Grätzel M (2001) In: Balzani V (ed) Electron transfer in chemistry, vol 5.<br />

Wiley-VCH, Weinheim, p 589<br />

168. Berglund Baudin H, Davidsson J, Serroni S, Juris A, Balzani V, Campagna S, Hammarström<br />

L (2002) J Phys Chem A 106:4312<br />

169. Andersson J, Puntoriero F, Serroni S, Yartsev A, Pascher T, Polivka T, Campagna S,<br />

Sundström V (2004) Chem Phys Lett 386:336<br />

170. Andersson J, Puntoriero F, Serroni S, Yartsev A, Pascher T, Polivka T, Campagna S,<br />

Sundström V (2004) Faraday Discuss 127:295<br />

171. Cannizzo A, Van Nourik F, Gawelda W, Bressler C, Chergui M (2006) Angew Chem<br />

Int Ed 45:3174<br />

172. Webb A, Knorr F, McHale J (2001) J Raman Spectrosc 32:481<br />

173. Wallin S, Davidsson J, Modin J, Hammarström L (2005) J Phys Chem A 109:4697<br />

174. Maestri M, Armaroli N, Balzani V, Constable EC, Thompson AMW (1995) Inorg<br />

Chem 34:2759<br />

175. Harriman A, Ziessel R (1996) Chem Commun, p 1707<br />

176. Benniston AC, Harriman A, Grosshenny V, Ziessel R (1997) New J Chem 21:405<br />

177. Mamo A, Stefio I, Poggi A, Tringali C, Di Pietro C, Campagna S (1997) New J Chem<br />

21:1173<br />

178. El-Ghayoury A, Harriman A, Khatyr A, Ziessel R (2000) J Phys Chem A 104:1512<br />

179. Polson MIJ, Medlycott EA, Mikelsons L, Taylor NJ, Watanabe M, Tanaka Y, Loiseau F,<br />

Passalacqua R, Campagna S (2004) Chem Eur J 10:3640<br />

180. Duatti M, Tasca S, Lynch FC, Bohlen H, Vos JG, Stragni S, Ward MD (2003) Inorg<br />

Chem 42:8377<br />

181. Medlycott EA, Hanan GS (2006) Coord Chem Rev 250:1763<br />

182. Benniston AC, Grosshenny V, Harriman A, Ziessel R (1994) Angew Chem Int Ed<br />

Engl 33:1884<br />

183. Hissler M, El-Ghayoury A, Harriman A, Ziessel R (1998) Angew Chem Int Ed 37:1717<br />

184. Barbieri A, Ventura B, Barigelletti F, De Nicola A, Quesada M, Ziessel R (2004)<br />

43:7359<br />

185. Benniston AC, Harriman A, Li P, Sams CA (2005) J Am Chem Soc 127:2553<br />

186. Galletta M, Puntoriero F, Campagna S, Chiorboli C, Quesada M, Goeb S, Ziessel S<br />

(2006) J Phys Chem A 110:4348<br />

187. Benniston AC, Harriman A (2006) Inorg Chim Acta 359:753<br />

188. Constable EC, Housecr<strong>of</strong>t CE, Cargill Thompson A, Passaniti P, Silvi S, Maestri M,<br />

Credi A (2007) Inorg Chim Acta 360:1102<br />

189. Abrahamsson M, Jäger M, Osterman T, Erikson L, Persson P, Becker HC, Johansson<br />

O, Hammarström L (2006) J Am Chem Soc 128:12616<br />

190. Dodsworth ES, Lever ABP (1985) Chem Phys Lett 119:61<br />

191. Dodsworth ES, Lever ABP (1986) Chem Phys Lett 124:152<br />

192. Rillema DP, Mack KB (1982) Inorg Chem 21:3849<br />

193. Juris A, Belser P, Barigelletti F, von Zelewsky A, Balzani V (1986) Inorg Chem 25:256<br />

194. Campagna S, Serroni S, Bodige S, MacDonnell FM (1999) Inorg Chem 38:692<br />

195. Warnmark K, Heyke O, Thomas JA, Lehn JM (1996) Chem Commun, p 2603<br />

196. Ishow E, Gourdon A, Launay JP, Chiorboli C, Sc<strong>and</strong>ola F (1999) Inorg Chem 38:1504<br />

197. Chiorboli C, Bignozzi CA, Sc<strong>and</strong>ola F, Ishow E, Gourdon A, Launay JP (1999) Inorg<br />

Chem 38:2402<br />

198. Lopez R, Leiva AM, Zuloaga F, Loeb B, Norambuena B, Omberg KM, Schoonover JR,<br />

Striplin D, Devenney M, Meyer TJ (1999) Inorg Chem 38:2924<br />

199. Torieda H, Yoshimura A, Nozaki K, Sakai S, Ohno T (2002) J Phys Chem A 106:11034


206 S. Campagna et al.<br />

200. Amouyal E, Homsi A, Chambron JC, Sauvage JP (1990) J Chem Soc Dalton Trans,<br />

p 1841<br />

201. Friedman AE, Chambron JC, Sauvage JP, Turro NJ, Barton JK (1990) J Am Chem Soc<br />

118:4960<br />

202. Tuite E, Lincoln P, Norden B (1997) J Am Chem Soc 119:239<br />

203. Nair RB, Cullum BM, Murphy CJ (1997) Inorg Chem 36:962<br />

204. Olson EJC, Hu D, Hormann A, Jonkman AM, Arkin MR, Stemp EDA, Barton JK, Barbara<br />

PF (1997) J Am Chem Soc 119:11458<br />

205. Albano G, Belser P, De Cola L, G<strong>and</strong>olfi MT (1999) Chem Commun, p 1171<br />

206. Flamigni L, Encinas S, MacDonnell FM, Kim KJ, Puntoriero F, Campagna S (2000)<br />

Chem Commun, p 1185<br />

207. Torieda H, Nozali K, Yoshimura A, Ohno T (2004) J Phys Chem A 108:4819<br />

208. Balzani V, Moggi L, Sc<strong>and</strong>ola F (1987) In: Balzani V (ed) Supramolecular photochemistry.<br />

Reidel, Dordrecht, p 1<br />

209. Balzani V, Sc<strong>and</strong>ola F (1991) Supramolecular photochemistry. Horwood, Chichester<br />

210. Chiorboli C, Indelli MT, Sc<strong>and</strong>ola F (2005) Top Curr Chem 257:63<br />

211. Schlike B, Belser P, De Cola L, Sabbioni E, Balzani V (1999) J Am Chem Soc 121:4207<br />

212. Welter S, Salluce N, Belser P, Groeneveld M, De Cola L (2005) Coord Chem Rev<br />

249:1360<br />

213. Indelli MT, Chiorboli C, Flamigni L, De Cola L, Sc<strong>and</strong>ola F (2007) Inorg Chem<br />

(in press)<br />

214. Cavazzini M, Pastorelli P, Quici S, Loiseau F, Campagna S (2005) Chem Commun,<br />

p 5266<br />

215. Barigelletti F, Flamigni L, Collin JP, Sauvage JP (1997) Chem Commun, p 333<br />

216. Collin JP, Gavina P, Hietz V, Sauvage JP (1998) Eur J Inorg Chem 1<br />

217. Barigelletti F, Flamigni L (2000) Chem Soc Rev 29:1<br />

218. Barigelletti F, Flamigni L, Guardigli M, Juris A, Beley M, Chodorowski-Kimmes S,<br />

Collin JP, Sauvage JP (1996) Inorg Chem 35:136<br />

219. Polson M, Chiorboli C, Fracasso S, Sc<strong>and</strong>ola F (2007) Photochem Photobiol Sci 6:438<br />

220. Indelli MT, Sc<strong>and</strong>ola F, Collin JP, Sauvage JP, Sour A (1996) Inorg Chem 25:303<br />

221. Indelli MT, Sc<strong>and</strong>ola F, Flamigni L, Collin JP, Sauvage JP, Sour A (1997) Inorg Chem<br />

26:4247<br />

222. Helms A, Heiler D, McLendon G (1991) J Am Chem Soc 113:4325<br />

223. Onuchic JN, Beratan DN (1987) J Am Chem Soc 109:6771<br />

224. Lainé PP, Bedioui F, Loiseau F, Chiorboli C, Campagna S (2006) J Am Chem Soc<br />

128:7510<br />

225. Lainé PP, Loiseau F, Campagna S, Ci<strong>of</strong>ini I, Adamo C (2006) Inorg Chem 45:5538<br />

226. Grosshenny V, Harriman A, Ziessel R (1995) Angew Chem Int Ed Engl 34:1100<br />

227. Paddon-Row MN (2001) In: Balzani V (ed) Electron transfer in chemistry, vol 3.<br />

Wiley-VCH, Weinheim, p 179<br />

228. Sc<strong>and</strong>ola F, Chiorboli C, Indelli MT, Rampi MA (2001) In: Balzani V (ed) Electron<br />

transfer in chemistry, vol. 3. Wiley-VCH, Weinheim, p 337<br />

229. Grosshenny V, Harriman A, Ziessel R (1995) Angew Chem Int Ed Engl 34:2705<br />

230. Ziessel R, Hissler M, El-ghayoury A, Harriman A (1998) Coord Chem Rev 177:1251<br />

231. Belser P, Dux R, Baak M, De Cola L, Balzani V (1995) Angew Chem Int Ed Engl<br />

34:595<br />

232. De Cola L, Balzani V, Dux R, Baak M, Belser P (1995) Supramol Chem 5:297<br />

233. Chiorboli C, Rodgers MAJ, Sc<strong>and</strong>ola F (2003) J Am Chem Soc 125:483<br />

234. Welter S, Laforet F, Cecchetto E, Vergeer F, De Cola L (2005) ChemPhysChem 6:2417


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 207<br />

235. Torieda H, Yoshimura A, Nozaki K, Sakai S, Ohno T (2002) J Phys Chem A 106:11034<br />

236. Torieda H, Yoshimura A, Nozaki K, Ohno T (2004) J Phys Chem A 108:2148<br />

237. Ruben M, Rojo J, Romero-Salguero FJ, Uppadine LH, Lehn JM (2004) Angew Chem<br />

Int Ed 43:3644<br />

238. Credi A, Balzani V, Campagna S, Hanan GS, Arana CR, Lehn JM (1995) Chem Phys<br />

Lett 243:105<br />

239. Bassani DM, Lehn JM, Serroni S, Puntoriero F, Campagna S (2003) Chem Eur J<br />

9:5936<br />

240. Stadler AM, Puntoriero F, Campagna S, Kyritsakas N, Welter R, Lehn JM (2005)<br />

Chem Eur J 11:3997<br />

241. Juris A (2003) Annu Rep Prog Chem Sect C 99:177<br />

242. Balzani V, Ceroni P, Maestri M, Saudan C, Vicinelli V (2003) Top Curr Chem 228:159<br />

243. Balzani V, Vögtle F (2003) CR Chim 6:867<br />

244. Balzani V, Juris A, Puntoriero F, Campagna S (2007) In: Gleria M, De Jaeger R (eds)<br />

Inorganic polymers, chap 19. Nova Science, New York<br />

245. Balzani V, Campagna S, Denti G, Serroni S, Juris A, Venturi M (1998) Acc Chem Res<br />

31:26<br />

246. Pleovets M, Vögtle F, De Cola L, Balzani V (1999) New J Chem 23:63<br />

247. Vögtle F, Pleovets M, Nieger M, Azzellini GC, Credi A, De Cola L, De Marchis V, Venturi<br />

M, Balzani V (1999) J Am Chem Soc 121:6290<br />

248. McClenaghan ND, Passalacqua R, Loiseau F, Campagna S, Verheyde B, Hameurlaine<br />

A, Dehaen W (2003) J Am Chem Soc 125:5356<br />

249. Zhou XL, Tyson DS, Castellano FN (2000) Angew Chem Int Ed 39:4301<br />

250. Tyson DS, Luman CR, Castellano FN (2002) Inorg Chem 41:3578<br />

251. Serroni S, Denti G, Campagna S, Juris A, Ciano M, Balzani V (1992) Angew Chem<br />

Int Ed Engl 31:1493<br />

252. Campagna S, Denti G, Serroni S, Juris A, Venturi M, Ricevuto V, Balzani V (1995)<br />

Chem Eur J 1:211<br />

253. Serroni S, Juris A, Venturi M, Campagna S, Resino Resino I, Denti G, Credi A,<br />

Balzani V (1997) J Mater Chem 7:1227<br />

254. Denti G, Campagna S, Serroni S, Ciano M, Balzani V (1992) J Am Chem Soc 114:2944<br />

255. Campagna S, Denti G, Sabatino L, Serroni S, Ciano M, Balzani V (1990) Inorg Chem<br />

29:4750<br />

256. Serroni S, Campagna S, Denti G, Keyes TE, Vos JG (1996) Inorg Chem 35:4513<br />

257. Campagna S, Denti G, Serroni S, Ciano M, Balzani V (1991) Inorg Chem 30:3728<br />

258. Denti G, Serroni S, Campagna S, Ricevuto V, Juris A, Ciano M, Balzani V (1992)<br />

Inorg Chim Acta 198–200:507<br />

259. Denti G, Campagna S, Sabatino L, Serroni S, Ciano M, Balzani V (1990) Inorg Chim<br />

Acta 176:175<br />

260. Campagna S, Denti G, Serroni S, Juris A, Ciano M, Balzani V (1991) Inorg Chem<br />

31:2982<br />

261. Leveque J, Moucheron C, Kirsch-De Mesmaeker A, Loiseau F, Serroni S, Puntoriero F,<br />

Campagna S, Nierengarten H, Van Dorsselaer A (2004) Chem Commun, p 877<br />

262. Campagna S, Denti G, Sabatino L, Serroni S, Ciano M, Balzani V (1989) J Chem Soc<br />

Chem Commun, p 1500<br />

263. Sommovigo M, Denti G, Serroni S, Campagna S, Mingazzini C, Mariotti C, Juris A<br />

(2001) Inorg Chem 40:3318<br />

264. Puntoriero F, Serroni S, Galletta M, Juris A, Licciardello A, Chiorboli C, Campagna S,<br />

Sc<strong>and</strong>ola F (2005) ChemPhysChem 6:129


208 S. Campagna et al.<br />

265. Tachibana Y, Moser JE, Grätzel M, Klug DR, Durrant JR (1996) J Phys Chem<br />

100:20056<br />

266. Benkö G, Kallioinen J, Korppi-Tommola JEI, Yartsev AP, Sundström V (2002) J Am<br />

Chem Soc 124:489<br />

267. Kallioinen J, Benkö G, Sundström V, Korppi-Tommola JEI, Yartsev AP (2002) J Phys<br />

Chem B 106:4396<br />

268. Tachibana Y, Nazeeruddin MDK, Grätzel M, Klug DR, Durrant JR (2002) Chem Phys<br />

285:127<br />

269. Asbury JB, Anderson NA, Hao EC, Ai X, Lian TQ (2003) J Phys Chem 107:7376<br />

270. Benkö G, Kallioinen J, Myllyperkiö P, Trif F, Korppi-Tommola JEI, Yartsev AP, Sundström<br />

V (2004) J Phys Chem B 108:2862<br />

271. Wenger B, Grätzel M, Moser JE (2005) J Am Chem Soc 127:12150<br />

272. McClenaghan ND, Loiseau F, Puntoriero F, Serroni S, Campagna S (2001) Chem<br />

Commun, p 2634<br />

273. Larsen J, Puntoriero F, Pascher T, McClenaghan ND, Campagna S, Sundström V,<br />

˚Akesson E (2007) Chem Phys Chem (in press)<br />

274. Leveque J, Elias B, Moucheron C, Kirsch-De Mesmaeker A (2005) Inorg Chem 44:393<br />

275. Börje A, Köthe O, Juris A (2002) J Chem Soc Dalton Trans, p 843<br />

276. Constable EC, H<strong>and</strong>el RW, Housecr<strong>of</strong>t CE, Farràn Morales A, Ventura B, Flamigni L,<br />

Barigelletti F (2005) Chem Eur J 11:4024<br />

277. Arm KJ, Williams JAG (2005) Chem Commun, p 230<br />

278. Campagna S, Serroni S, Puntoriero F, Loiseau F, De Cola L, Kleverlaan CJ, Becher J,<br />

Sørensen AP, Hascoat P, Thorup N (2002) Chem Eur J 8:4461<br />

279. Puntoriero F, Nastasi F, Cavazzini M, Quici S, Campagna S (2007) Coord Chem Rev<br />

251:536<br />

280. Balzani V, Credi A, Venturi M (2003) Molecular devices <strong>and</strong> machines. Wiley-VCH,<br />

Weinheim<br />

281. Schanze KS, Walters KA (1998) In: Ramamurthy V, Schanze KS (eds) Organic <strong>and</strong><br />

inorganic photochemistry. Dekker, New York, p 75<br />

282. Collin JP, Guillerez S, Sauvage JP, Barigelletti F, De Cola L, Flamigni L, Balzani V<br />

(1991) Inorg Chem 30:4230<br />

283. Danielson E, Elliott CM, Merkert JW, Meyer TJ (1987) J Am Chem Soc 109:2519<br />

284. Cooley LF, Larson SL, Elliott CM, Kelley DF (1991) J Phys Chem 95:10694<br />

285. Larson SL, Elliott CM, Kelley DF (1995) J Phys Chem 99:6530<br />

286. Klumpp T, Linsenmann M, Larson SL, Limoges BR, Buerssner D, Krissinel B, Elliott<br />

CM, Steiner UE (1999) J Am Chem Soc 121:4092<br />

287. Klumpp T, Linsenmann M, Larson SL, Limoges BR, Buerssner D, Krissinel B, Elliott<br />

CM, Steiner UE (1999) J Am Chem Soc 121:1076<br />

288. Weber JM, Rawls MT, MacKenzie VJ, Limoges BR, Elliott CM (2007) J Am Chem Soc<br />

129:313<br />

289. Larson SL, Elliott CM, Kelley DF (1996) Inorg Chem 35:2070<br />

290. Schmehl RH, Ryu CK, Elliott CM, Headford CLE, Ferrere S (1990) Adv Chem Ser<br />

226:211<br />

291. Cooley LF, Headford CEL, Elliott CM, Kelley DF (1988) J Am Chem Soc 112:6673<br />

292. Ryu CK, Wang R, Schmehl RH, Ferrere S, Ludwikow M, Merkert JW, Headford CEL,<br />

Elliott CM (1992) J Am Chem Soc 114:430<br />

293. Yonemoto EH, Saupe GB, Schmehl RH, Hubig SM, Riley RL, Iverson BL, Mallouk TE<br />

(1994) J Am Chem Soc 116:4786<br />

294. Meckenburg SL, Peek BM, Schoonover JR, McCafferty DJ, Wall CJ, Erickson BW,<br />

Meyer TJ (1993) J Am Chem Soc 115:5479


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 209<br />

295. Meckenburg SL, McCafferty DJ, Schoonover JR, Peek BM, Erickson BW, Meyer TJ<br />

(1994) Inorg Chem 33:2974<br />

296. Hu YZ, Tsukiji S, Shinkai S, Oishi S, Hamachi I (2000) J Am Chem Soc 122:241<br />

297. Huynh MHV, Dattelbaum DM, Meyer TJ (2005) Coord Chem Rev 249:457<br />

298. McCafferty DG, Bishop BM, Wall CG, Hughes SG, Meckenburg SL, Meyer TJ, Erickson<br />

BW (1995) Tetrahedron 51:1093<br />

299. McCafferty DG, Friesen DA, Danielson E, Wall CG, Saederholm MJ, Erickson BW,<br />

Meyer TJ (1993) Proc Natl Acad Sci USA 93:8200<br />

300. Striplin DR, Reece SY, McCafferty DG, Wall CG, Friesen DA, Erickson BW, Meyer TJ<br />

(2004) J Am Chem Soc 126:5282<br />

301. Slate CA, Striplin DR, Moss JA, Chen PY, Erickson BW, Meyer TJ (1998) J Am Chem<br />

Soc 120:4885<br />

302. Serron SA, Aldridge SA III, Fleming CN, Danell RM, Baik MH, Sykora M, Dattelbaum<br />

DM, Meyer TJ (2004) J Am Chem Soc 126:14506<br />

303. Serron SA, Aldridge SA III, Danell RM, Meyer TJ (2000) Tetrahedron 41:4039<br />

304. Strouse GF, Worl LA, Younathan JN, Meyer TJ (1989) J Am Chem Soc 111:9101<br />

305. Jones WE Jr, Baxter SM, Strouse GF, Meyer TJ (1993) J Am Chem Soc 115:7363<br />

306. Worl LA, Jones WE Jr, Strouse GF, Younathan JN, Danielson E, Maxwell KA, Sykora<br />

M, Meyer TJ (1999) Inorg Chem 28:2705<br />

307. Dupray LM, Devenny M, Striplin DR, Meyer TJ (1997) J Am Chem Soc 119:10243<br />

308. Fleming CN, Dupray LM, Papanikolas JM, Meyer TJ (2002) J Phys Chem 106:2328<br />

309. Worl LA, Strouse GF, Younathan JN, Baxter SM, Meyer TJ (1990) J Am Chem Soc<br />

112:7571<br />

310. Baxter SM, Jones WE Jr, Danielson E, Worl LA, Strouse GF, Younathan JN, Meyer TJ<br />

(1991) Coord Chem Rev 111:47<br />

311. Sykora M, Maxwell KA, DeSimone JM, Meyer TJ (2000) Proc Natl Acad Sci USA<br />

97:7687<br />

312. Konduri R, Ye H, MacDonnell FM, Serroni S, Campagna S, Rajeshwar K (2002)<br />

Angew Chem Int Ed 41:3185<br />

313. Chiorboli C, Fracasso S, Sc<strong>and</strong>ola F, Campagna S, Serroni S, Konduri R, MacDonnell<br />

FM (2003) Chem Commun, p 1658<br />

314. Chiorboli C, Fracasso S, Ravaglia M, Sc<strong>and</strong>ola F, Campagna S, Wouters K, Konduri R,<br />

MacDonnell FM (2005) Inorg Chem 44:1658<br />

315. Konduri R, de Tacconi NR, Rajeshwar K, MacDonnell FM (2004) J Am Chem Soc<br />

126:11621<br />

316. de Tacconi NR, Lezna RO, Konduri R, Ongeri F, Rajeshwar K, MacDonnell FM (2005)<br />

Chem Eur J 11:4327<br />

317. Wouters KL, de Tacconi NR, Konduri R, Lezna RO, MacDonnell FM (2006) Photosynth<br />

Res 87:1<br />

318. Molnar SM, Nallas G, Bridgewater JS, Brewer KJ (1994) J Am Chem Soc 116:5206<br />

319. Holder AA, Swavey S, Brewer KJ (2004) Inorg Chem 43:303<br />

320. Elvington M, Brewer KJ (2006) Inorg Chem 45:5242<br />

321. Barber J (2003) Q Rev Biophys 36:71<br />

322. Ferreira KN, Iverson TM, Maghlaoui K, Barber J, Iwata S (2004) Science 303:1831<br />

323. Iwata S, Barber J (2004) Curr Opin Struct Biol 14:447<br />

324. McEvoy JP, Brudvig G (2006) Chem Rev 106:4455<br />

325. Hammarström L (2003) Curr Opin Chem Biol 7:666<br />

326. Berg KE, Tran A, Raymond MK, Abrahamsson M, Wolny J, Redon S, Andersson M,<br />

Sun L, Styring S, Hammarström L, T<strong>of</strong>lund H, Akermark B (2001) Eur J Inorg Chem<br />

1019


210 S. Campagna et al.<br />

327. Abrahamsson M, Baudin HB, Tran A, Philouze C, Berg KE, Raymond MK, Sun L, Akermark<br />

B, Styring S, Hammarström L (2002) Inorg Chem 41:1534<br />

328. Magnusson A, Frapart Y, Abrahamsson M, Horner O, Akermark B, Sun L, Girerd JJ,<br />

Hammarström L, Styring S (1999) J Am Chem Soc 121:89<br />

329. Burdinski D, Wieghardt K, Steenken S (1999) J Am Chem Soc 121:10781<br />

330. Sjodin M, Styring S, Sun L, Akermark B, Hammarström L (2000) J Am Chem Soc<br />

122:3932<br />

331. Sun L, Raymond MK, Magnusson A, LeGourriérec D, Tamm M, Abrahamsson M,<br />

Huang-Kenez P, Martensson J, Stenhagen G, Hammarström L, Akermark B, Styring S<br />

(2000) J Inorg Biochem 78:15<br />

332. Burdinski D, Eberhard B, Wieghardt K (2000) Inorg Chem 39:105<br />

333. Huang P, Magnusson A, Lomoth R, Abrahamsson M, Tamm M, Sun L, van Rotterdam<br />

B, Park J, Hammarström L, Akermark B, Styring S (2002) J Inorg Biochem<br />

91:150<br />

334. Borgström M, Shaikh N, Johansson O, Anderlund MF, Styring S, Akermark B, Magnusson<br />

A, Hammarström L (2005) J Am Chem Soc 127:17504<br />

335. Balzani V, Moggi L, Manfrin MF, Bolletta F, Gleria M (1975) Science 189:852<br />

336. Lehn JM, Sauvage JP (1977) Nouv J Chim 1:449<br />

337. Kalyanasundaram K, Grätzel M (1979) Angew Chem Int Ed Engl 18:701<br />

338. Creutz C, Sutin N (1975) Proc Natl Acad Sci USA 72:2858<br />

339. Kalyanasundaram K, Kiwi J, Grätzel M (1978) Helv Chim Acta 61:2720<br />

340. Meyer TJ (1989) Acc Chem Res 22:163<br />

341. Campagna S, Serroni S, Puntoriero F, Di Pietro C, Ricevuto V (2001) In: Balzani V<br />

(ed) Electron transfer in chemistry, vol. 5. Wiley-VCH, Weinheim, p 168<br />

342. Alstrum-Acevedo JH, Brennaman MK, Meyer TJ (2005) Inorg Chem 44:6802<br />

343. Armaroli N, Balzani V (2007) Angew Chem Int Ed 46:52<br />

344. Wasielewski MR (1992) Chem Rev 92:435<br />

345. Gust D, Moore TA, Moore AL (1993) Acc Chem Res 26:198<br />

346. Ozawa H, Haga MA, Sakai K (2006) J Am Chem Soc 128:4926<br />

347. Rau S, Schäfer B, Gleick D, Anders E, Rudolph M, Friedrich M, Görls H, Henry W,<br />

Vos JG (2006) Angew Chem Int Ed 45:6215<br />

348. Brewer KJ, Elvington M (2006) US Patent 20060120954A1<br />

349. Gholankhass B, Mametsuka H, Koite K, Tanabe T, Furue M, Ishitani O (2005) Inorg<br />

Chem 44:2326<br />

350. Hawecker J, Lehn JM, Ziessel R (1983) J Chem Soc Chem Commun, p 536<br />

351. Walther D, Ruben M, Rau S (1999) Coord Chem Rev 182:67<br />

352. Inagaki A, Edure S, Yatsuda S, Akita M (2005) Chem Commun, p 5468<br />

353. Osawa M, Hoshino M, Wakatsuki Y (2001) Angew Chem Int Ed 40:3472<br />

354. Rau S, Walther D, Vos JG (2007) Dalton Trans, p 915<br />

355. Ballardini R, Balzani V, Credi A, G<strong>and</strong>olfi MT, Venturi M (2001) Acc Chem Res<br />

34:445<br />

356. Balzani V (2003) Photochem Photobiol Sci 2:459<br />

357. Venturi M, Balzani V, Ballardini R, Credi A, G<strong>and</strong>olfi MT (2004) Int J Photoenerg 6:1<br />

358. Balzani V, Credi A, Venturi M (2007) Nano Today 2:18<br />

359. Balzani V, Clemente-Leon M, Credi A, Ferrer B, Venturi M, Fllod A, Stoddart JF<br />

(2006) Proc Natl Acad Sci USA 103:1178<br />

360. Mobian P, Kern J-P, Sauvage J-P (2004) Angew Chem Int Ed 43:2392<br />

361. Collin J-P, Jouvenot D, Koizumi M, Sauvage J-P (2005) Eur J Inorg Chem 1850<br />

362. Kirsch-De Mesmaeker A, Lecomte JP, Kelly JM (1996) Top Curr Chem 177:25


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 211<br />

363. Jenkins Y, Friedman AE, Turro NJ, Barton JK (1992) Biochemistry 31:10809<br />

364. Friedman AE, Chambron JC, Sauvage JP, Turro NJ, Barton JK (1984) Nucleic Acids<br />

Res 13:6016<br />

365. Tossi AB, Kelly JM (1989) Photochem Photobiol 49:545<br />

366. Ortmans I, Moucheron C, Kirsch-De Mesmaeker A (1998) Coord Chem Rev 168:233<br />

367. Kelly JM, Feeney M, Jacquet L, Kirsch-De Mesmeaker A, Lecomte JP (1997) Pure<br />

Appl Chem 69:767<br />

368. Piérard F, Del Guerzo A, Kirsch-De Mesmaeker A, Demeunynck M, Lhomme J (2001)<br />

Phys Chem Chem Phys 3:2911<br />

369. Vos JG, Kelly JM (2006) Dalton Trans, p 4839<br />

370. Hiort C, Lincoln P, Nordén B (1993) J Am Chem Soc 115:3448<br />

371. Lincoln P, Broo A, Nordén B (1996) J Am Chem Soc 118:2644<br />

372. Ujj L, Coates CG, Kelly JM, Kruger P, McGarvey JJ, Atkinson GH (2002) J Phys<br />

Chem B 106:4854<br />

373. Coates CG, Callaghan PL, McGarvey JJ, Kelly JM, Kruger P, Higgins ME (2000) J Raman<br />

Spectrosc 31:283<br />

374. Turro C, Bossman SH, Jenkins Y, Barton JK, Turro NJ (1995) J Am Chem Soc<br />

117:9026<br />

375. Sabatini E, Nikol HD, Gray HB, Anson FC (1996) J Am Chem Soc 118:1158<br />

376. Nair RB, Cullum BM, Murphy CJ (1997) Inorg Chem 36:962<br />

377. Guo XQ, Castellano FN, Li L, Lakowitz JR (1998) Biophys Chem 71:51<br />

378. Brennaman MK, Alstrum-Acevedo JH, Fleming CN, Jang P, Meyer TJ, Papanikolas JM<br />

(2002) J Am Chem Soc 124:15094<br />

379. Pourtois G, Beljonne D, Moucheron C, Schumm S, Kirsch-De Mesmaeker A, Lazzaroni<br />

R, Brédas JL (2004) J Am Chem Soc 126:683<br />

380. Onfelt B, Lincoln P, Nordéen B, Baskin JS, Zewail AH (2000) Proc Natl Acad Sci USA<br />

97:5708<br />

381. Coates CG, Ol<strong>of</strong>sson J, Coletti M, McGarvey JJ, Onfelt B, Lincoln P, Nordén B, Tuite E,<br />

Matousek P, Parker AW (2001) J Phys Chem B 105:12653<br />

382. Coates CG, Callaghan P, McGarvey JJ, Kelly JM, Jacquet L, Kirsch-De Mesmaeker A<br />

(2002) J Mol Struct 598:15<br />

383. O’Donoghue K, Kelly JM, Kruger PE (2004) Dalton Trans, p 13<br />

384. Moucheron C, Kirsch-De Mesmaeker A, Kelly JM (1997) J Photochem Photobiol B<br />

40:91<br />

385. Ortmans I, Elias B, Kelly JM, Moucheron C, Kirsch-De Mesmaeker A (2004) Dalton<br />

Trans, p 668<br />

386. Jacquet L, Kelly JM, Kirsch-De Mesmaeker A (1995) J Chem Soc Chem Commun,<br />

p 913<br />

387. Lentzen O, Defrancq E, Constant JF, Schumm S, Garcia-Fresnadillo D, Moucheron C,<br />

Dumy P, Kirsch-De Mesmaeker A (2004) J Biol Inorg Chem 9:100<br />

388. Gicquel E, Boisdenghien A, Defranc E, Moucheron C, Kirsch-De Mesmaeker A<br />

(2006) Chem Commun, p 2764<br />

389. Jenkins Y, Barton JK (1992) J Am Chem Soc 114:8736<br />

390. Hurley DJ, Tor Y (1998) J Am Chem Soc 120:2194<br />

391. Telser J, Cruickshank A, Schanze KS, Netzel TL (1989) J Am Chem Soc 111:7221<br />

392. Grimm GN, Boutorine AS, Lincoln P, Nordén B, Hélène C (2002) ChemBioChem<br />

3:324<br />

393. Lentzen O, Constant JF, Defrancq E, Prevost M, Schumm S, Moucheron C, Dumy P,<br />

Kirsch-De Mesmaeker A (2003) ChemBioChem 4:195


212 S. Campagna et al.<br />

394. Crean CW, Kavanagh YT, O’Keefe CM, Lawler MP, Stevenson C, Davies JH, Boyle PH,<br />

Kelly JM (2002) Photochem Photobiol Sci 1:1024<br />

395. Vargas-Baca I, Mitra D, Zulyniak HJ, Banerjee J, Sleiman HF (2001) Angew Chem Int<br />

Ed 40:4629<br />

396. Mitra D, Di Cesare N, Sleiman HF (2004) Angew Chem Int Ed 43:5804<br />

397. Chen B, Sleiman HF (2004) Macromolecules 37:5866<br />

398. Pauly M, Kayser I, Schmitz M, Dicato M, Del Guerzo A, Kolber I, Moucheron C,<br />

Kirsch-De Mesmaeker A (2002) Chem Commun, p 1086<br />

399. Ortmans I, Content S, Boutonnet N, Kirsch-De Mesmaeker A, Bannwarth W, Constant<br />

JF, Defrancq E, Lhomme J (1999) Chem Eur J 5:2712<br />

400. Garcia-Fresnadillo D, Boutonnet N, Schumm S, Moucheron C, Kirsch-De Mesmaeker<br />

A, Defrancq E, Constant JF, Lhomme J (2002) Biophys J 82:978<br />

401. O’Regan B, Grätzel M (1991) Nature 335:737<br />

402. Hagfeldt A, Grätzel M (1992) Chem Rev 95:49<br />

403. Grätzel M (2005) Inorg Chem 44:6841<br />

404. Meyer G (2005) Inorg Chem 44:6852<br />

405. Bignozzi CA, Argazzi R, Caramori S (2007) Inorganic <strong>and</strong> bioinorganic chemistry.<br />

In: Bertini I (ed) Encyclopaedia <strong>of</strong> life supporting systems (EOLSS), developed under<br />

the auspices <strong>of</strong> UNESCO. Eolss, Oxford (in press) http://www.eolss.net<br />

406. Gerischer H, Tributsch H (1968) Ber Bunsenges Phys Chem 72:437<br />

407. DeSilvestro J, Grätzel M, Kavan L, Moser JE, Augustynski J (1985) J Am Chem Soc<br />

107:2988<br />

408. Memming R (1984) Prog Surf Sci 17:7<br />

409. Nazeeruddin MK, Pechy P, Renouard T, Zakeeruddin SM, Humphry-Baker R,<br />

Comte P, Liska P, Cevey L, Costa E, Shklover V, Spiccia L, Deacon GB, Bignozzi CA,<br />

Grätzel M (2001) J Am Chem Soc 123:1613<br />

410. Nazeeruddin MK, Wang Q, Cevey L, Aranyos V, Liska P, Figgemeier E, Klein C, Hirata<br />

N, Koops S, Haque SA, Durrant JR, Hagfeldt A, Lever ABP, Grätzel M (2006)<br />

Inorg Chem 45:787<br />

411. Grätzel M (2001) Pure Appl Chem 73:459<br />

412. Cao F, Oskam G, Searson PC, Stipkala J, Farzhad F, Heimer TA, Meyer GJ (1995)<br />

J Phys Chem 99:11974<br />

413. Redmond G, Grätzel M, Fitzmaurice D (1993) J Phys Chem 97:6951<br />

414. Boschlo G, Fitzmaurice D (1999) J Phys Chem B 103:7860<br />

415. Nazeeruddin MK, Kay A, Rodicio R, Humphry-Baker R, Muller E, Liska P, Vlachopoulos<br />

M, Grätzel M (1993) J Am Chem Soc 115:6382<br />

416. Argazzi R, Bignozzi CA, Hasselmann GM, Meyer GJ (1998) Inorg Chem 37:4533<br />

417. Rensmo H, Sodergen S, Patthey L, Westermak L, Vayssieres L, Kohle O, Bruhwiler PA,<br />

Hagfeldt A, Siegbahn H (1997) Chem Phys Lett 274:51<br />

418. Nazeeruddin MK, Pechy P, Grätzel M (1997) Chem Commun, p 1705<br />

419. Zakeeruddin SM, Nazeeruddin MK, Pechy P, Rotzinger F, Humphry-Baker R, Kalyanasundaram<br />

K, Grätzel M, Shklover V, Haibach T (1997) Inorg Chem 36:5937<br />

420. Argazzi R, Bignozzi CA, Heimer TA, Meyer GJ (1997) Inorg Chem 36:2<br />

421. Indelli MT, Bignozzi CA, Harriman A, Schoonover JR, Sc<strong>and</strong>ola F (1994) J Am Chem<br />

Soc 116:3768<br />

422. Argazzi R, Bignozzi CA, Heimer TA, Castellano FN, Meyer GJ (1997) J Phys Chem B<br />

101:2591<br />

423. Hirata N, Lagref JJ, Palomares EJ, Durrant JR, Nazeeruddin MK, Grätzel M, Di Censo<br />

D (2004) Chem Eur J 10:595


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium 213<br />

424. de Silva AP, Tecilla P (eds) (2005) Thematic issue on fluorescent sensors. J Mater<br />

Chem 15:2617–2976<br />

425. Schmittel M, Lin HW (2007) Angew Chem Int Ed 46:893<br />

426. Demas JN, DeGraff BA (1991) Anal Chem 63:829A<br />

427. Demas JN, DeGraff BA, Coleman P (1999) Anal Chem 71:793A<br />

428. Demas JN, DeGraff BA (2001) Coord Chem Rev 211:317<br />

429. Borisov SM, Vasylevska AS, Krause C, Wolfbeis O (2006) Adv Funct Mater 16:1536<br />

430. Welter S, Brunner K, H<strong>of</strong>straat JW, De Cola L (2003) Nature 421:54<br />

431. Ford WE, Rodgers MAJ (1992) J Phys Chem 96:2917<br />

432. Wilson GJ, Launikonis A, Sasse WHF, Mau AWH (1997) J Phys Chem A 101:4860<br />

433. Simon JA, Curry SL, Schmehl RH, Schartz TR, Piotrowiak P, Jin X, Thummel RP<br />

(1997) J Am Chem Soc 119:11012<br />

434. Harriman A, Hissler M, Khatyr A, Ziessel R (1999) Chem Eur J 5:3366<br />

435. Tyson DS, Lauman CR, Zhou X, Castellano FN (2001) Inorg Chem 40:4063<br />

436. McClenaghan ND, Barigelletti F, Maubert B, Campagna S (2002) Chem Commun,<br />

p 602<br />

437. Passalacqua R, Loiseau F, Campagna S, Fang YQ, Hanan GS (2003) Angew Chem Int<br />

Ed 42:1608<br />

438. Leroy-Lhez S, Belin C, D’Aleo A, Williams R, De Cola L, Fages F (2003) Supramol<br />

Chem 15:627<br />

439. Maubert B, McClenhagan ND, Indelli MT, Campagna S (2003) J Phys Chem A<br />

107:447<br />

440. Wang XY, Del Guerzo A, Schmehl RH (2004) J Photochem Photobiol C 5:55<br />

441. McClenaghan ND, Leydet Y, Maubert B, Indelli MT, Campagna S (2005) Coord Chem<br />

Rev 249:1336<br />

442. Wang J, Fang YQ, Bourget-Merle L, Polson MIJ, Hanan GS, Juris A, Loiseau F, Campagna<br />

S (2006) Chem Eur J 12:8539<br />

443. Strouse GF, Schoonover JR, Duesing R, Boyde S, Jones WE, Meyer TJ (1995) Inorg<br />

Chem 34:473<br />

444. Fang YQ, Taylor NJ, Hanan GS, Loiseau F, Passalacqua R, Campagna S, Nierengerten<br />

H, Van Dorsselaer A (2002) J Am Chem Soc 124:7912<br />

445. Polson MIJ, Loiseau F, Campagna S, Hanan GS (2006) Chem Commun, p 1301<br />

446. Fang YQ, Taylor NJ, Laverdière F, Hanan GS, Loiseau F, Nastasi F, Campagna S, Nierengerten<br />

H, Leize-Wagner E, Van Dorsselaer A (2007) Inorg Chem 46:2854<br />

447. Encinas S, Simpson NRM, Andrews P, Ward MD, White CW, Armaroli N, Barigelletti<br />

F, Houlton A (2000) New J Chem 24:987<br />

448. Ward MD, White CM, Barigelletti F, Armaroli N, Calogero G, Flamigni L (1998)<br />

Coord Chem Rev 171:481<br />

449. Rau S, Schäfer B, Schebesta S, Grübing A, Poppitz W, Walther D, Duati M,<br />

Browne WR, Vos JG (2003) Eur J Inorg Chem, p 1503<br />

450. Loiseau F, Marzanni G, Quici S, Indelli MT, Campagna S (2003) Chem Commun,<br />

p 286<br />

451. Bergamini G, Saudan C, Ceroni P, Maestri M, Balzani V, Gorka M, Lee SK, van Heyst J,<br />

Vögtle F (2004) J Am Chem Soc 126:16466<br />

452. Falz JA, Williams RM, Dilva MJJP, De Cola L, Pikramenou Z (2006) J Am Chem Soc<br />

128:4520<br />

453. Glazer EC, Madge D, Tor Y (2005) J Am Chem Soc 127:4190<br />

454. Treadway JA, Moss JA, Meyer TJ (1999) Inorg Chem 38:4386<br />

455. Gallaghaer LA, Serron SA, Wen X, Hornstein BJ, Dattelbaum DM, Schoonover JR,<br />

Meyer TJ (2005) Inorg Chem 44:2089


214 S. Campagna et al.<br />

456. Sykora M, Petruska MA, Alstrum-Acevedo J, Bezel I, Meyer TJ, Klimov VI (2006)<br />

J Am Chem Soc 128:9984<br />

457. Kim YI, Salim S, Huq MJ, Mallouk TE (1991) J Am Chem Soc 113:9561<br />

458. Kim YI, Atherton SJ, Brigham ES, Mallouk TE (1993) J Phys Chem 97:11802<br />

459. Saupe GB, Mallouk TE, Kim W, Schmehl RH (1997) J Phys Chem 101:2508<br />

460. Hara, Waraksa CC, Lean JT, Lewis BA, Mallouk TE (2000) J Phys Chem A 104:5275<br />

461. Morris ND, Suzuki M, Mallouk TE (2004) J Phys Chem A 108:9115<br />

462. Haga MA, Ali MDM, Koseki S, Fujimoto K, Yoshimura A, Nozaki K, Ohno T, Nakajima<br />

K, Stufkens D (1996) Inorg Chem 35:3335<br />

463. Di Pietro C, Serroni S, Campagna S, G<strong>and</strong>olfi MT, Ballardini R, Fanni S, Browne W,<br />

Vos JG (2002) Inorg Chem 41:2871<br />

464. Fanni S, Di Pietro C, Serroni S, Campagna S, Vos JG (2000) Inorg Chem Commun<br />

3:42<br />

465. Jukes TF, Adamo V, Hartl F, Belser P, De Cola L (2005) Coord Chem Rev 249:1327<br />

466. Belser P, De Cola L, Hartl F, Adamo V, Bozic B, Chriqui Y, Iyer VM, Jukes TF, Kühni J,<br />

Querol M, Roma S, Salluce N (2006) Adv Funct Mater 16:195<br />

467. Saes M, Bressler C, Abela R, Grolimund D, Johnson SL, Heimann PA, Chergui M<br />

(2003) Phys Rev Lett 90:47403<br />

468. Gawelda W, Johnson M, de Groot FMF, Abela R, Bressler C, Chergui M (2006) J Am<br />

Chem Soc 128:5001<br />

469. Benfatto M, Della Longa S, Hatada H, Hayakawa K, Gawelda W, Bressler C, Chergui M<br />

(2006) J Phys Chem B 110:14035


Top Curr Chem (2007) 280: 215–255<br />

DOI 10.1007/128_2007_137<br />

© Springer-Verlag Berlin Heidelberg<br />

Published online: 27 June 2007<br />

<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong><br />

<strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium<br />

Maria Teresa Indelli · Claudio Chiorboli · Franco Sc<strong>and</strong>ola (✉)<br />

Dipartimento di Chimica dell’Università, ISOF-CNR sezione di Ferrara, 44100 Ferrara,<br />

Italy<br />

snf@unife.it<br />

1 Introduction ................................... 216<br />

2 Mononuclear Species .............................. 218<br />

2.1 PolypyridineComplexes ............................ 218<br />

2.2 CyclometalatedComplexes ........................... 223<br />

3 Polynuclear <strong>and</strong> Supramolecular Species ................... 226<br />

3.1 HomobinuclearComplexes........................... 226<br />

3.2 Dyads....................................... 227<br />

3.2.1 Photoinduced Electron Transfer in Ru(II)-Rh(III) Polypyridine Dyads . . . 228<br />

3.2.2 Photoinduced Electron Transfer in Porphyrin-Rh(III) Conjugates . . . . . 234<br />

3.3 Triads<strong>and</strong>OtherComplexSystems ...................... 235<br />

3.4 PhotoinducedElectronCollection ....................... 239<br />

4 RhodiumComplexesasDNAIntercalators .................. 241<br />

4.1 SpecificBindingtoDNA<strong>and</strong>Photocleavage.................. 241<br />

4.2 Rh(III) Complexes in DNA-Mediated Long-Range Electron Transfer . . . . 245<br />

4.2.1 Rh(III) Complexes as Acceptors in Electron Transfer Reactions ....... 245<br />

4.2.2 Long Range Oxidative DNA Damage by Excited Rh(III) Complexes . . . . 248<br />

5 Conclusion .................................... 250<br />

References ....................................... 251<br />

Abstract Rhodium(III) polypyridine complexes <strong>and</strong> their cyclometalated analogues display<br />

photophysical properties <strong>of</strong> considerable interest, both from a fundamental viewpoint<br />

<strong>and</strong> in terms <strong>of</strong> the possible applications. In mononuclear polypyridine complexes,<br />

the photophysics <strong>and</strong> photochemistry are determined by the interplay between LC <strong>and</strong><br />

MC excited states, with relative energies depending critically on the metal coordination<br />

environment. In cyclometalated complexes, the covalent character <strong>of</strong> the C – Rh bonds<br />

makes the lowest excited state classification less clear cut, with strong mixing <strong>of</strong> LC,<br />

MLCT, <strong>and</strong> LLCT character being usually observed. In redox reactions, Rh(III) polypyridine<br />

units can behave as good electron acceptors <strong>and</strong> strong photo-oxidants. These<br />

properties are exploited in polynuclear complexes <strong>and</strong> supramolecular systems containing<br />

these units. In particular, Ru(II)-Rh(III) dyads have been actively investigated for the<br />

study <strong>of</strong> photoinduced electron transfer, with specific interest in driving force, distance,<br />

<strong>and</strong> bridging lig<strong>and</strong> effects. Among systems <strong>of</strong> higher nuclearity undergoing photoinduced<br />

electron transfer, <strong>of</strong> particular interest are polynuclear complexes where rhodium<br />

dihalo polypyridine units, thanks to their Rh(III)/Rh(I) redox behavior, can act as twoelectron<br />

storage components. A large amount <strong>of</strong> work has been devoted to the use <strong>of</strong>


216 M.T. Indelli et al.<br />

Rh(III) polypyridine complexes as intercalators for DNA. In this role, they have proven<br />

to be very versatile, being used for direct str<strong>and</strong> photocleavage marking the site <strong>of</strong> intercalation,<br />

to induce long-distance photochemical damage or dimer repair, or to act as<br />

electron acceptors in long-range electron transfer processes.<br />

Keywords DNA intercalators · Electron transfer · <strong>Photophysics</strong> ·<br />

Polynuclear complexes · Rhodium<br />

Abbreviations<br />

bpy 2,2 ′ -bipyridine<br />

bzq Benzo(h)-quinoline<br />

chrysi 5,6-chrysenequinone diimine<br />

dpb 2,3-bis(2-pyridyl)benzoquinoxaline<br />

DPB 4,4 ′ -diphenylbipyridine<br />

dpp 2,3-bis(2-pyridyl)pyrazine<br />

dppz dipyridophenazine<br />

dpq 2,3-bis(2-pyridyl)quinoxaline<br />

HAT 1,4,5,8,9,12-hexaazatriphenylene<br />

Me2bpy 4,4 ′ -dimethyl-2,2 ′ -bipyridine<br />

Me2phen 4,7-dimethyl-1,10-phenanthroline<br />

Me2trien diamino-4,7-diazadecane<br />

ox Oxalato<br />

phen 1,10-phenanthroline<br />

phi 9,10-phenanthrenequinonediimine<br />

PPh3 triphenylphosphine<br />

ppy 2-phenylpyridine,<br />

TAP 1,4,5,8-tetraazaphenanthrene<br />

thpy 2-(2-thienyl)-pyridine<br />

tpy 2,2 ′ :6 ′ ,2 ′′ -terpyridine<br />

1<br />

Introduction<br />

Although not as popular as other transition metals, e.g., ruthenium, rhodium<br />

has received considerable attention in the field <strong>of</strong> inorganic photochemistry.<br />

Few specific reviews on rhodium photochemistry are available, however. The<br />

literature preceding 1970 was reviewed in the classical book <strong>of</strong> Carassiti <strong>and</strong><br />

Balzani [1]. The photochemistry <strong>of</strong> polypyridine metal complexes, including<br />

those <strong>of</strong> rhodium, has been reviewed by Kalyanasundaram in 1992 [2]. Several<br />

rhodium-containing species are considered in the extensive review written in<br />

1996 by Balzani <strong>and</strong> coworkers on luminescent <strong>and</strong> redox active polynuclear<br />

complexes [3]. A number <strong>of</strong> photochemical investigations are included in the<br />

1997 review article <strong>of</strong> Hannon on rhodium complexes [4]. Rhodium complexes<br />

are included in more recent reviews dealing with photoinduced processes in<br />

covalently linked systems containing metal complexes [5, 6].


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 217<br />

In general terms, three main classes <strong>of</strong> rhodium complexes have attracted<br />

the attention <strong>of</strong> photochemists: (i) amino complexes <strong>and</strong> substituted derivatives;<br />

(ii) multiply bridged dirhodium complexes; (iii) rhodium polypyridine<br />

<strong>and</strong> related complexes.<br />

Rhodium(III) halo/amino complexes have been actively investigated in the<br />

late 1970s <strong>and</strong> in the 1980s. They have low-lying metal centered (MC) excited<br />

stats <strong>of</strong> d – d type, <strong>and</strong> can be considered paradigmatic representatives<br />

<strong>of</strong> the lig<strong>and</strong>-field photochemistry <strong>of</strong> d 6 metal complexes. The subject has<br />

been summarized <strong>and</strong> clearly discussed by Ford <strong>and</strong> coworkers in their 1983<br />

review articles [7, 8]. In more recent times, however, despite a number <strong>of</strong> interesting<br />

investigations [9–18], the activity in the field seems to have slowed<br />

down considerably.<br />

Multiply bridged dirhodium complexes constitute a large family <strong>of</strong> compounds,<br />

the structure <strong>and</strong> properties <strong>of</strong> which depend strongly upon the oxidation<br />

state <strong>of</strong> the metals. The Rh(I)-Rh(I) species <strong>of</strong> formula Rh2(bridge)4 2+ ,<br />

where the two d 8 metal centers are bridged by four bidentate lig<strong>and</strong>s (e.g.,<br />

diisocyanoalkanes) in square planar coordination, have dσ ∗ → pσ excited<br />

states with a greatly shortened metal–metal bond [19] that emit efficiently in<br />

fluid solution [20]. In the Rh(II)-Rh(II) species <strong>of</strong> formula Rh2(bridge)4X2 n+ ,<br />

the two d 7 metal centers are bridged by four bidentate lig<strong>and</strong>s (e.g., diisocyanoalkanes,<br />

acetate) <strong>and</strong> complete their pseudo-octahedral coordination<br />

with a metal–metal bond <strong>and</strong> two-axial monodentate lig<strong>and</strong>s (e.g., X<br />

=Cl,Brn = 2). These dirhodium complexes have long-lived excited states<br />

<strong>of</strong> dπ ∗ → dσ ∗ type, which do not emit in fluid solution but can undergo<br />

a variety <strong>of</strong> bimolecular energy <strong>and</strong> electron transfer reactions [21].<br />

Dirhodium tetracarboxylato units <strong>of</strong> this type have also been used as building<br />

blocks for a variety <strong>of</strong> supramolecular systems <strong>of</strong> photophysical interest<br />

[22–24]. Particularly interesting triply bridged dirhodium complexes <strong>of</strong><br />

type X2Rh(bridge)3RhX2, LRh(bridge)3RhX2, <strong>and</strong>LRh(bridge)3RhL (bridge<br />

= bis(difluorophosphino)methylamine, X = Br, L = PPh3) have been developed<br />

recently by Nocera [25]. These Rh(II)-Rh(II), Rh(0)-Rh(II) <strong>and</strong><br />

Rh(0)-Rh(0) species, all possessing excited states <strong>of</strong> dπ ∗ → dσ ∗ type, can be<br />

interconverted photochemically by means <strong>of</strong> two-electron redox processes.<br />

Such two-electron photoprocesses provide the basis for a recently developed<br />

light-driven hydrogen production system [26], with a Rh(0)-Rh(II) mixed<br />

valence species playing the role <strong>of</strong> key photocatalysts [27]. The interest in<br />

multiply bridged dirhodium systems is now largely driven by their potential<br />

<strong>and</strong> implications for photocatalytic purposes.<br />

As for other transition metals, polyimine lig<strong>and</strong>s (in particular, polypyridines<br />

<strong>and</strong> their cyclometalated analogues) have played a major role in the design<br />

<strong>of</strong> rhodium complexes <strong>of</strong> photophysical interest. This is due to an ensemble<br />

<strong>of</strong> factors, including chemical robustness, synthetic flexibility, electronic<br />

structure, excited-state <strong>and</strong> redox tunability. Thus, rhodium polypyridine<br />

<strong>and</strong> related complexes have been extensively studied from a photophysical


218 M.T. Indelli et al.<br />

viewpoint, both as simple molecular species or as components <strong>of</strong> supramolecular<br />

systems featuring energy/electron transfer processes. Also, rhodium<br />

polypyridine complexes have played a major role in the active research field<br />

<strong>of</strong> DNA metal complex interactions. In recognition <strong>of</strong> the relevance <strong>of</strong> these<br />

systems, this review will be essentially focused on the photochemistry <strong>and</strong><br />

photophysics <strong>of</strong> complexes <strong>of</strong> rhodium with polypyridine-type lig<strong>and</strong>s <strong>and</strong><br />

<strong>of</strong> supramolecular systems that that contain such units as molecular components.<br />

2<br />

Mononuclear Species<br />

2.1<br />

Polypyridine Complexes<br />

The fundamental features <strong>of</strong> the photophysics <strong>of</strong> Rh(III) polypyridine<br />

complexes have been extensively discussed in the book <strong>of</strong> Kalyanasundaram<br />

[2] <strong>and</strong> only a few general aspects are recalled here. The tris(1,10phenanthroline)rhodium(III)<br />

ion, Rh(phen)3 3+ (1), can be used to exemplify<br />

the typical photophysical behavior <strong>of</strong> this class <strong>of</strong> complexes. Rh(phen)3 3+<br />

exhibits in 77 Kmatricesanintense(Φ,ca.1),long-lived(τ,ca.50 ms), structured<br />

emission (λ = 465, 485, 524, 571 nm) assigned as lig<strong>and</strong>-centered (LC)<br />

phosphorescence, i.e., emission from a π–π ∗ triplet state essentially localized<br />

on the phenanthroline lig<strong>and</strong>s [28–31]. As is shown by high-resolution<br />

spectroscopy, the LC excitation is not delocalized, but rather confined to<br />

a single lig<strong>and</strong> [32, 33]. In room-temperature fluid solutions, Rh(phen)3 3+ is<br />

practically non-emitting (see below). The LC triplet state can be nevertheless<br />

easily monitored by transient absorption spectroscopy (λmax = 490 nm,<br />

εmax = 4000 M –1 cm –1 , τ = 250 ns) [34]. The temperature-dependent behavior<br />

is explained on the basis <strong>of</strong> decay <strong>of</strong> the LC triplet via a thermally activated<br />

process involving an upper metal-centered (MC) state [34–36]. Indeed, the


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 219<br />

properties <strong>of</strong> the very weak emission measured from room-temperature solutions<br />

<strong>of</strong> Rh(phen)3 3+ (b<strong>and</strong>shape, lifetimes) are consistent with a small<br />

amount <strong>of</strong> MC excited state being in thermal equilibrium with the lowest LC<br />

state [34]. In related 2,2 ′ -bipyridine complexes, changes in LC-MC energy gap<br />

<strong>and</strong> emission spectral pr<strong>of</strong>ile can be induced by methyl substitution in the<br />

3,3 ′ positions, as a consequence <strong>of</strong> changes in the degree <strong>of</strong> planarity <strong>of</strong> the<br />

lig<strong>and</strong> [37].<br />

The interplay <strong>of</strong> LC <strong>and</strong> MC triplets in the photophysics <strong>of</strong> this type <strong>of</strong><br />

complexes has been investigated in detail by studying the series <strong>of</strong> mixedlig<strong>and</strong><br />

complexes cis-Rh(phen)2XY n+ (X=Y=CN,n =1;X=Y=NH3, n =3;<br />

X=NH3, Y=Cl,n = 2), where the energy <strong>of</strong> the MC states is controlled by<br />

the lig<strong>and</strong> field strength <strong>of</strong> the X, Y ancillary lig<strong>and</strong>s [38]. While at room temperature<br />

all the complexes are very weakly emissive, at 77 K, the di-cyano <strong>and</strong><br />

di-amino complexes give the typical, structured LC emission, whereas the<br />

amino-chloro complex exhibits a broad emission <strong>of</strong> MC type (Fig. 1a). This<br />

can be readily explained on the basis <strong>of</strong> the energy diagram <strong>of</strong> Fig. 1b, where<br />

the relative energies <strong>of</strong> the LC triplet (appreciably constant for the three complexes)<br />

<strong>and</strong> <strong>of</strong> the lowest MC state (dependent on the lig<strong>and</strong> field strength <strong>of</strong><br />

the ancillary lig<strong>and</strong>s) are schematically depicted. For the amino-chloro complex<br />

the MC state is the lowest excited state <strong>of</strong> the system. For the di-amino<br />

case the situation is similar to that <strong>of</strong> the Rh(phen)3 3+ complex, with LC as<br />

the lowest excited state but with MC sufficiently close in energy to provide<br />

an efficient thermally activated decay path for LC (actually, in an appropriate<br />

temperature regime the two states are in equilibrium). In the di-cyano<br />

complex, the MC state is sufficiently high in energy that the LC state has<br />

a substantial lifetime (1.2 µs) even in room-temperature solution [38]. The<br />

actual energy gap between the LC <strong>and</strong> MC states (2000 cm –1 for the di-cyano<br />

Fig. 1 a 77 K emission spectra <strong>of</strong> cis-Rh(phen)2XY n+ complexes with different X, Y ancillary<br />

lig<strong>and</strong>s. b Rationalization <strong>of</strong> the emission properties in terms <strong>of</strong> relative energies <strong>of</strong><br />

lig<strong>and</strong>-centered (LC) <strong>and</strong> metal-centered (MC) excited states (adapted from [38])


220 M.T. Indelli et al.<br />

complex) can be evaluated by measuring the activation energy <strong>of</strong> the photosolvation<br />

reaction originating from the MC state in low-temperature glycerol<br />

matrices [39].<br />

As it is obvious from the trend sketched in Fig. 1b, the cis-Rh(phen)2Cl2 +<br />

complex has again a lowest MC excited state. It emits at ca. 710 nm [31] <strong>and</strong>, in<br />

keeping with the typical lig<strong>and</strong>-field photoreactivity <strong>of</strong> d 6 metal complexes [7],<br />

undergoes in fluid solution photosolvation <strong>of</strong> the chloride lig<strong>and</strong>s [40–42].<br />

This type <strong>of</strong> reactivity that has been exploited by Morrison <strong>and</strong> coworkers [43]<br />

in an extensive series <strong>of</strong> studies on photoinduced binding to DNA <strong>and</strong> potential<br />

applications <strong>of</strong> this class <strong>of</strong> complexes as photo-toxic agents.<br />

Anumber<strong>of</strong>analogues<strong>of</strong>thecis-Rh(NN)2Cl2 + complexes, where the NN<br />

represents various bidentate nitrogen donors, such as e.g., 2, 3 behave similarly<br />

to cis-Rh(phen)2Cl2 + , i.e., have lowest excited states <strong>of</strong> MC character <strong>and</strong><br />

emit accordingly [44, 45].<br />

Thesameline<strong>of</strong>reasoningcanbeappliedtorationalizethephotophysical<br />

properties <strong>of</strong> the bis(2,2 ′ :6 ′ ,2 ′′ -terpyridine)rhodium(III) ion, Rh(tpy) 3+<br />

2 (4),<br />

<strong>and</strong> analogous complexes. It is well known that, owing to a more unfavorable


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 221<br />

bite angle, two tpy lig<strong>and</strong>s provide a lower lig<strong>and</strong> field strength compared<br />

to three phen (or bpy) lig<strong>and</strong>s [46]. Accordingly, whereas Rh(bpy) 3+<br />

3 is a LC<br />

emitter, Rh(tpy) 3+ <strong>and</strong> related compounds only show at 77 Kbroadstructure-<br />

2<br />

less emissions <strong>of</strong> MC type [47].<br />

In heteroleptic bis-imine Rh(III) complexes, multiple emissions originating<br />

from LC states localized on different lig<strong>and</strong>s can be present at<br />

77 K. For example, in complexes <strong>of</strong> type [Rh(bpy)n(phen)3–n] 3+ , the LC<br />

emissions localized on bpy <strong>and</strong> phen are spectrally very similar but can<br />

be distinguished on the basis <strong>of</strong> their different lifetimes in time-resolved<br />

experiments [30, 48]. Clear indication that the excitation is trapped at either<br />

a bipyridyl or a phenanthroline lig<strong>and</strong> in the phosphorescent triplet state<br />

can also be obtained from phosphorescence microwave double resonance<br />

(PMDR) experiments [49]. The heteroleptic complex Rh(phi)2(phen) 3+ (5)<br />

has been particularly developed as a DNA photocleavage agent (see Sect. 4).<br />

The complexes absorb strongly in UV region with a significant broad absorption<br />

in the visible range, which tails to ca. 500 nm [50]. The spectrum<br />

is dominated by overlapping <strong>of</strong> the lig<strong>and</strong> centered (LC) transitions <strong>of</strong> the<br />

component lig<strong>and</strong>s with the phi centered b<strong>and</strong>s at lower energy. These b<strong>and</strong>s<br />

are strongly pH dependent with shifts to the blue upon increasing the pH.<br />

At 77 K the complex exhibits lig<strong>and</strong> centered (LC) dual emission from both<br />

phi <strong>and</strong> phen lig<strong>and</strong>s. At room temperature no emission can be detected<br />

whereas a long-lived excited state (τ ≈ 200 ns in polar solvent) has been<br />

observed by transient absorption. On the basis <strong>of</strong> several experimental evidences<br />

this excited state, lying in energy at ca. 2 eV above the ground state,<br />

is assigned by the authors to be intralig<strong>and</strong> charge transfer in nature (ILCT).<br />

Quenching experiments with organic electron donors clearly indicate that the<br />

ILCT triplet state is a strong oxidizing agent with E1/2( ∗Rh3+ /Rh2+ )=2.0 V<br />

vs. NHE [50]. Multiple LC emissions have also been suggested to occur in<br />

heteroleptic complexes containing bipyridine or phenanthroline <strong>and</strong> pyridyl<br />

triazole lig<strong>and</strong>s [51].<br />

While for the vast majority <strong>of</strong> Rh(III) polypyridine complexes the photophysics<br />

<strong>and</strong> photochemistry are dominated by LC <strong>and</strong> MC states, in a few


222 M.T. Indelli et al.<br />

cases lig<strong>and</strong>-to-metal charge transfer (LMCT) photochemistry is observed.<br />

A clear example is provided by the Rh(bpy)2(ox) + complex (6) [52].<br />

The spectrum <strong>of</strong> the colorless complex 6 is characterized by an intense<br />

b<strong>and</strong> at ca. 300 nm assigned to oxalato-to-rhodium LMCT transitions. Upon<br />

UV irradiation, the following photoreaction is observed:<br />

Rh III (ox)(bpy)2 + + hν → Rh I (bpy)2 + + 2CO2 . (1)<br />

The Rh(bpy)2 + product is formed rapidly (risetime in pulsed experiments,<br />

ca. 10 ns), probably via a sequence <strong>of</strong> processes comprising photochemical intramolecular<br />

electron transfer from the oxalate lig<strong>and</strong> to Rh(III) followed by<br />

the decomposition <strong>of</strong> the oxidized lig<strong>and</strong> into CO2 <strong>and</strong> CO2 – radical (Eq. 2)<br />

<strong>and</strong> thermal reduction <strong>of</strong> the Rh(II) center to Rh(I) by the reactive CO2 – radical<br />

(Eq. 3) [52].<br />

RhIII (ox)(bpy)2 + + hν → RhII (CO2 – )(bpy)2 + +CO2<br />

(2)<br />

RhII (CO2 – )(bpy)2 + → RhI (bpy)2 + +CO2 . (3)<br />

The violet Rh(bpy)2 + product, with intense MLCT visible absorption, is<br />

a tetrahedrally distorted d 8 square planar complex [53]. This Rh(I) species,<br />

which can also be obtained by chemical [54], electrochemical [55], or radiation<br />

chemical [56, 57] reduction <strong>of</strong> Rh(III) complexes, is <strong>of</strong> great interest from<br />

the catalytic viewpoint. It undergoes facile oxidative addition by molecular<br />

hydrogen [58], to give the corresponding Rh(III) dihydride (Eq. 4). The reaction<br />

is fully reversible upon<br />

Rh(bpy)2 + +H2 ⇆ cis-Rh III (bpy)2(H)2 + (4)<br />

removal <strong>of</strong> molecular hydrogen from the system. Interestingly, the release <strong>of</strong><br />

molecular hydrogen from the dihydride complex can be obtained photochemically<br />

(Eq. 5). This photoreaction provides a<br />

cis-RhIII (bpy)2(H)2 –→Rh(bpy)2 + +H2<br />

hν<br />

(5)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 223<br />

convenient means to perturb the equilibrium <strong>of</strong> Eq. 4 <strong>and</strong> to study the kinetics<br />

<strong>of</strong> hydrogen uptake in the thermal relaxation <strong>of</strong> the perturbed system.<br />

A detailed picture <strong>of</strong> the transition state <strong>of</strong> this interesting reaction has been<br />

obtained by such type <strong>of</strong> experiments [53].<br />

2.2<br />

Cyclometalated Complexes<br />

Lig<strong>and</strong>s such as 2-phenylpyridine, 2-(2-thienyl)-pyridine, or benzo(h)quinoline,<br />

in their ortho-deprotonated forms (ppy, 7;thpy,8;bzq,9), can bind<br />

in a bidentate N^C fashion to a variety <strong>of</strong> transition metals, including Rh(III).<br />

These complexes are generally indicated as cyclometalated (or orthometalated)<br />

complexes. The spectroscopy <strong>and</strong> photophysics <strong>of</strong> cyclometalated<br />

complexes [59] are usually very different form those <strong>of</strong> the corresponding<br />

polypyridine complexes, the main reason being the much stronger σ donor<br />

character <strong>of</strong> C – relative to N. The consequences are (i) a high degree <strong>of</strong> covalency<br />

in the carbon–metal bond, (ii) a strongly enhanced ease <strong>of</strong> oxidation<br />

<strong>of</strong> the metal, (iii) high-energy MC excited states, (iv) relatively low-energy<br />

MLCT excited states.<br />

The photophysics <strong>of</strong> Rh(III) cyclometalated complexes, though not as<br />

developed as that <strong>of</strong> analogous Ir(III) species (see Chap. 9), has been actively<br />

investigated in the last two decades. While some tris- [60] <strong>and</strong> monocyclometalated<br />

[61] compounds have been synthesized <strong>and</strong> studied, for synthetic<br />

reasons bis-cyclometalated complexes <strong>of</strong> Rh(III) are by far more common<br />

in the literature. All the bis-cyclometalated complexes have a C,C cis<br />

geometry [62]. The Rh(ppy)2(bpy) + (10) complex can be used here to exemplify<br />

the main photophysical features <strong>of</strong> this class <strong>of</strong> compounds.


224 M.T. Indelli et al.<br />

The absorption spectrum <strong>of</strong> 10 (Fig. 2) shows LC transitions <strong>of</strong> the bpy<br />

<strong>and</strong> ppy lig<strong>and</strong>s in the 240–310 nm range. In addition, a prominent b<strong>and</strong><br />

is present at 366 nm, which is attributed to MLCT transitions [63, 64]. The<br />

presence <strong>of</strong> MLCT transitions at relatively low energy, a feature completely<br />

absent in analogous polypyridine complexes, is the result <strong>of</strong> the strongly<br />

electron donating character <strong>of</strong> the cyclometalating lig<strong>and</strong>, which makes the<br />

formally Rh(III) center relatively easy to oxidize (irreversible wave observed<br />

at ca. + 1.1 V vs. NHE) [63]. The complex, which is only very weakly emissive<br />

at room temperature, exhibits an intense, long-lived (τ, 170 µs), structured<br />

emission at 77 K (Fig. 2). The emission has been attributed to ppy-based LC<br />

phosphorescence, although various degrees <strong>of</strong> mixing between LC <strong>and</strong> with<br />

MLCT triplet states have been invoked on the basis <strong>of</strong> absence <strong>of</strong> dual emissions<br />

[64], lifetime considerations [63], <strong>and</strong> high-resolution spectroscopy [65,<br />

66]. In fact, because <strong>of</strong> the strong covalency <strong>of</strong> the carbon–metal bonds, the<br />

HOMO in this class <strong>of</strong> complexes has a mixed character, being delocalized<br />

on the central metal <strong>and</strong> on the cyclometalating lig<strong>and</strong>. This has been clearly<br />

shown by a number <strong>of</strong> recent TD/DFT calculations performed on Rh(III) [67]<br />

<strong>and</strong> related Ir(III) [68, 69] cyclometalated complexes. Therefore, a classifica-<br />

Fig. 2 Absorption spectrum (dashed line) <strong>and</strong> emission spectrum (room-temperature,<br />

continuous line; 77 K, dotted line) <strong>of</strong> Rh(ppy)2(bpy) + in 4/1 EtOH/MeOH (adapted<br />

from [64])


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 225<br />

tion <strong>of</strong> the excited states in classical, localized terms (MC, LC, MLCT) is not<br />

generally applicable for cyclometalated complexes.<br />

The general behavior <strong>of</strong> other Rh(N^C)2(N^N) + complexesissimilar<br />

to that described above for Rh(ppy)2(bpy) + , with some easily underst<strong>and</strong>able<br />

specific differences. For instance, the main effect <strong>of</strong> substituting thpy<br />

(8) for ppy is an enhancement <strong>of</strong> emission intensity <strong>and</strong> lifetime in roomtemperature<br />

solutions [63]. This is likely a result <strong>of</strong> the lower energy <strong>of</strong><br />

the emission <strong>and</strong> the consequent lower efficiency <strong>of</strong> the thermally activated<br />

decay [70] via higher MC states. The remarkably long-lived (4.4–6.6 µs)<br />

emission observed in fluid solution for an analog <strong>of</strong> 10 with aldehyde substituents<br />

on the phenyl ring <strong>of</strong> the cyclometalating lig<strong>and</strong> [71] is probably<br />

justified by the same type <strong>of</strong> argument. On the other h<strong>and</strong>, the photophysical<br />

behavior <strong>of</strong> 10 is only slightly affected by substitution <strong>of</strong> bpy with similar<br />

N^N lig<strong>and</strong>s, such as, e.g., 1,10-phenanthroline, 2,2 ′ -biquinoline [72],<br />

or 4-amino-3,5-bis(2-pyridyl)-4H-1,2,4-triazole) (11) [73]. However, when<br />

strong π-deficient lig<strong>and</strong>s such as 1,4,5,8-tetraazaphenanthrene (TAP, 12)<br />

or 1,4,5,8,9,12-hexaazatriphenylene (HAT, 13) are used as N^N lig<strong>and</strong> [74],<br />

a definite switch in behavior is observed, with the presence <strong>of</strong> broad structureless<br />

77 K emissions that have a clear charge transfer character. Indeed, for<br />

this type <strong>of</strong> complexes, TD/DFT calculations show that the HOMO involves<br />

the metal <strong>and</strong> the cyclometalating lig<strong>and</strong>-carbon bonds, but the LUMO is now<br />

exclusively localized on the non-cyclometalating lig<strong>and</strong> (Fig. 3) [67]. Thus,<br />

in this case the emission is best considered as having a mixed LLCT/MLCT<br />

character.


226 M.T. Indelli et al.<br />

Fig. 3 Frontier orbitals <strong>of</strong> Rh(ppy)2TAP + . From [67]<br />

3<br />

Polynuclear <strong>and</strong> Supramolecular Species<br />

3.1<br />

Homobinuclear Complexes<br />

A few homobinuclear lig<strong>and</strong>-bridged Rh(III) polypyridine complexes have<br />

been studied [3, 75, 76]. Their photochemical interest is rather limited, however,<br />

as they behave generally like their mononuclear analogues, with minor<br />

differences in spectral shifts <strong>and</strong> lifetimes. An interesting type <strong>of</strong> systems,<br />

which bring together the complexities <strong>of</strong> lig<strong>and</strong>-bridged species <strong>and</strong> multiply<br />

bridged metal–metal bonded rhodium dimers, has recently been reported by<br />

Campagna et al. [77]. In (14) two quadruply bridged Rh(II)-Rh(II) dimers are<br />

the “molecular components” <strong>of</strong> a higher-order two-component system held<br />

together by the complex bis-naphthyridine-type lig<strong>and</strong>. As already observed<br />

for some related simple rhodium dimers (e.g., Rh2(CH3COO)4(PPh3)2) [21],<br />

the “binuclear” compound has a non-emissive but long-lived excited state in<br />

room-temperature solution. In the case <strong>of</strong> 14, the long-lived state has been assigned<br />

as an MLCT (metal–metal π ∗ to naphthyridine π ∗ )excitedstate[77].


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 227<br />

3.2<br />

Dyads<br />

Heteronuclear bimetallic species containing rhodium polypyridine complexes<br />

are more interesting, as the Rh(III) unit can be involved in intercomponent<br />

processes, particularly <strong>of</strong> the electron transfer type. The thermodynamic<br />

requirements for the participation <strong>of</strong> Rh(III) polypyridine complexes in electron<br />

transfer processes are summarized in Fig. 4, where Rh(III), ∗ Rh(III), <strong>and</strong><br />

Rh(II) represent the ground state, the triplet LC excited state (see Sect. 2.1),<br />

<strong>and</strong> the one-electron reduced form, respectively, <strong>and</strong> the values <strong>of</strong> excitedstate<br />

energy [31] <strong>and</strong> reduction potential [78] refer to Rh(phen)3 3+ (1). From<br />

these figures, it is apparent that Rh(III) polypyridine complexes can behave<br />

as extremely powerful photochemical oxidants <strong>and</strong> relatively good electron<br />

transfer quenchers. On the other h<strong>and</strong>, because <strong>of</strong> the high excited-state energy,<br />

these complexes are also good potential energy donors.<br />

Fig. 4 Typical redox energy level diagram for Rh(III) polypyridine complexes. Values<br />

(reduction potential vs. SCE) appropriate for Rh(phen)3 3+ (1)<br />

As a matter <strong>of</strong> fact, Rh(III) polypyridine complexes have been extensively<br />

used in bimolecular electron transfer processes, either as photoexcited<br />

molecule [79, 80] or as quencher [81–83], with motivations <strong>of</strong> both fundamental<br />

(testing electron transfer-free energy relationships) [80] <strong>and</strong> applied<br />

nature (photoinduced hydrogen evolution from water) [81, 82]. Here, on the<br />

other h<strong>and</strong>, we focus our attention on photoinduced processes where the reactants<br />

are pre-assembled in some kind <strong>of</strong> supramolecular system. The most<br />

common photoinduced processes taking place in simple two-component systems<br />

(<strong>of</strong>ten called “dyads”) involving a Rh(III) polypyridine unit are shown<br />

in Eqs. 6–8:<br />

∗ Rh(III) –Q → Rh(III) – ∗ Q (6)<br />

∗ Rh(III) –Q → Rh(II) –Q + (7)<br />

∗ P – Rh(III) → P + – Rh(II) . (8)


228 M.T. Indelli et al.<br />

The dyads generically indicated in Eqs. 6–8 are actually heterobinuclear complexes<br />

when, as it is <strong>of</strong>ten the case, the photosensitizer P or the quencher Q<br />

are transition metal complex moieties, <strong>and</strong> the chemical linkage is provided<br />

by a bridging lig<strong>and</strong>. A number <strong>of</strong> examples <strong>of</strong> such processes are discussed<br />

in the following sections.<br />

3.2.1<br />

Photoinduced Electron Transfer in Ru(II)-Rh(III) Polypyridine Dyads<br />

Though not exclusively, dyads containing Rh(III) polypyridine units have<br />

<strong>of</strong>ten involved as P or Q (Eqs. 7, 8) the chromophores par excellence <strong>of</strong> inorganic<br />

photochemistry, namely Ru(II) polypyridine complexes. The general<br />

behavior <strong>of</strong> Ru(II)-Rh(III) polypyridine dyads can be discussed taking dyad<br />

15 as an example [84].<br />

The absorption spectrum <strong>of</strong> dyad 15, as compared with those <strong>of</strong> the<br />

Ru(Me2phen)2(Me2bpy) 2+ <strong>and</strong> Rh(Me2bpy)3 3+ molecular components, is<br />

shown in Fig. 5. It shows that the spectra <strong>of</strong> the molecular components are<br />

strictly additive, as expected for weak intercomponent interaction, <strong>and</strong> that<br />

selective (100%) excitation <strong>of</strong> the Ru(II) chromophore can be easily performed<br />

in the visible region, whereas partial excitation <strong>of</strong> the Rh(III) component<br />

(ca. 70% at300 nm) can be achieved in the ultraviolet. The energy<br />

level diagram for this dyad (Fig. 6) shows that, besides the photophysical<br />

processes taking place within each molecular component, a number <strong>of</strong> intercomponent<br />

processes are thermodynamically allowed. They include:<br />

∗Ru(II)-Rh(III) → Ru(III)-Rh(II) electron transfer from excited Ru(II)<br />

(a)<br />

Ru(II)-∗Rh(III) → Ru(III)-Rh(II) electron transfer to excited Rh(III)<br />

(b)<br />

Ru(II)-∗Rh(III) → ∗Ru(II)-Rh(III) energy transfer from Rh(III) to Ru(II)<br />

(c)<br />

Ru(III)-Rh(II) → Ru(II)-Rh(III) back electron transfer . (d)


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 229<br />

Fig. 5 Absorption spectra <strong>of</strong> the Ru(II)-Rh(III) dyad 15 (dotted line, c), <strong>and</strong> <strong>of</strong> the<br />

Ru(Me2phen)2(Me2bpy) 2+ (dashed line, b) <strong>and</strong>Rh(Me2bpy)3 3+ (continuous line, a) molecular<br />

components<br />

Fig. 6 Energy level diagram <strong>and</strong> photophysical processes for the Ru(II)-Rh(III) dyad 15<br />

For dyad 15, all these processes could be time-resolved by nanosecond <strong>and</strong> picosecond<br />

techniques under the appropriate experimental conditions (visible<br />

excitation for a, UV excitation for b <strong>and</strong> d, rigid matrix for c), leading to the


230 M.T. Indelli et al.<br />

detailed kinetic picture <strong>of</strong> Fig. 6 [84]. The widely different rates <strong>of</strong> the three<br />

ET processes (a, b, d) can be rationalized in terms <strong>of</strong> predominant driving<br />

force effects [84], as shown schematically in Fig. 7.<br />

Fig. 7 Free-energy correlation <strong>of</strong> rate constants for the three electron transfer processes<br />

<strong>of</strong> dyad 15<br />

Since most Ru(II)-Rh(III) polypyridine dyads have very similar energetics,<br />

the qualitative features illustrated above for dyad 15 can be safely generalized<br />

to this whole class <strong>of</strong> compounds. For example, in dyad 16 [85]therateconstants<br />

<strong>of</strong> processes a, b,<strong>and</strong>d are slower by a factor <strong>of</strong> ca. 3 but have the same<br />

relative magnitudes as for dyad 15. The slower rates are likely related to the<br />

longer aliphatic bridge, although for this <strong>and</strong> related [86] dyads, the flexibility<br />

<strong>of</strong> the bridges limits the validity <strong>of</strong> such comparisons.<br />

Within this general type <strong>of</strong> behavior <strong>of</strong> Ru(II)-Rh(III) dyads, a number <strong>of</strong><br />

experimental studies have been specifically aimed at investigating the role <strong>of</strong><br />

the bridge in determining electron transfer rates. As has been the case for<br />

other types <strong>of</strong> bimetallic dyads [87–90], particular attention has been devoted<br />

to Ru(II)-Rh(III) dyads with modular bridges involving p-phenylene<br />

spacer units [6, 91, 92]. The dyads in Chart 1 provide a homogeneous series


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 231<br />

with appreciably constant energetics but differing in the number (1–3) <strong>of</strong><br />

p-phenylene spacers in the bridge <strong>and</strong>, in one <strong>of</strong> the two dyads with three<br />

spacers, for the presence <strong>of</strong> alkyl substituents on the central spacer. Upon<br />

excitation <strong>of</strong> the Ru(II)-based chromophore, the rate constants <strong>of</strong> photoinduced<br />

electron transfer (process a in the above general scheme) have been<br />

measured by time-resolved emission <strong>and</strong> transient absorption techniques in<br />

the nanosecond <strong>and</strong> picosecond time domains [6, 92]. The values for Ru-ph-<br />

Rh, Ru-ph2-Rh, <strong>and</strong>Ru-ph3-Rh (Chart 1), when plotted as a function <strong>of</strong> the<br />

metal–metal distance r (Fig. 8), display an exponential decrease (Eq. 9):<br />

k = k(0) exp(– βr). (9)<br />

Chart 1


232 M.T. Indelli et al.<br />

Fig. 8 Distance dependence <strong>of</strong> photoinduced electron transfer rates in the dyads <strong>of</strong><br />

Chart 1: Ru-ph-Rh, Ru-ph2-Rh, Ru-ph3-Rh (dots), Ru-ph ′ 3-Rh (triangle)<br />

This is the behavior predicted for electron transfer in the superexchange<br />

regime [5,93,95] if the distance dependence <strong>of</strong> the reorganizational energy<br />

term can be neglected. The β valueobtainedfromtheslope<strong>of</strong>thelinein<br />

Fig. 8, 0.65 ˚A –1 , should be regarded as an upper limiting value for the attenuation<br />

factor <strong>of</strong> the intercomponent electronic coupling (Eq. 9). This β value<br />

is in the range found for other oligophenylene-containing systems (organic<br />

dyads [96, 97], metal–molecules–metal junctions [98]). This underlines the<br />

goodability<strong>of</strong>thistype<strong>of</strong>bridgestomediatedonor–acceptorelectroniccoupling<br />

(for comparison, β is typically 0.8–1.2 ˚A –1 for rigid aliphatic bridges). In<br />

this regard, it is instructive to compare the electron transfer rate constant observed<br />

for Ru-ph-Rh (k = 3.0 × 109 s –1 ) with that mentioned above for dyad<br />

15 containing an aliphatic bis-methylene bridge (k = 1.7 × 108 s –1 ). Despite the<br />

longer metal–metal distance (15.5 ˚A for Ru-ph-Rh relative to 13.5 ˚A for 15),<br />

the reaction is faster across the phenylene spacer by more than one order <strong>of</strong><br />

magnitude.<br />

An interesting result [6, 92] is the fact that dyad Ru-ph ′ 3-Rh, which is iden-<br />

tical to Ru-ph3-Rh except for the presence <strong>of</strong> two solubilizing hexyl groups on<br />

the central phenylene ring, is one order <strong>of</strong> magnitude slower than its unsubstituted<br />

analog (Fig. 8). This is related to the notion that in a superexchange<br />

mechanism the rate is sensitive to the electronic coupling between adjacent<br />

modules <strong>of</strong> the spacer [5, 93, 95], <strong>and</strong> that in polyphenylene bridges this coupling<br />

is a sensitive function <strong>of</strong> the twist angle between adjacent spacers [99].<br />

While the planes <strong>of</strong> unsubstituted adjacent phenylene units form angles <strong>of</strong>.<br />

20 ◦ –40 ◦ [100, 101], ring substitution leads to a substantial increase in the<br />

twist angle (to ca. 70 ◦ ) [100] <strong>and</strong>, as a consequence, to a slowing down <strong>of</strong> the<br />

electron transfer process.<br />

A number <strong>of</strong> Ru(II)-Rh(III) dyads have been reported where little or any<br />

photoinduced electron transfer quenching <strong>of</strong> the Ru(II)-based MLCT emis-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 233<br />

sion takes place [75, 76, 102]. In the case <strong>of</strong> dyad 17, the plausible reason is<br />

that, owing to the comparable reduction potentials <strong>of</strong> the diphenylpyrazine<br />

bridging lig<strong>and</strong> <strong>and</strong> Ru(III) center, the driving force for intramolecular electron<br />

transfer is too small [102]. In the cases <strong>of</strong> dyads 18 <strong>and</strong> 19, the presence<br />

<strong>of</strong> cyclometalated ancillary lig<strong>and</strong>s makes the formally Rh(III) center very<br />

difficult to reduce <strong>and</strong> relatively easy to oxidize, thus yielding MLCT states at<br />

comparable energies on the two units [75, 76].<br />

Low driving force arguments could also apply to the dyad 20, where<br />

relatively slow quenching <strong>of</strong> the Ru(II) MLCT emission (estimated k,<br />

ca. 3.5 × 10 7 s –1 ) was observed <strong>and</strong> attributed to intramolecular electron<br />

transfer [103]. Here, however, a relevant aspect is also the presence a Rh(III)


234 M.T. Indelli et al.<br />

moiety with <strong>of</strong> coordinated chloride lig<strong>and</strong>s. In fact, contrary to what happens<br />

for common Rh(III) polypyridine units (e.g., Rh(bpy)3 3+ ,Rh(phen)3 3+ )<br />

where one-electron reduction is a quasi-reversible process, mixed-lig<strong>and</strong><br />

units containing halide ions (e.g., Rh(bpy)2Cl2 + ,Rh(phen)2Cl2 + )undergo<br />

strongly irreversible two-electron reductions accompanied by prompt halide<br />

lig<strong>and</strong> loss [78, 104]. While the use <strong>of</strong> these units as electron acceptors can<br />

be <strong>of</strong> interest towards photoinduced electron collection <strong>and</strong> multi-electron<br />

catalysis (see Sect. 3.4), from a kinetic viewpoint the large reorganizational<br />

energies involved are likely to lead to slow electron transfer rates.<br />

3.2.2<br />

Photoinduced Electron Transfer in Porphyrin-Rh(III) Conjugates<br />

Though structurally quite different, the porphyrin-Rh(III) conjugates thoroughly<br />

studied by Harriman et al. [105] behave with regard to photoinduced<br />

electron transfer rather similarly to the above-discussed Ru(II)-Rh(III) dyads.<br />

The systems contain a zinc porphyrin unit connected directly with one (21)or<br />

via a phenylene spacer with two (22) rhodium terpyridine units.<br />

The electron transfer processes thermodynamically allowed in these systems<br />

are indicated in Eqs. 10–14, where both 21 <strong>and</strong> 22 are schematized as


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 235<br />

dyads (indeed, 22 is a triad only in a formal sense), ZnP, Rh, <strong>and</strong> ph represent<br />

the zinc porphyrin <strong>and</strong> rhodium terpyridine molecular components <strong>and</strong> the<br />

phenylene spacer, respectively. The singlet excited state is considered for the<br />

zinc porphyrin chromophore <strong>and</strong> the triplet state for the rhodium terpyridine<br />

unit.<br />

∗ + ◦<br />

ZnP-ph-Rh(III) → ZnP -ph-Rh(II) ∆G =–0.71 eV (10)<br />

ZnP-ph-Rh(III) ∗ → ZnP + -ph-Rh(II) ∆G ◦ =–0.81 eV (11)<br />

ZnP + -ph-Rh(II) → ZnP-ph-Rh(III) ∆G ◦ =–1.47 eV (12)<br />

∗ + ◦<br />

ZnP-Rh(III) → ZnP -Rh(II) ∆G =–0.58 eV (13)<br />

ZnP-Rh(III) ∗ → ZnP + -Rh(II) ∆G ◦ =–0.80 eV (14)<br />

In 22, where a phenylene spacer is interposed between the two molecular<br />

components, both photoinduced electron transfer following excitation <strong>of</strong> the<br />

zinc porphyrin (Eq. 10) <strong>and</strong> charge recombination (Eq. 12) have been time<br />

resolved, with values in acetonitrile <strong>of</strong> 3.2 × 1011 s –1 <strong>and</strong> 8.3 × 109 s –1 ,respectively.<br />

The wide difference in rates is attributed to the different kinetic<br />

regimes <strong>of</strong> the two processes, photoinduced electron transfer (Eq. 10) being<br />

almost activationless, <strong>and</strong> charge recombination (Eq. 12) lying deep into the<br />

Marcus inverted region [105]. For the directly linked system 21 (as well as for<br />

its free-base analogue), the disappearance <strong>of</strong> the porphyrin excited state, presumablybyphotoinducedelectrontransfer(Eq.13),isextremelyfast.Infact,<br />

the excited state lifetime <strong>of</strong> 21, ca.0.7 ps in acetonitrile, is comparable to the<br />

longitudinal relaxation time <strong>of</strong> the solvent, implying that the electron transfer<br />

process in this system is controlled by solvent reorientation [105].<br />

3.3<br />

Triads <strong>and</strong> Other Complex Systems<br />

In dyads, including those discussed in the previous sections, photoinduced<br />

electron transfer is always followed by fast charge recombination. This greatly<br />

limits the use <strong>of</strong> dyads for practical purposes such as, e.g., conversion <strong>of</strong> light


236 M.T. Indelli et al.<br />

into chemical energy. A strategy to overcome this problem, largely inspired<br />

by the architecture <strong>of</strong> natural photosynthetic reaction centers, is that <strong>of</strong> going<br />

to more complex supramolecular systems, triads, etc., in which a sequence <strong>of</strong><br />

electron transfer steps is used to achieve long-range charge separation. The<br />

simplest <strong>of</strong> such system is a triad, as schematically illustrated in Fig. 9 for two<br />

possible reaction schemes. In Fig. 9a, the consecutive ET steps are from the<br />

excited chromophore, P, to an acceptor molecular component, A, <strong>and</strong> from<br />

a donor unit, D, to the oxidized chromophore. In Fig. 9b, the two consecutive<br />

electron transfer steps are from the excited chromophore, P, to a primary<br />

acceptor, A, <strong>and</strong> from the primary to a secondary acceptor unit, A ′ .These<br />

strategies have been extensively implemented using organic [106, 107] <strong>and</strong>, to<br />

a lesser extent, inorganic [108–110] molecular components.<br />

Fig. 9 Two types <strong>of</strong> triads for photoinduced charge separation. Molecular components:<br />

P (chromophore), D (donor), A (acceptor), A ′ (secondary acceptor). Electron transfer<br />

processes: cs (primary PET), cr (primary charge recombination), cs ′ (secondary charge<br />

separation), cr ′ (final charge recombination)<br />

A number <strong>of</strong> systems involving Rh(III) molecular components that behave<br />

in some respects as triads (or pseudo-triads) are discussed in this section.<br />

The trimetallic species 23 has been synthesized <strong>and</strong> studied by Petersen<br />

<strong>and</strong> coworkers [111] as a possible supramolecular system for photoinduced<br />

multi-step charge separation. This system comprises a Fe(II) electron donor,<br />

a Ru(II) photoexcitable chromophore, <strong>and</strong> a Rh(III) unit carrying a “monoquat”<br />

acceptor as lig<strong>and</strong>. Two different types <strong>of</strong> bridging lig<strong>and</strong> are present<br />

in 23, a bipyrimidine between Fe(II) <strong>and</strong> Ru(II) <strong>and</strong> a dipyridylpyrazine between<br />

Ru(II) <strong>and</strong> Rh(III). All the other combinations <strong>of</strong> bridging lig<strong>and</strong>s be-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 237<br />

tween the three metal centers have been produced as well [111]. The ideal aim<br />

<strong>of</strong> the molecular design was to achieve complete photoinduced charge separation,<br />

with the positive hole on the iron donor <strong>and</strong> the electron on monoquat<br />

acceptor. Apparently, however, the redox properties <strong>of</strong> the molecular components<br />

were not quite ideal. Thus, the Ru(II)-based MLCT excited state is<br />

efficiently quenched. In the transient product obtained, however, the oxidized<br />

site is indeed the iron center but the reduced site seems to be a polypyridine<br />

lig<strong>and</strong> (in 23, either the bridging dipyridylpyrazine or the terpyridine lig<strong>and</strong>).<br />

In 23, this charge transfer state reverts to the ground state in 37 ns [111].<br />

The simple chromophore-quencher system 24 also contains a quaternarized<br />

electron acceptor attached to a Rh(III) polypyridine unit. This dyad<br />

was designed [112] to study intramolecular charge shift processes, using<br />

a photochemical inter/intramolecular reaction scheme <strong>of</strong> the type shown in<br />

Eqs. 15–19.<br />

The dyad, schematized as Rh(III)-DQ, undergoes Rh(III)-localized photoexcitation<br />

(Eq. 15). The excited state is then involved in reductive quenching<br />

(Eq. 16) by a suitable electron donor, 1,3,5-trimethoxybenzene, indicated<br />

as TMB. The reduced dyad originates from the quenching process with the extra<br />

electron on the metal complex moiety, i.e., on the thermodynamically less


238 M.T. Indelli et al.<br />

favored (by ca. 0.2 eV) site. Therefore, in competition with primary bimolecular<br />

charge recombination (Eq. 17), it is expected to undergo intramolecular<br />

electron transfer (charge shift process, Eq. 18) from the metal complex to DQ.<br />

The system will be finally converted back to ground-state reactants by secondary<br />

charge recombination (Eq. 19):<br />

Rh(III)-DQ + hν → ∗ Rh(III)-DQ (15)<br />

∗ +<br />

Rh(III)-DQ + TMB → Rh(II)-DQ + TMB (16)<br />

Rh(II)-DQ + TMB + → Rh(III)-DQ + TMB (17)<br />

Rh(II)-DQ → Rh(III)-DQ –<br />

(18)<br />

Rh(III)-DQ – +TMB + → Rh(III)-DQ + TMB . (19)<br />

The system performs indeed as predicted. The rate constants <strong>of</strong> all the processes<br />

in the above scheme have been experimentally determined, except for<br />

that <strong>of</strong> Eq. 17, inferred from experiments on appropriate model rhodium<br />

systems (without DQ pendant unit) [112]. In particular, the intramolecular<br />

charge shift process (Eq. 18) has been observed <strong>and</strong> time-resolved (k =<br />

3 × 107 s –1 ) by laser flash photolysis. It can be noticed that, from a formal<br />

viewpoint, the stepwise photoinduced electron transfer taking place in this<br />

dyad/quencher system is reminiscent <strong>of</strong> that <strong>of</strong> a triad for charge separation.<br />

A system that, despite the chemical <strong>and</strong> physical differences, bears a close<br />

similarity to charge separating triads, is the heterogeneous assembly depicted<br />

in Fig. 10 [113]. It is based on a Ru(II)-Rh(III) polypyridine dyad <strong>of</strong> the<br />

same type as 15, that has been functionalized with carboxyl groups at the<br />

Rh(III) units. This gives the dyad the capability to adsorb on nanocrystalline<br />

titanium dioxide <strong>and</strong> to act as a photosensitizer in Graetzel-type photoelec-<br />

Fig. 10 Schematic picture <strong>of</strong> a Rh(III)-Ru(II) dyad anchored on nanocrystalline TiO2 <strong>and</strong><br />

<strong>of</strong> its behavior as a heterotriad system (from [113])


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 239<br />

Fig. 11 Energy-level diagram <strong>and</strong> photophysical processes for the “heterotriad” <strong>of</strong> Fig. 10<br />

trochemical cells. When photocurrent action spectra are measured with this<br />

dyad sensitizer, it is seen that light absorption by the Ru(II) chromophore<br />

leads to electron injection into the semiconductor. Furthermore, a detailed<br />

analysis <strong>of</strong> the transient behavior <strong>of</strong> the system indicates that the dyad performs<br />

a stepwise charge injection process, i.e., intramolecular Ru → Rh<br />

electron transfer followed by electron injection from the Rh unit into the<br />

semiconductor (Fig. 10). The first process has comparable rates <strong>and</strong> efficiencies<br />

as for the free dyads in solution. The second step is 40% efficient, because<br />

<strong>of</strong> competing primary recombination (Fig. 11). When the final recombination<br />

between injected electrons <strong>and</strong> oxidized Ru(III) centers is studied, a remarkable<br />

slowing down is obtained relative to st<strong>and</strong>ard systems containing simple<br />

mononuclear sensitizers. Stepwise charge separation <strong>and</strong> slow recombination<br />

between remote sites are distinctive features <strong>of</strong> charge separating triads<br />

(Fig. 9b). Therefore, the system can be considered as a “heterotriad” with the<br />

TiO2 nanocrystal playing the role <strong>of</strong> the terminal electron acceptor [113].<br />

3.4<br />

Photoinduced Electron Collection<br />

Central to the problem <strong>of</strong> light energy conversion into fuels (e.g., solar water<br />

splitting or light-driven carbon dioxide reduction) is the concept that the<br />

fuel-generating reactions are multielectron processes [3]. Progress in this<br />

field is therefore related to the design <strong>of</strong> systems capable <strong>of</strong> performing<br />

photoinduced electron collection [114]. A supramolecular system for photoinduced<br />

electron collection (PEC) can be constructed by coupling components<br />

capable <strong>of</strong> causing photoinduced electron transfer processes with compo-


240 M.T. Indelli et al.<br />

Fig. 12 Block diagram representation <strong>of</strong> a photochemical molecular device for photoinduced<br />

electron collection <strong>and</strong> two-electron redox catalysis<br />

nents capable <strong>of</strong> storing electrons <strong>and</strong> using them in multielectron redox<br />

processes. A possible PEC scheme is shown in Fig. 12.<br />

In this scheme, P are electron transfer photosensitizers, C an electron store<br />

component, <strong>and</strong> BL rigid bridging lig<strong>and</strong>s. D is a sacrificial electron donor,<br />

<strong>and</strong> A2 is a two-electron reduced product (e.g., H2 starting from 2H + ). The<br />

key molecular component C must have the ability to store two electrons following<br />

photoinduced electron transfer from P, <strong>and</strong> to deliver them to the<br />

substrate A + in a low-activation two-electron process that leads to the desired<br />

product A2.<br />

Very few homogeneous systems for photoinduced electron collection<br />

(PEC) have been reported by now [115–122]. Trimetallic complexes incorporating<br />

polyazine bridging lig<strong>and</strong>s have been designed <strong>and</strong> studied by<br />

Brewer’s group for potential applications as PEC devices [123–126]. Most<br />

<strong>of</strong> these complexes have general formula [(bpy)2Ru(BL)]2MCl2 n+ with BL =<br />

2,3-bis(2-pyridyl)pyrazine (dpp), 2,3-bis(2-pyridyl)quinoxaline (dpq), or 2,3bis(2-pyridyl)benzoquinoxaline<br />

(dpb) <strong>and</strong> M = Ir(III) or Rh(III) [123, 125].<br />

The first functioning PEC system, [(bpy)2Ru(dpb)]2IrCl2 5+ ,wasreported<br />

in 1994, employing π systems <strong>of</strong> polyazine bridging lig<strong>and</strong>s to collect electrons<br />

[123]. Very recently, an analogous supramolecular trimetallic species<br />

has been reported where the central Ir-based moiety is replaced by a Rh(III)<br />

complex [127]. This new system, [(bpy)2Ru(dpp)]2RhCl2 5+ (25), was obtained<br />

coupling two Ru chromophoric units which play the role <strong>of</strong> P, through


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 241<br />

polyazine bridges (BL), to a central Rh core (C). In the presence <strong>of</strong> dimethylaniline<br />

(DMA) as sacrificial electron donor (D), 25 undergoes two-electron<br />

photoreduction at the rhodium center, producing the stable Rh(I) form<br />

[(bpy)2Ru(dpp)]2Rh 5+ via loss <strong>of</strong> two chlorides. This result is demonstrated<br />

by the observation that the spectroscopic changes associated with the photochemical<br />

reduction are identical with those seen in the electrochemical reduction<br />

experiments. On the basis <strong>of</strong> quenching experiments in the presence<br />

<strong>of</strong> different concentrations <strong>of</strong> DMA, the authors suggest that the monoreduced<br />

Rh(II) species can be formed by two alternative pathways following<br />

excitation <strong>of</strong> the peripheral Ru(II)-base chromophores: i) photoinduced electron<br />

transfer from the excited Ru(II)-based units followed by regeneration <strong>of</strong><br />

the Ru(II)-based chromophores by oxidation <strong>of</strong> the sacrificial donor or ii) bimolecular<br />

quenching <strong>of</strong> the excited Ru(II)-based units by the sacrificial donor<br />

followed by reduction <strong>of</strong> the central Rh(III). No clear indication is given as<br />

to the mechanism for the formation <strong>of</strong> the stable two-electron-reduced Rh(I)<br />

product from the Rh(II) intermediate. According to the authors, the ability<br />

<strong>of</strong> [(bpy)2Ru(dpp)]2RhCl2 5+ to undergo photoreduction at the rhodium<br />

center by multiple electrons <strong>and</strong> the fact that the photoreduced product<br />

[(bpy)2Ru(dpp)]2Rh 5+ is coordinatively unsaturated <strong>and</strong> thus available to interact<br />

with substrates are promising features in view <strong>of</strong> potential applications<br />

in light energy harvesting to produce fuels [127].<br />

4<br />

Rhodium Complexes as DNA Intercalators<br />

4.1<br />

Specific Binding to DNA <strong>and</strong> Photocleavage<br />

The development <strong>and</strong> the study <strong>of</strong> transition metal complexes able to bind<br />

selectively to DNA sites, emulating the behavior <strong>of</strong> the DNA-binding proteins,<br />

ranks among the most fascinating <strong>and</strong> challenging issues in the field <strong>of</strong><br />

current chemical <strong>and</strong> biological research [128–131]. This topic has been extensively<br />

investigated by Barton <strong>and</strong> coworkers, who devoted an impressive<br />

number <strong>of</strong> studies to the use <strong>of</strong> Rh(III) complexes with lig<strong>and</strong>s containing<br />

nitrogen donors as DNA binding agents [129, 130]. Most <strong>of</strong> these studies center<br />

around complexes <strong>of</strong> the lig<strong>and</strong> 9,10-phenanthrenequinonediimine (phi)<br />

(26) [130, 132–138].<br />

The phi Rh(III) complexes are indeed excellent DNA intercalators given<br />

the flat aromatic heterocyclic moiety <strong>of</strong> the phi lig<strong>and</strong> that deeply inserts <strong>and</strong><br />

stacks in between the DNA base pairs (binding affinity constants range from<br />

10 6 –10 9 M –1 ) [132, 139, 140]. The photophysical properties <strong>of</strong> phi Rh(III)<br />

complexes [50] have been discussed in Sect. 2.1. When bound to DNA, upon<br />

photoactivation with UV light, they are able to promote DNA str<strong>and</strong> cleav-


242 M.T. Indelli et al.<br />

age, thanks to the outst<strong>and</strong>ing oxidant properties <strong>of</strong> their excited states. The<br />

photocleavage ability <strong>of</strong>fers a strategy for the use <strong>of</strong> these rhodium complexes<br />

as DNA targets. The approach used by the Barton’s laboratory is the following:<br />

i) upon ultraviolet excitation, the excited state <strong>of</strong> DNA-bound rhodium<br />

complex promotes the scission <strong>of</strong> DNA sugar-phosphate backbone through<br />

oxidative degradation <strong>of</strong> the sugar moiety; ii) biochemical methods (e.g., gel<br />

electrophoresis) are used to determine where the str<strong>and</strong> scission occurred<br />

<strong>and</strong> therefore where, along the str<strong>and</strong>, the complex was bound. This method<br />

provides a powerful tool to mark specifically the sites <strong>of</strong> binding [129].<br />

A variety <strong>of</strong> articles focused on DNA photocleavage by phi complexes containing<br />

different lig<strong>and</strong>s in ancillary positions [130, 139–142]. The structural<br />

formula <strong>of</strong> the most extensively characterized complexes are reported below<br />

(27, 28, 29, 30).<br />

Irradiation with UV light <strong>of</strong> Rh(phen)2(phi) 3+ (28) <strong>and</strong>Rh(bpy)(phi)2 3+<br />

(30) intercalated in DNA leads to direct DNA str<strong>and</strong> scission with products


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 243<br />

consistent with 3 ′ -hydrogen abstraction from the deoxyribose sugar adjacent<br />

to the binding site [140, 141]. The photochemistry <strong>of</strong> these phi complexes<br />

intercalated in DNA has also been studied as a function <strong>of</strong> irradiation<br />

wavelength [143, 144]. Interestingly, the results showed that light <strong>of</strong> different<br />

wavelengths induces selectively different chemical reactions. In particular,<br />

the irradiation <strong>of</strong> the DNA-bound Rh(phi)2(L) 3+ complexes with UV light<br />

(λ = 313 nm) leads, as discussed above, to direct scission <strong>of</strong> the DNA-sugar<br />

backbone [141, 144]. If instead the complexes are excited with low-energy<br />

light (λ ≥ 365 nm) oxidative damage to the DNA bases is observed. The mechanism<br />

<strong>of</strong> these two photoprocesses is not discussed in great detail [130, 143].<br />

It is proposed, however, which direct DNA scission takes place by hydrogen<br />

abstraction from the sugar by the phi lig<strong>and</strong> radical <strong>of</strong> a lig<strong>and</strong>-to-metal<br />

charge transfer (LMCT) state [130]. On the other h<strong>and</strong>, the oxidative damage<br />

is attributed to the population <strong>of</strong> a powerful oxidizing excited state (ILCT [50]<br />

or LC [143]) with longer wavelength light.<br />

Photocleavage experiments have been used pr<strong>of</strong>itably for establishing how<br />

the phi complexes are associated to DNA [129, 130, 141]. Confirmation <strong>of</strong><br />

site selectivity <strong>and</strong> greater structural definition were obtained later from<br />

high-resolution NMR studies [134–138, 145]. The important result is that<br />

all the phi complexes bind DNA noncovalently through intercalation in the<br />

major groove where the phi lig<strong>and</strong> is inserted between the base pairs so<br />

as to maximize stacking interactions. More recently a full crystal structure<br />

<strong>of</strong> ∆– Rh ((R,R)-Me2trien)2(phi) 3+ ((R,R)-Me2trien = 2R,9R-diamino-4,7diazadecane)<br />

bound to a DNA octamer provided a direct evidence <strong>of</strong> the


244 M.T. Indelli et al.<br />

intercalation through the major groove [146]. A series <strong>of</strong> systematic NMR <strong>and</strong><br />

photocleavage studies clearly showed that the binding <strong>of</strong> complexes containing<br />

different ancillary lig<strong>and</strong>s occurs at a different specific DNA sequence.<br />

This site specificity results from both shape-selective steric interactions as<br />

well as stabilizing van der Waals <strong>and</strong> hydrogen bonding contacts. In particular,<br />

Rh(NH3)4phi 3+ <strong>and</strong> related amine complexes bind to d(TGGCCA)2<br />

duplex through hydrogen bonding between the ancillary amine lig<strong>and</strong>s <strong>and</strong><br />

DNA bases [130, 136]. Evidence for specific intercalation was found also for<br />

Rh(phen)2phi 3+ in the hexanucleotide d(GTCGAC)2 [134]. In this case Barton<br />

proposed that the site specificity was based upon shape-selection. Since<br />

thephenanthrolinelig<strong>and</strong>sprovidestericbulkabove<strong>and</strong>belowtheplane<strong>of</strong><br />

the phi lig<strong>and</strong>, the stacking occurs at sites which are more open in the major<br />

grove. The most striking example <strong>of</strong> site-specific recognition by shape<br />

selection with bulky ancillary lig<strong>and</strong>s was found for Rh(DPB)2phi 3+ (DPB<br />

=4,4 ′ -diphenylbpy) [140]. For all the complexes studied enantioselectivity<br />

favoring the intercalation <strong>of</strong> the ∆-isomer was observed [130, 147]. Further<br />

control <strong>of</strong> sequence specificity has been achieved by using derivatives <strong>of</strong><br />

Rh(phen)(phi)2 3+ complexes where the nonintercalating phenanthroline lig<strong>and</strong><br />

has been functionalized with pendant guanidinium group or with short<br />

oligopeptides [148, 149]. For metal-peptide complexes photocleavage experiments<br />

showed that the polypeptide chain is essential in directing the complex<br />

to a specific DNA sequence [149].<br />

Among the rhodium intercalators explored as probes <strong>of</strong> DNA structure,<br />

Barton selected the Rh(bpy)2chrysi 3+ (chrysi = 5,6-chrysenequinone diimine,<br />

31) complex as an ideal c<strong>and</strong>idate for mismatches recognition [150–152].<br />

The specific recognition is based on the size <strong>of</strong> the intercalating lig<strong>and</strong>: the<br />

chrysene ring system is too large to intercalate in normal B-form DNA but it<br />

can do so at destabilized mismatch sites. The authors point out that sterically<br />

dem<strong>and</strong>ing intercalators such as Rh(bpy)2chrysi 3+ may have application both<br />

in mutation detection systems <strong>and</strong> as mismatch-specific chemotherapeutic<br />

agents.<br />

Recently mixed-metal trimetallic complexes have been designed <strong>and</strong><br />

studied by Brewer to obtain supramolecular system capable <strong>of</strong> DNA photocleavage<br />

[153, 154]. These complexes <strong>of</strong> general formula [{(bpy)2M(dpp)}2<br />

RhCl2](PF6)5 with M = Ru(II) or Os(II) couple ruthenium or osmium chromophoric<br />

units to a central rhodium(III) core. When excited with visible light<br />

into their intense MLCT b<strong>and</strong>s, these complexes exhibit DNA photocleavage


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 245<br />

property. The authors discussed the role <strong>of</strong> the supramolecular architecture<br />

<strong>and</strong> in particular <strong>of</strong> the rhodium(III) unit on the photoreactivity<br />

4.2<br />

Rh(III) Complexes in DNA-Mediated Long-Range Electron Transfer<br />

DNA-mediated electron transfer has been an active <strong>and</strong> much debated topic<br />

<strong>of</strong> research [155]. Several articles have dealt with DNA-mediated photoinduced<br />

electron transfer reactions involving metal complexes as photoactive<br />

units [156]. In this context, <strong>of</strong> particular interest are the studies <strong>of</strong> Barton<br />

<strong>and</strong> coworkers that used rhodium(III) intercalator complexes not only as<br />

ground-state but also as excited-state electron acceptor in electron transfer<br />

(ET) reactions through the DNA [144].<br />

4.2.1<br />

Rh(III) Complexes as Acceptors in Electron Transfer Reactions<br />

In 1993, Murphy et al. [157]. reported the surprising result that an efficient<br />

<strong>and</strong> rapid photoinduced electron transfer occurs over a large separation distance<br />

(> 40 ˚A) between DNA metallointercalators that are covalently tethered<br />

to opposite 5 ′ -ends <strong>of</strong> a 15-base pair DNA duplex (Fig. 13).<br />

In this oligomeric assembly Ru(phen)2dppz 2+ (dppz = dipyridophenazine)<br />

plays the role <strong>of</strong> excited electron donor <strong>and</strong> the Rh(phi)2(phen) 3+ is the electron<br />

acceptor. Both donor <strong>and</strong> acceptor bind to DNA with high affinities<br />

(> 10 6 M –1 ) by intercalation through the dppz [158, 159] <strong>and</strong> phi lig<strong>and</strong>s, re-<br />

Fig. 13 A 15-base pair DNA duplex carrying covalently tethered Ru(II) <strong>and</strong> Rh(III) intercalators


246 M.T. Indelli et al.<br />

spectively. A clear advantage <strong>of</strong> the tethered Ru/DNA/Rh system is that both<br />

the donor <strong>and</strong> acceptor are covalently held in a well-defined fixed distance<br />

range. On the other h<strong>and</strong>, the report <strong>of</strong> Murphy et al. was limited by the<br />

exclusive use <strong>of</strong> steady-state emission spectroscopy. A lower limit for the photoinduced<br />

electron transfer rate (> 3 × 10 9 s –1 ) has been obtained measuring<br />

the quenching <strong>of</strong> the Ru(II) metal-to-lig<strong>and</strong> charge transfer (MLCT) emission<br />

by the tethered Rh(III) acceptor.<br />

In a subsequent investigation, untethered Ru/DNA/Rh <strong>and</strong> related systems<br />

were studied by Barton et al. [160] using ultrafast laser spectroscopy. The<br />

study was focused mainly on the system shown in Fig. 14 constituted by ∆-<br />

Ru(phen)2dppz 2+ as excited donor, ∆-Rh(phi)2bpy 3+ as acceptor intercalated<br />

in the calf thymus DNA with the aim to determine the rate <strong>of</strong> excited-state<br />

electron transfer (ket) that occurs from the lowest-lying MLCT state <strong>of</strong> the Ru<br />

donor, <strong>and</strong> the recombination electron transfer reaction (krec).<br />

Fig. 14 Photoinduced electron transfer processes taking place between Ru(phen)2dppz 2+<br />

<strong>and</strong> Rh(phi)2bpy 3+ DNA intercalators<br />

Efficient <strong>and</strong> rapid quenching <strong>of</strong> luminescence <strong>of</strong> the Ru complex in the<br />

presence <strong>of</strong> Rh complex, even at surprisingly low acceptor loading on the<br />

DNA duplex was observed. All the experimental observations were consistent<br />

with complete intercalation <strong>of</strong> the donor <strong>and</strong> acceptor in DNA. A comparative<br />

experiment employing Ru(NH3) 3+<br />

6 complex as electron acceptor, clearly<br />

indicates that much less efficient quenching is observed when the quencher<br />

is groove bound rather intercalated. To deepen the underst<strong>and</strong>ing <strong>of</strong> the<br />

mechanism <strong>of</strong> the electron transfer processes, the authors examined the photoinduced<br />

charge separation (ket) <strong>and</strong> recombination electron transfer (krec)<br />

reactions on the picosecond time scale by monitoring both the kinetics <strong>of</strong><br />

the emission decay <strong>and</strong> the kinetics <strong>of</strong> the recovery <strong>of</strong> ground state absorption<br />

<strong>of</strong> Ru(II) donor (Fig. 14). Time-correlated single photon counting failed<br />

to detect the lifetime <strong>of</strong> the excited state, clearly indicating that luminescent<br />

quenching by electron transfer proceeds faster (ket > 3 × 10 10 s –1 ) than the<br />

time resolution <strong>of</strong> the instrument (ca. 50 ps). Ultrafast transient absorption<br />

measurements, on the other h<strong>and</strong>, revealed bleaching <strong>of</strong> the MLCT b<strong>and</strong> <strong>of</strong><br />

the Ru(II) complex in a picosecond time scale, assigned by the authors to the


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 247<br />

Table 1 Rate constants for charge recombination following electron transfer from DNAbound,<br />

photoexcited donors to ∆-Rh(phi)2(bpy) 3+ a<br />

Donor DNA krec [10 9 s –1 ] –∆G ◦ [V]<br />

∆-Ru(phen)2(dppz) 2+ Calf thymus 9.2 1.66<br />

rac-Ru(bpy)2(dppz) 2+ Calf thymus 7.1 1.69<br />

∆-Ru(dmp)2(dppz) 2+ Calf thymus 11 1.59<br />

∆-Ru(phen)2(F2dppz) 2+ Calf thymus 7.7 1.68<br />

rac-Ru(phen)2(Me2dppz) 2+ Calf thymus 9.2 1.67<br />

∆-Os(phen)2(dppz) 2+ Calf thymus 11 1.21<br />

Λ-Ru(phen)2(dppz) 2+ Calf thymus 4.5<br />

∆-Ru(phen)2(dppz) 2+ Poly-d(AT) 7.4<br />

∆-Ru(phen)2(dppz) 2+ Poly-d(GC) 0.21<br />

a Based on [144]<br />

presence <strong>of</strong> Ru(III) oxidized donor. The rate constants for charge recombination<br />

process (krec) were obtained from the decay <strong>of</strong> this signal. The data<br />

for seven donor–acceptor pairs are given in Table 1. Within this series, the<br />

driving force (∆G ◦ ) is comparable but the donors vary with respect to intercalating<br />

lig<strong>and</strong>, ancillary lig<strong>and</strong>s, chirality, <strong>and</strong> metal center. Despite such<br />

a range <strong>of</strong> chemical properties, the rate observed is centered around 10 10 s –1 .<br />

A significant difference in rate is observed, however, when the absolute<br />

configuration <strong>of</strong> the donor is varied. For the right-h<strong>and</strong>ed ∆-Ru(phen)2dppz 2+<br />

the value is 2.5 times higher with respect the left-h<strong>and</strong>ed enantiomer indicating<br />

a deeper stacking <strong>of</strong> this complex into the double helix. This result,<br />

according to the authors, clearly suggests that the electron transfer process<br />

required the intervening aromatic base pairs. The notion <strong>of</strong> highly efficient<br />

ET through the stack <strong>of</strong> DNA base is also strongly supported by the finding<br />

that the largest change in electron transfer rate is observed when the sequence<br />

<strong>of</strong> the DNA bridge is changed: for the same donor <strong>and</strong> acceptor reactants the<br />

rate with poly d(AT) is 30 times higher than with poly d(GC). This is an importantresultthatindicatesthattheπ-stacked<br />

bases <strong>of</strong> the DNA provide an<br />

effective pathway for electron transfer reactions. However, the crucial point <strong>of</strong><br />

this study that involves an untethered Ru/DNA/Rh system concerns the distances<br />

over which fast ET occurs. The question is: do the donor <strong>and</strong> acceptor<br />

complexes contact each other, or does electron transfer occur at long range?<br />

Two models were considered by the authors to interpret the experimental results:<br />

i) a cooperative binding model with ET over short D–A distance <strong>and</strong><br />

ii) a r<strong>and</strong>om binding model with ET over long distances. On the basis <strong>of</strong><br />

DNA photocleavage experiments, the first hypothesis was reject in favor <strong>of</strong><br />

a rapid long range ET with a shallow distance dependence [160]. On the other<br />

h<strong>and</strong>, soon thereafter, Barbara [161] reinterpreted the experimental results on<br />

a quantitative basis using computational simulation procedures <strong>and</strong> demon-


248 M.T. Indelli et al.<br />

strated the failure <strong>of</strong> long-distance electron transfer model to account for the<br />

data. Concurrently, Tuite <strong>and</strong> coworkers [162] arrived at a similar conclusion<br />

for the ET quenching <strong>of</strong> Ru(phen)2dppz 2+ emission in a very similar untethered<br />

DNA/metallointercalator system. In summary, considerable controversy<br />

persists in the estimates <strong>of</strong> the distances over which fast ET may occur in such<br />

type <strong>of</strong> untethered systems [144].<br />

4.2.2<br />

Long Range Oxidative DNA Damage by Excited Rh(III) Complexes<br />

It is well known from a large variety <strong>of</strong> experimental studies <strong>and</strong> calculations<br />

that guanine (G) is the most easily oxidized <strong>of</strong> the nucleic acid<br />

bases [144, 163, 164]. Barton <strong>and</strong> coworkers have extensively exploited the<br />

ability <strong>of</strong> Rh(III)-phi complexes to induce oxidative damage specifically at<br />

the 5 ′ -G <strong>of</strong> the 5 ′ -GG-3 ′ doublets, when irradiated with low-energy light.<br />

A first investigation was carried out using a 15-base duplex (Rh-DNA) which<br />

possesses an end-tethered Rh(phi)2bpy 3+ complex in one str<strong>and</strong> <strong>and</strong> two<br />

5 ′ -GG-3 ′ sites in the complementary str<strong>and</strong>. (Fig. 15). The peculiarity <strong>of</strong><br />

this Rh-DNA assembly is that the rhodium complex is spatially separated<br />

in a well-defined manner from the potential sites <strong>of</strong> oxidation. Damage to<br />

DNA was demonstrated to occurred as a result <strong>of</strong> excitation <strong>of</strong> the intercalated<br />

rhodium complex, followed by long-range (30–40 ˚A) electron transfer<br />

through the DNA base pair stack [165]. The strategy used to analyze the<br />

mechanism is illustrated in Fig. 16.<br />

Fig. 15 A15-base duplex with an end-tethered Rh(phi)2bpy 3+ complex in one str<strong>and</strong> <strong>and</strong><br />

two 5 ′ -GG-3 ′ sites in the complementary str<strong>and</strong> [165]<br />

The Rh-DNA assembly was first irradiated at 313 nm to induce direct<br />

str<strong>and</strong> cleavage. This photocleavage step marks the site <strong>of</strong> intercalation, <strong>and</strong><br />

permits determination <strong>of</strong> the distance separating the rhodium complex from<br />

potential sites <strong>of</strong> damage. Rh-DNA samples were then irradiated with low<br />

energy light at 365 nm, treated with hot piperidine, which promotes str<strong>and</strong><br />

cleavage at the damaged sites, <strong>and</strong> examined by gel electrophoresis. This<br />

treatment reveals the position <strong>and</strong> yield <strong>of</strong> damage. The results clearly in-


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 249<br />

Fig. 16 Use <strong>of</strong> the tethered Rh(III) complex to (i) mark the site <strong>of</strong> intercalation by direct<br />

str<strong>and</strong> cleavage (313-nm irradiation) <strong>and</strong> (ii) promote damage via long-range electron<br />

transfer (365-nm irradiation) [165]<br />

dicated that both the proximal <strong>and</strong> distal 5 ′ -GG-3 ′ doublets were equally<br />

damaged <strong>and</strong> the reaction was intraduplex. Two possible mechanisms for this<br />

process were discussed: i) concerted long-range electron transfer; <strong>and</strong> ii) oxidation<br />

<strong>of</strong> a base near the intercalated Rh acceptor followed by hole migration<br />

to the two GG sites. Sensitivity <strong>of</strong> the reaction to the intervening base pair<br />

stack was also observed. In subsequent studies, oxidation has been reported<br />

at sites that are up to 200 ˚A away from the site <strong>of</strong> intercalation <strong>of</strong> the photoactive<br />

rhodium complex [166].<br />

The photooxidant properties <strong>of</strong> the phi rhodium (III) complexes have also<br />

been used to repair thymine dimers [167, 168], the most common photo-<br />

Fig. 17 DNA duplex containing a thymine dimer with tethered Rh(III) complex for photoinduced<br />

repair studies [167, 168]


250 M.T. Indelli et al.<br />

chemical lesion in DNA. Investigations <strong>of</strong> photoinitiated repair <strong>of</strong> duplexes<br />

containing a single thymine dimer lesion were carried out with visible light<br />

(400 nm) using both nontethered <strong>and</strong> tethered complexes (Fig. 17).<br />

The quantum yield for photorepair with a Rh(III)-tethered complex is substantially<br />

(about ca. 30 fold) reduced compared to the noncovalently bound<br />

complex. Since the repair efficiency does not appear to be very sensitive to the<br />

distance between intercalated rhodium complex <strong>and</strong> the thymine dimer, the<br />

authors suggest that the observed disparity likely results from differences in<br />

π-stacking. In addition, evidences that the repair efficiency diminished with<br />

disruption <strong>of</strong> the intervening π-stack confirm that the DNA helix mediates<br />

this long-range oxidative repair reaction.<br />

5<br />

Conclusion<br />

A large number <strong>of</strong> rhodium(III) polypyridine complexes <strong>and</strong> their cyclometalated<br />

analogues have been investigated from the viewpoint <strong>of</strong> photochemistry,<br />

photophysics <strong>and</strong> <strong>of</strong> their possible applications.<br />

As mononuclear species, Rh(III) polypyridine complexes display interesting<br />

photophysical properties, with lowest excited states <strong>of</strong> LC type for tris<br />

bis-chelated species, <strong>and</strong> increasing role <strong>of</strong> MC states for mixed-lig<strong>and</strong> halopolypyridine<br />

species. In Rh(III) cyclometalated complexes, the covalent character<br />

<strong>of</strong> the C – Rh bonds makes the excited state classification less clearcut,<br />

with strong mixing <strong>of</strong> LC, MLCT, <strong>and</strong> LLCT character.<br />

Many polynuclear <strong>and</strong> supramolecular systems containing Rh(III) polypyridine<br />

<strong>and</strong> related units have been synthesized <strong>and</strong> studied, taking advantage<br />

<strong>of</strong> the favorable properties <strong>of</strong> these units as good electron acceptors <strong>and</strong><br />

strong photo-oxidants. In particular, Ru(II)-Rh(IIII) dyads have been actively<br />

investigated for the study <strong>of</strong> photoinduced electron transfer, with specific<br />

interest in driving force, distance, <strong>and</strong> bridging lig<strong>and</strong> effects. A limited number<br />

<strong>of</strong> supramolecular systems <strong>of</strong> higher nuclearity have also been produced.<br />

Among these, <strong>of</strong> particular interest are trinuclear species containing rhodium<br />

dihalo polypyridine units, which can act as two-electron storage components<br />

thanks to their Rh(III)/Rh(I) redox behavior.<br />

Finally, a large amount <strong>of</strong> work has been devoted to the use <strong>of</strong> Rh(III)<br />

polypyridine complexes as intercalators for DNA. In this role, they have<br />

shown a very versatile behavior, being used for direct str<strong>and</strong> photocleavage<br />

marking the site <strong>of</strong> intercalation, to induce long-distance photochemical<br />

damage or dimer repair, or to act as electron acceptors in long-range electron<br />

transfer processes.<br />

Acknowledgements Financial support from MUR (PRIN 2006) is gratefully acknowledged.


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 251<br />

References<br />

1. Carassiti V, Balzani V (1970) <strong>Photochemistry</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>. Academic<br />

Press, New York<br />

2. Kalyanasundaram K (1992) <strong>Photochemistry</strong> <strong>of</strong> Polypyridine <strong>and</strong> Porphyrin Complexes.<br />

Academic Press, London<br />

3. Balzani V, Juris A, Venturi M, Campagna S, Serroni S (1996) Chem Rev 96:759<br />

4. Hannon MJ (1997) Coord Chem Rev 162:477<br />

5. Sc<strong>and</strong>ola F, Chiorboli C, Indelli MT, Rampi MA (2001) Covalently linked systems<br />

containing metal complexes. In: Balzani V (ed) Electron Transfer in Chemistry,<br />

Vol III, Chap 2.3. Wiley-VCH, Weinheim, p 337<br />

6. Chiorboli C, Indelli MT, Sc<strong>and</strong>ola F (2005) Top Curr Chem 257:63<br />

7. Ford PC (1983) J Chem Ed 60:829<br />

8. Ford PC, Wink D, DiBenedetto J (1983) Prog Inorg Chem 40:213<br />

9. Miller DB, Miller PK, Kane-Maguire NAP (1983) Inorg Chem 22:3831<br />

10. Frink ME, Ford PC, Skibsted LH (1984) Acta Chem Sc<strong>and</strong> A38:795<br />

11. Skibsted LH, Hancock MP, Magde D, Sexton DA (1984) Inorg Chem 23:3735<br />

12. Skisbsted LH (1985) Coord Chem Rev 64:343<br />

13. Kane-Maguire NAP, Wallace KC, Cobranchi DP, Derrick JM, Speece DG (1986) Inorg<br />

Chem 25:2101<br />

14. Skibsted LH, Hancock MP, Magde D, Sexton DA (1987) Inorg Chem 26:1708<br />

15. McClure LJ, Ford PC (1992) J Phys Chem 96:6640<br />

16. Carlos RM, Frink ME, Tfouni E, Ford PC (1992) Inorg Chim Acta 193:159<br />

17. Islam K, Ikeda AN, Nozaki KT, Ohno T (1998) J Chem Phys 109:4900<br />

18. Forster LS, Rund JV (2003) Inorg Chem Commun 6:78<br />

19. Coppens P, Gerlits O, Vorontsov II, Kovalevsky AY, Chen Y-S, Graber T, Novozhilova<br />

IV (2004) Chem Commun, p 2144<br />

20. Miskowski VM, Rice SF, Gray HB, Milder SJ (1993) J Phys Chem 97:4277<br />

21. Bradley PM, Bursten BE, Turro C (2001) Inorg Chem 40:1376<br />

22. Cotton FA, Lin C, Murillo CA (2001) Acc Chem Res 34:759<br />

23. Lo Schiavo S, Serroni S, Puntoriero F, Tresoldi G, Piraino P (2002) Eur J Inorg Chem<br />

79<br />

24. Cooke MW, Hanan GS, Loiseau F, Campagna S, Watanabe M, Tanaka Y (2005) Angew<br />

Chem Int Ed 44:4881<br />

25. Heyduk AF, Macintosh AM, Nocera DG (1999) J Am Chem Soc 121:5023<br />

26. Heyduk AF, Nocera DG (2001) Science 293:1639<br />

27. Gray TG, Nocera DG (2005) Chem Commun, p 1540<br />

28. Hillis JE, De Armond MK (1971) J Lumin 4:273<br />

29. Carstens DHW, Crosby GA (1970) J Mol Spectrosc 34:113<br />

30. Crosby GA, Elfring WH Jr (1976) J Phys Chem 80:2206<br />

31. De Armond MK, Hillis JE (1971) J Chem Phys 54:2247<br />

32. Humbs W, Yersin H (1996) Inorg Chem 35:2220<br />

33. Humbs W, Strasser J, Yersin H (1997) J Lumin 72–74:677<br />

34. Indelli MT, Carioli A, Sc<strong>and</strong>ola F (1984) J Phys Chem 88:2685<br />

35. Bolletta F, Rossi A, Barigelletti F, Dellonte S, Balzani V (1981) Gazz Chim Ital 111:155<br />

36. Ohno T, Kato S (1984) Bull Chem Soc Jpn 57:3391<br />

37. Nishizawa M, Suzuki TM, Sprouse S, Watts RJ, Ford PC (1984) Inorg Chem 23:1837<br />

38. Indelli MT, Sc<strong>and</strong>ola F (1990) Inorg Chem 29:3056<br />

39. Brozik JA, Crosby GA (2005) Coord Chem Rev 249:1310<br />

40. Broomhead JA, Grumley W (1968) Chem Commun, p 1211


252 M.T. Indelli et al.<br />

41. Muir MM, Huang WL (1973) Inorg Chem 12:1831<br />

42. Loganathan D, Rodriguez JH, Morrison H (2003) J Am Chem Soc 125:5640<br />

43. Menon EL, Perera R, Navarro M, Kuhn RJ, Morrison H (2004) Inorg Chem 43:5373–<br />

5381<br />

44. Huang WL, Lee JR, Shi SY, Tsai CY (2003) Trans Met Chem 28:381<br />

45. Tseng MC, Li FK, Huang JH, Su WL, Wang SP, Huang WL (2006) Inorg Chim Acta<br />

359:401<br />

46. Calvert JM, Caspar JV, Binstead RA, Westmorel<strong>and</strong> TD, Meyer TJ (1982) J Am Chem<br />

Soc 104:6620<br />

47. Frink ME, Sprouse SD, Goodwin HA, Watts RJ, Ford PC (1988) Inorg Chem 27:1283<br />

48. Watts RJ, Van Houten J (1978) J Am Chem Soc 100:1718<br />

49. Westra J, Glasbeek M (1992) J Lumin 53:92<br />

50. Turro C, Evenzahav A, Bossmann SH, Barton JK, Turro NJ (1996) Inorg Chim Acta<br />

243:101<br />

51. Burke HM, Gallagher JF, Indelli MT, Vos JG (2004) Inorg Chim Acta 357:2989<br />

52. Shinozaki K, Takahashi N (1996) Inorg Chem 35:3917<br />

53. Fujita E, Brunschwig BS, Creutz C, Muckerman JT, Sutin N, Szalda D, van Eldik R<br />

(2006) Inorg Chem 45:1595<br />

54. Martin B, McWhinnie WR, Waind GM (1961) J Inorg Nucl Chem 23:207<br />

55. Chou M, Creutz C, Mahajan D, Sutin N, Zipp AP (1982) Inorg Chem 21:3989<br />

56. Mulazzani QG, Venturi M, H<strong>of</strong>fman MZ (1982) J Phys Chem 86:242<br />

57. Schwarz HA, Creutz C (1983) Inorg Chem 22:707<br />

58. Yan SG, Brunschwig BS, Creutz C, Fujita E, Sutin N (1998) J Am Chem Soc 120:10553<br />

59. Maestri M, Balzani V, Deuschel-Cornioley C, von Zelewsky A (1992) Adv Photochem<br />

17:1<br />

60. Colombo MG, Brunold TC, Riedener T, Guedel HU, Fortsch M, Buergi H-B (1994)<br />

Inorg Chem 33:545<br />

61. Constable EC, Leese TA (1990) Polyhedron 9:1613<br />

62. Maeder U, von Zelewsky A, Stoeckli-Evans H (1992) Helv Chim Acta 75:1320<br />

63. Maestri M, S<strong>and</strong>rini D, Balzani V, Mader U, von Zelewsky A (1987) Inorg Chem<br />

26:1323<br />

64. Ohsawa Y, Sprouse S, King KA, De Armond MK, Hanck KW, Watts RJ (1987) J Phys<br />

Chem 91:1047<br />

65. Zilian A, Maeder U, von Zelewski A, Guedel HU (1989) J Am Chem Soc 111:3855<br />

66. Zilian A, Guedel HU (1991) Coord Chem Rev 111:33<br />

67. Ghizdavu L, Lentzen O, Schumm S, Brodkorb A, Moucheron C, Kirsch-De Mesmaeker<br />

A (2003) Inorg Chem 42:1935<br />

68. Hay PJ (2002) J Phys Chem A 106:1634<br />

69. Polson M, Ravaglia M, Fracasso S, Garavelli M, Sc<strong>and</strong>ola F (2005) Inorg Chem<br />

44:1282<br />

70. Barigelletti F, S<strong>and</strong>rini D, Maestri M, Balzani V, von Zelewsky A, Chassot L, Jolliet P,<br />

Maeder U (1988) Inorg Chem 27:3644<br />

71. Lo KK-W, Li C-K, Lau K-W, Zhu N (2003) Dalton Trans 24:4682<br />

72. S<strong>and</strong>rini D, Maestri M, Balzani V, Mader U, von Zelewsky A (1988) Inorg Chem<br />

27:2640<br />

73. Calogero G, Giuffrida G, Serroni S, Ricevuto V, Campagna S (1995) Inorg Chem<br />

34:541<br />

74. Didier P, Ortmans I, Kirsch-De Mesmaeker A, Watts RJ (1993) Inorg Chem 32:5239<br />

75. Van Diemen JH, Hage R, Haasnoot JG, Lempers HEB, Reedijk J, Vos JG, De Cola L,<br />

Barigelletti F, Balzani V (1992) Inorg Chem 31:3518


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 253<br />

76. Ortmans I, Didier P, Kirsch-De Mesmaeker A (1995) Inorg Chem 34:3695<br />

77. Petitjean A, Puntoriero F, Campagna S, Juris A, Lehn J-M, (2006) Eur J Inorg Chem<br />

3878<br />

78. Kew G, Hanck K, DeArmond MK (1975) J Phys Chem 79:1828<br />

79. Ballardini R, Varani G, Balzani V (1980) J Am Chem Soc 102:1719<br />

80. Indelli MT, Ballardini R, Sc<strong>and</strong>ola F (1984) J Phys Chem 88:2547<br />

81. Lehn J-M, Sauvage J-P (1977) Nouv J Chim 1:449<br />

82. Chan S-F, Chou M, Creutz C, Matsubara T, Sutin N (1981) J Am Chem Soc 103:369<br />

83. Creutz C, Keller AD, Sutin N, Zipp A P (1982) J Am Chem Soc 104:3618<br />

84. Indelli MT, Bignozzi CA, Harriman A, Schoonover JR, Sc<strong>and</strong>ola F (1994) J Am Chem<br />

Soc 116:3768<br />

85. Yoshimura A, Nozaki K, Ohno T (1997) Coord Chem Rev 159:375<br />

86. Furue M, Hirata M, Kinoshita S, Kushida T, Kamachi M (1990) Chem Lett 2065<br />

87. Collin J-P, Gavina P, Heitz V, Sauvage J-P (1998) Eur J Inorg Chem :1<br />

88. Welter S, Salluce N, Belser P, Groeneveld M, De Cola L (2005) Coord Chem Rev<br />

249:1360<br />

89. Welter S, Lafolet F, Cecchetto E, Vergeer F, De Cola L (2005) Chem Phys Chem 6:2417<br />

90. Barigelletti F, Flamigni I (2000) Chem Soc Rev 29:1<br />

91. Indelli MT, Sc<strong>and</strong>ola F, Flamigni L, Collin J-P, Sauvage J-P, Sour A (1997) Inorg<br />

Chem 36:4247<br />

92. Indelli MT, Chiorboli C, Flamigni L, De Cola L, Sc<strong>and</strong>ola F (2007) Inorg Chem<br />

(in press)<br />

93. McConnell HM (1961) J Chem Phys 35:508<br />

94. Newton MD (1991) Chem Rev 91:767<br />

95. Paddon-Row MN (2001) Covalently linked systems based on organic components.<br />

In: Balzani V (ed) Electron Transfer in Chemistry, Vol III, Chap 2.1. Wiley-VCH,<br />

Weinheim, p 179<br />

96. Helms A, Heiler D, McLendon G (1992) J Am Chem Soc 114:6227<br />

97. Weiss EA, Ahrens MJ, Sinks LE, Gusev AV, Ratner MA, Wasielewsi MR (2004) J Am<br />

Chem Soc 126:5577<br />

98. Wold DJ, Haag R, Rampi MA, Frisbie CD (2002) J Phys Chem B 106:2814<br />

99. Toutounji MM, Ratner MA (2000) J Phys Chem A 104:8566<br />

100. Tour JM, Lamba JJS (1993) J Am Chem Soc 115:4935<br />

101. Tsuzuki K, Tanabe J (1991) J Phys Chem 95:139<br />

102. Kalyanasundaram K, Graetzel M, Nazeeruddin MK (1992) J Phys Chem 96:5865<br />

103. Lee J-D, Vrana LM, Bullock ER, Brewer KJ (1998) Inorg Chem 37:3575<br />

104. Kew G, DeArmond K, Hanck K (1974) J Phys Chem 78:727<br />

105. Collin J-P, Harriman A, Heitz V, Odobel F, Sauvage J-P (1994) J Am Chem Soc<br />

116:5679<br />

106. Wasielewski MR (1992) Chem Rev 92:435<br />

107. Gust D, Moore TA, Moore AL (2001) Covalently linked systems containing porphyrin<br />

units. In: Balzani V (ed) Electron Transfer in Chemistry, Vol III, Chap 2.2. Wiley-<br />

VCH, Weinheim, p 273<br />

108. Larson SL, Elliott CM, Kelley DF (1995) J Phys Chem 99:6530<br />

109. Huynh MHV, Dattelbaum DM, Meyer TJ (2005) Coord Chem Rev 249:457<br />

110. Baran<strong>of</strong>f E, Collin J-P, Flamigni L, Sauvage J-P (2004) Chem Soc Rev 33:147<br />

111. Ronco SE, Thompson DW, Gahan SL, Petersen JD (1998) Inorg Chem 37:2020<br />

112. Indelli MT, Polo E, Bignozzi CA, Sc<strong>and</strong>ola F (1991) J Phys Chem 95:3889<br />

113. Kleverlaan CJ, Indelli MT, Bignozzi CA, Pavanin LA, Sc<strong>and</strong>ola F, Hasselman GM,<br />

Meyer GJ (2000) J Am Chem Soc 122:2840


254 M.T. Indelli et al.<br />

114. Balzani V, Sc<strong>and</strong>ola F (1991) Supramolecular <strong>Photochemistry</strong>. Horwood, Chichester<br />

115. Konduri R, Ye H, MacDonnell FM, Serroni S, Campagna S, Rajeshwar K (2002)<br />

Angew Chem Int Ed 41:3185<br />

116. Chiorboli C, Fracasso S, Ravaglia M, Sc<strong>and</strong>ola F, Campagna S, Wouters K, Konduri R,<br />

MacDonnell F (2005) Inorg Chem 44:8368<br />

117. Heyduk AF, Nocera DG (2001) Science 293:1639<br />

118. Esswein A, Veige A, Nocera D (2005) J Am Chem Soc 127:16641<br />

119. Chang CC, Pfennig B, Bocarsly AB (2000) Coord Chem Rev 208:33<br />

120. Watson DF, Tan HS, Schreiber E, Mordas CJ, Bocarsly AB (2004) J Phys Chem A<br />

108:3261<br />

121. Rau S, Schfer B, Gleich D, Anders E, Rudolph M, Friedrich M, Gorls H, Henry W,<br />

Vos JG (2006) Angew Chem Int Ed 45:6215<br />

122. Ozawa H, Haga M, Sakai K (2006) J Am Chem Soc 128:4926<br />

123. Molnar SM, Nallas G, Bridgewater JS, Brewer KJ (1994) J Am Chem Soc 116:5206<br />

124. Nallas GNA, Jones SW, Brewer KJ (1996) Inorg Chem 35:6974<br />

125. Molnar SM, Jensen GE, Vogler LM, Jones SW, Laverman L, Bridgewater JS, Richter<br />

MM, Brewer KJ (1994) J Photochem Photobiol A Chem 80:315<br />

126. Swavey S, Brewer KJ (2002) Inorg Chem 41:4044<br />

127. Elvington M, Brewer KJ (2006) Inorg Chem 45:5242<br />

128. Barton JK (1986) Science 233:727<br />

129. Pyle AM, Barton JK (1990) In: Lippard SJ (ed) Progress in Inorganic Chemistry:<br />

Bioinorganic Chemistry, Vol 38. Wiley, New York, p 413<br />

130. Erkkila KE, Odom DT, Barton JK (1999) Chem Rev 99:2777<br />

131. Jennette KW, Lippard SJ, Vassiliades GA, Bauer WR (1974) Proc Natl Acad Sci USA<br />

71:3839<br />

132. Crotz AH, Hudson BP, Barton JK (1993) J Am Chem Soc 115:12577<br />

133. Pyle AM, Chiang MY, Barton JK (1990) Inorg Chem 29:4487<br />

134. David SS, Barton JK (1993) J Am Chem Soc 115:2984<br />

135. Crotz AH, Kuo LY, Barton JK (1993) Inorg Chem 32:5963<br />

136. Collins J-C, Shield JK, Barton JK (1994) J Am Chem Soc 116:9840<br />

137. Crotz AH, Barton JK (1994) Inorg Chem 33:1940<br />

138. Hudson BP, Barton JK (1998) J Am Chem Soc 120:6877<br />

139. Crotz AH, Kuo LY, Shields TP, Barton JK (1993) J Am Chem Soc 115:3877<br />

140. Sitlani A, Dupureur C, Barton JK (1993) J Am Chem Soc 115:12589<br />

141. Sitlani A, Long EC, Pyle AM, Barton JK (1992) J Am Chem Soc 114:2303<br />

142. Kisko J, Barton JK (2000) Inorg Chem 39:4942<br />

143. Turro C, Hall DB, Chen W, Zuilh<strong>of</strong> H, Barton JK, Turro NJ (1998) J Phys Chem A<br />

102:5708<br />

144. Holmlin RE, D<strong>and</strong>liker PJ, Barton KJ (1997) Angew Chem Int Ed 36:2714<br />

145. Hudson BP, Dupureur CM, Barton JK (1995) J Am Chem Soc 117:9379<br />

146. Kielkopf CL, Erkkila KE, Hudson BP, Barton JK, Rees DC (2000) Nat Struct Biol 7:117<br />

147. Pyle AM, Morii T, Barton JK (1990) J Am Chem Soc 112:9432<br />

148. Sardesai NJ, Zimmermann, Barton JK (1994) J Am Chem Soc 116:7502<br />

149. Terbrueggen RH, Johann T W, Barton JK (1998) Inorg Chem 37:6874<br />

150. Jackson BA, Barton JK (1997) J Am Chem Soc 119:12986<br />

151. Jackson BA, Alekseyev VY, Barton JK (1999) Biochemistry 38:4655<br />

152. Kisko JL, Barton JK (2000) Inorg Chem 39:4942<br />

153. Holder AA, Swavey A, Brewer KJ (2004) Inorg Chem 43:303<br />

154. Swavey S, Brewer KJ (2002) Inorg Chem 41:6196


<strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Rhodium 255<br />

155. Lewis FD (2001) In: Balzani V (ed) Electron Transfer in Chemistry, Vol III, Chap 1.5.<br />

Wiley/VCH, Weinheim, p 105<br />

156. Stemp EDA, Barton JK (1996) In: Sigel A, Sigel H (eds) Metal Ions in Biological<br />

systems, Vol 33. Marcel Dekker, New York, p 325<br />

157. Murphy CJ, Arkin MR, Jenkins Y, Ghatlia ND, Bossmann S, Turro NJ, Barton JK<br />

(1993) Science 262:1025<br />

158. Friedman AE, Chambron JC, Sauvage JP, Turro NJ, Barton JK (1990) J Am Chem Soc<br />

112:4960<br />

159. Hort C, Lincoln P, Norden B (1993) J Am Chem Soc 115:3448<br />

160. Arkin MR, Stemp EDA, Holmlin RE, Barton JK, Hormann A, Olson EJC, Barbara PF<br />

(1996) Science 273:475<br />

161. Olson EJC, Hu D, Hormann A, Barbara PF (1997) J Phys Chem B 101:299<br />

162. Lincoln P, Tuite E, Norden B (1997) J Am Chem Soc 119:1454<br />

163. Saito I, Takayama M, Sugiyama H, Nakatani K (1995) J Am Chem Soc 117:6406<br />

164. Kittler L (1980) J Electroanalyt Chem 116:503<br />

165. Hall DB, Holmlin RE, Barton JK (1996) Nature 382:731<br />

166. Nunez ME, Hall DB, Barton JK (1999) Chem Biol 6:85<br />

167. D<strong>and</strong>liker PJ, Holmlin RE, Barton JK (1997) Science 275:1465<br />

168. D<strong>and</strong>liker PJ, Nunez ME, Barton JK (1998) Biochemistry 37:6491


Author Index Volumes 251–280<br />

Author Index Vols. 26–50 see Vol. 50<br />

Author Index Vols. 51–100 see Vol. 100<br />

Author Index Vols. 101–150 see Vol. 150<br />

Author Index Vols. 151–200 see Vol. 200<br />

Author Index Vols. 201–250 see Vol. 250<br />

Thevolumenumbersareprintedinitalics<br />

Accorsi G, see Armaroli N (2007) 280: 69–115<br />

Ajayaghosh A, George SJ, Schenning APHJ (2005) Hydrogen-Bonded Assemblies <strong>of</strong> Dyes<br />

<strong>and</strong> Extended π-Conjugated Systems. 258: 83–118<br />

Akai S, Kita Y (2007) Recent Advances in Pummerer Reactions. 274: 35–76<br />

Albert M, Fensterbank L, Lacôte E, Malacria M (2006) T<strong>and</strong>em Radical Reactions. 264:1–62<br />

Alberto R (2005) New Organometallic Technetium Complexes for Radiopharmaceutical<br />

Imaging. 252:1–44<br />

Alegret S, see Pividori MI (2005) 260:1–36<br />

Alfaro JA, see Schuman B (2007) 272: 217–258<br />

Amabilino DB, Veciana J (2006) Supramolecular Chiral Functional Materials. 265: 253–302<br />

Anderson CJ, see Li WP (2005) 252: 179–192<br />

Anslyn EV, see Collins BE (2007) 277: 181–218<br />

Anslyn EV, see Houk RJT (2005) 255: 199–229<br />

Appukkuttan P, Van der Eycken E (2006) Microwave-Assisted Natural Product Chemistry.<br />

266: 1–47<br />

Araki K, Yoshikawa I (2005) Nucleobase-Containing Gelators. 256: 133–165<br />

Armaroli N, Accorsi G, Cardinali Fç, Listorti A (2007) <strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong><br />

<strong>Coordination</strong> <strong>Compounds</strong>: Copper. 280: 69–115<br />

Armitage BA (2005) Cyanine Dye–DNA Interactions: Intercalation, Groove Binding <strong>and</strong><br />

Aggregation. 253: 55–76<br />

Arya DP (2005) Aminoglycoside–Nucleic Acid Interactions: The Case for Neomycin. 253:<br />

149–178<br />

Bailly C, see Dias N (2005) 253: 89–108<br />

Balaban TS, Tamiaki H, Holzwarth AR (2005) Chlorins Programmed for Self-Assembly. 258:<br />

1–38<br />

Baltzer L (2007) Polypeptide Conjugate Binders for Protein Recognition. 277: 89–106<br />

Balzani V, Bergamini G, Campagna S, Puntoriero F (2007) <strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong><br />

<strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Overview <strong>and</strong> General Concepts. 280:1–36<br />

Balzani V, Credi A, Ferrer B, Silvi S, Venturi M (2005) Artificial Molecular Motors <strong>and</strong><br />

Machines: Design Principles <strong>and</strong> Prototype Systems. 262:1–27<br />

Balzani V, see Campagna S (2007) 280: 117–214<br />

Barbieri CM, see Pilch DS (2005) 253: 179–204<br />

Barchuk A, see Daasbjerg K (2006) 263: 39–70<br />

Bargon J, see Kuhn LT (2007) 276: 25–68<br />

Bargon J, see Kuhn LT (2007) 276: 125–154<br />

Bayly SR, see Beer PD (2005) 255: 125–162


258 Author Index Volumes 251–280<br />

Beck-Sickinger AG, see Haack M (2007) 278: 243–288<br />

Beer PD, Bayly SR (2005) Anion Sensing by Metal-Based Receptors. 255: 125–162<br />

Bergamini G, see Balzani V (2007) 280:1–36<br />

Bergamini G, see Campagna S (2007) 280: 117–214<br />

Bertini L, Bruschi M, de Gioia L, Fantucci P, Greco C, Zampella G (2007) Quantum Chemical<br />

Investigations <strong>of</strong> Reaction Paths <strong>of</strong> Metalloenzymes <strong>and</strong> Biomimetic Models – The<br />

Hydrogenase Example. 268:1–46<br />

Bier FF, see Heise C (2005) 261:1–25<br />

Blum LJ, see Marquette CA (2005) 261: 113–129<br />

Boiteau L, see Pascal R (2005) 259: 69–122<br />

Bolhuis PG, see Dellago C (2007) 268: 291–317<br />

Borovkov VV, Inoue Y (2006) Supramolecular Chirogenesis in Host–Guest Systems Containing<br />

Porphyrinoids. 265: 89–146<br />

Boschi A, Duatti A, Uccelli L (2005) Development <strong>of</strong> Technetium-99m <strong>and</strong> Rhenium-188 Radiopharmaceuticals<br />

Containing a Terminal Metal–Nitrido Multiple Bond for Diagnosis<br />

<strong>and</strong> Therapy. 252: 85–115<br />

Braga D, D’Addario D, Giaffreda SL, Maini L, Polito M, Grepioni F (2005) Intra-Solid <strong>and</strong><br />

Inter-Solid Reactions <strong>of</strong> Molecular Crystals: a Green Route to Crystal Engineering. 254:<br />

71–94<br />

Bräse S, see Jung N (2007) 278:1–88<br />

Braverman S, Cherkinsky M (2007) [2,3]Sigmatropic Rearrangements <strong>of</strong> Propargylic <strong>and</strong><br />

Allenic Systems. 275: 67–101<br />

Brebion F, see Crich D (2006) 263:1–38<br />

Breinbauer R, see Mentel M (2007) 278: 209–241<br />

Breit B (2007) Recent Advances in Alkene Hydr<strong>of</strong>ormylation. 279: 139–172<br />

Brizard A, Oda R, Huc I (2005) Chirality Effects in Self-assembled Fibrillar Networks. 256:<br />

167–218<br />

Broene RD (2007) Reductive Coupling <strong>of</strong> Unactivated Alkenes <strong>and</strong> Alkynes. 279: 209–248<br />

Bromfield K, see Ljungdahl N (2007) 278: 89–134<br />

Bruce IJ, see del Campo A (2005) 260: 77–111<br />

Bruschi M, see Bertini L (2007) 268:1–46<br />

Bur SK (2007) 1,3-Sulfur Shifts: Mechanism <strong>and</strong> Synthetic Utility. 274: 125–171<br />

Campagna S, Puntoriero F, Nastasi F, Bergamini G, Balzani V (2007) <strong>Photochemistry</strong> <strong>and</strong><br />

<strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>: Ruthenium. 280: 117–214<br />

Campagna S, see Balzani V (2007) 280:1–36<br />

del Campo A, Bruce IJ (2005) Substrate Patterning <strong>and</strong> Activation Strategies for DNA Chip<br />

Fabrication. 260: 77–111<br />

Cardinali F, see Armaroli N (2007) 280: 69–115<br />

Carney CK, Harry SR, Sewell SL, Wright DW (2007) Detoxification Biominerals. 270: 155–185<br />

Castagner B, Seeberger PH (2007) Automated Solid Phase Oligosaccharide Synthesis. 278:<br />

289–309<br />

Chaires JB (2005) Structural Selectivity <strong>of</strong> Drug-Nucleic Acid Interactions Probed by Competition<br />

Dialysis. 253: 33–53<br />

Cherkinsky M, see Braverman S (2007) 275: 67–101<br />

Chiorboli C, Indelli MT, Sc<strong>and</strong>ola F (2005) Photoinduced Electron/Energy Transfer Across<br />

Molecular Bridges in Binuclear Metal Complexes. 257: 63–102<br />

Chiorboli C, see Indelli MT (2007) 280: 215–255<br />

Coleman AW, Perret F, Moussa A, Dupin M, Guo Y, Perron H (2007) Calix[n]arenes as<br />

Protein Sensors. 277: 31–88


Author Index Volumes 251–280 259<br />

Cölfen H (2007) Bio-inspired Mineralization Using Hydrophilic Polymers. 271:1–77<br />

Collin J-P, Heitz V, Sauvage J-P (2005) Transition-Metal-Complexed Catenanes <strong>and</strong> Rotaxanes<br />

in Motion: Towards Molecular Machines. 262: 29–62<br />

Collins BE, Wright AT, Anslyn EV (2007) Combining Molecular Recognition, Optical Detection,<br />

<strong>and</strong> Chemometric Analysis. 277: 181–218<br />

Collyer SD, see Davis F (2005) 255: 97–124<br />

Commeyras A, see Pascal R (2005) 259: 69–122<br />

Coquerel G (2007) Preferential Crystallization. 269:1–51<br />

Correia JDG, see Santos I (2005) 252: 45–84<br />

Costanzo G, see Saladino R (2005) 259: 29–68<br />

Cotarca L, see Zonta C (2007) 275: 131–161<br />

Credi A, see Balzani V (2005) 262:1–27<br />

Crestini C, see Saladino R (2005) 259: 29–68<br />

Crich D, Brebion F, Suk D-H (2006) Generation <strong>of</strong> Alkene Radical Cations by Heterolysis<br />

<strong>of</strong> β-Substituted Radicals: Mechanism, Stereochemistry, <strong>and</strong> Applications in Synthesis.<br />

263:1–38<br />

Cuerva JM, Justicia J, Oller-López JL, Oltra JE (2006) Cp2TiCl in Natural Product Synthesis.<br />

264: 63–92<br />

DaasbjergK,SvithH,GrimmeS,GerenkampM,Mück-LichtenfeldC,GansäuerA,Barchuk<br />

A (2006) The Mechanism <strong>of</strong> Epoxide Opening through Electron Transfer: Experiment<br />

<strong>and</strong> Theory in Concert. 263: 39–70<br />

D’Addario D, see Braga D (2005) 254: 71–94<br />

Danishefsky SJ, see Warren JD (2007) 267: 109–141<br />

Darmency V, Renaud P (2006) Tin-Free Radical Reactions Mediated by Organoboron <strong>Compounds</strong>.<br />

263: 71–106<br />

Davis F, Collyer SD, Higson SPJ (2005) The Construction <strong>and</strong> Operation <strong>of</strong> Anion Sensors:<br />

Current Status <strong>and</strong> Future Perspectives. 255: 97–124<br />

Deamer DW, Dworkin JP (2005) Chemistry <strong>and</strong> Physics <strong>of</strong> Primitive Membranes. 259:1–27<br />

Debaene F, see Winssinger N (2007) 278: 311–342<br />

Dellago C, Bolhuis PG (2007) Transition Path Sampling Simulations <strong>of</strong> Biological Systems.<br />

268: 291–317<br />

Deng J-Y, see Zhang X-E (2005) 261: 169–190<br />

Dervan PB, Poulin-Kerstien AT, Fechter EJ, Edelson BS (2005) Regulation <strong>of</strong> Gene Expression<br />

by Synthetic DNA-Binding Lig<strong>and</strong>s. 253:1–31<br />

Dias N, Vezin H, Lansiaux A, Bailly C (2005) Topoisomerase Inhibitors <strong>of</strong> Marine Origin<br />

<strong>and</strong> Their Potential Use as Anticancer Agents. 253: 89–108<br />

DiMauro E, see Saladino R (2005) 259: 29–68<br />

Dittrich M, Yu J, Schulten K (2007) PcrA Helicase, a Molecular Motor Studied from the<br />

Electronic to the Functional Level. 268: 319–347<br />

Dobrawa R, see You C-C (2005) 258: 39–82<br />

Du Q, Larsson O, Swerdlow H, Liang Z (2005) DNA Immobilization: Silanized Nucleic Acids<br />

<strong>and</strong> Nanoprinting. 261: 45–61<br />

Duatti A, see Boschi A (2005) 252: 85–115<br />

Dupin M, see Coleman AW (2007) 277: 31–88<br />

Dworkin JP, see Deamer DW (2005) 259:1–27<br />

Edelson BS, see Dervan PB (2005) 253:1–31<br />

Edwards DS, see Liu S (2005) 252: 193–216<br />

Ernst K-H (2006) Supramolecular Surface Chirality. 265: 209–252


260 Author Index Volumes 251–280<br />

Ersmark K, see Wannberg J (2006) 266: 167–197<br />

Escudé C, Sun J-S (2005) DNA Major Groove Binders: Triple Helix-Forming Oligonucleotides,<br />

Triple Helix-Specific DNA Lig<strong>and</strong>s <strong>and</strong> Cleaving Agents. 253: 109–148<br />

Evans SV, see Schuman B (2007) 272: 217–258<br />

Van der Eycken E, see Appukkuttan P (2006) 266:1–47<br />

Fages F, Vögtle F, ˇZinić M (2005) Systematic Design <strong>of</strong> Amide- <strong>and</strong> Urea-Type Gelators with<br />

Tailored Properties. 256: 77–131<br />

Fages F, see Žinić M (2005) 256: 39–76<br />

Faigl F, Schindler J, Fogassy E (2007) Advantages <strong>of</strong> Structural Similarities <strong>of</strong> the Reactants<br />

in Optical Resolution Processes. 269: 133–157<br />

Fan C-A, see Gansäuer A (2007) 279: 25–52<br />

Fantucci P, see Bertini L (2007) 268:1–46<br />

Fechter EJ, see Dervan PB (2005) 253:1–31<br />

Fensterbank L, see Albert M (2006) 264:1–62<br />

Fernández JM, see Moonen NNP (2005) 262: 99–132<br />

Fern<strong>and</strong>o C, see Szathmáry E (2005) 259: 167–211<br />

Ferrer B, see Balzani V (2005) 262:1–27<br />

De Feyter S, De Schryver F (2005) Two-Dimensional Dye Assemblies on Surfaces Studied<br />

by Scanning Tunneling Microscopy. 258: 205–255<br />

Fischer D, Geyer A (2007) NMR Analysis <strong>of</strong> Bioprotective Sugars: Sucrose <strong>and</strong> Oligomeric<br />

(1→2)-α-d-glucopyranosyl-(1→2)-β-d-fruct<strong>of</strong>uranosides. 272: 169–186<br />

Flood AH, see Moonen NNP (2005) 262: 99–132<br />

Fogassy E, see Faigl F (2007) 269: 133–157<br />

Fricke M, Volkmer D (2007) Crystallization <strong>of</strong> Calcium Carbonate Beneath Insoluble Monolayers:<br />

Suitable Models <strong>of</strong> Mineral–Matrix Interactions in Biomineralization? 270:1–41<br />

Fujimoto D, see Tamura R (2007) 269: 53–82<br />

Fujiwara S-i, Kambe N (2005) Thio-, Seleno-, <strong>and</strong> Telluro-Carboxylic Acid Esters. 251: 87–<br />

140<br />

Gansäuer A, see Daasbjerg K (2006) 263: 39–70<br />

Garcia-Garibay MA, see Karlen SD (2005) 262: 179–227<br />

Gelinck GH, see Grozema FC (2005) 257: 135–164<br />

Geng X, see Warren JD (2007) 267: 109–141<br />

Gansäuer A, Justicia J, Fan C-A, Worgull D, Piestert F (2007) Reductive C–C Bond Formation<br />

after Epoxide Opening via Electron Transfer. 279: 25–52<br />

George SJ, see Ajayaghosh A (2005) 258: 83–118<br />

Gerenkamp M, see Daasbjerg K (2006) 263: 39–70<br />

Gevorgyan V, see Sromek AW (2007) 274: 77–124<br />

Geyer A, see Fischer D (2007) 272: 169–186<br />

Giaffreda SL, see Braga D (2005) 254: 71–94<br />

Giernoth R (2007) Homogeneous Catalysis in Ionic Liquids. 276:1–23<br />

de Gioia L, see Bertini L (2007) 268:1–46<br />

Di Giusto DA, King GC (2005) Special-Purpose Modifications <strong>and</strong> Immobilized Functional<br />

Nucleic Acids for Biomolecular Interactions. 261: 131–168<br />

Greco C, see Bertini L (2007) 268:1–46<br />

Greiner L, Laue S, Wöltinger J, Liese A (2007) Continuous Asymmetric Hydrogenation. 276:<br />

111–124<br />

Grepioni F, see Braga D (2005) 254: 71–94<br />

Grimme S, see Daasbjerg K (2006) 263: 39–70


Author Index Volumes 251–280 261<br />

Grozema FC, Siebbeles LDA, Gelinck GH, Warman JM (2005) The Opto-Electronic Properties<br />

<strong>of</strong> Isolated Phenylenevinylene Molecular Wires. 257: 135–164<br />

Guiseppi-Elie A, Lingerfelt L (2005) Impedimetric Detection <strong>of</strong> DNA Hybridization: Towards<br />

Near-Patient DNA Diagnostics. 260: 161–186<br />

Guo Y, see Coleman AW (2007) 277: 31–88<br />

Haack M, Beck-Sickinger AG (2007) Multiple Peptide Synthesis to Identify Bioactive Hormone<br />

Structures. 278: 243–288<br />

Haase C, Seitz O (2007) Chemical Synthesis <strong>of</strong> Glycopeptides. 267:1–36<br />

Hahn F, Schepers U (2007) Solid Phase Chemistry for the Directed Synthesis <strong>of</strong> Biologically<br />

Active Polyamine Analogs, Derivatives, <strong>and</strong> Conjugates. 278: 135–208<br />

Hansen SG, Skrydstrup T (2006) Modification <strong>of</strong> Amino Acids, Peptides, <strong>and</strong> Carbohydrates<br />

through Radical Chemistry. 264: 135–162<br />

Harmer NJ (2007) The Fibroblast Growth Factor (FGF) – FGF Receptor Complex: Progress<br />

Towards the Physiological State. 272: 83–116<br />

Harry SR, see Carney CK (2007) 270: 155–185<br />

Heise C, Bier FF (2005) Immobilization <strong>of</strong> DNA on Microarrays. 261:1–25<br />

Heitz V, see Collin J-P (2005) 262: 29–62<br />

Herrmann C, Reiher M (2007) First-Principles Approach to Vibrational Spectroscopy <strong>of</strong><br />

Biomolecules. 268: 85–132<br />

Higson SPJ, see Davis F (2005) 255: 97–124<br />

Hirao T (2007) Catalytic Reductive Coupling <strong>of</strong> Carbonyl <strong>Compounds</strong> – The Pinacol Coupling<br />

Reaction <strong>and</strong> Beyond. 279: 53–75<br />

Hirayama N, see Sakai K (2007) 269: 233–271<br />

Hirst AR, Smith DK (2005) Dendritic Gelators. 256: 237–273<br />

Holzwarth AR, see Balaban TS (2005) 258:1–38<br />

Homans SW (2007) Dynamics <strong>and</strong> Thermodynamics <strong>of</strong> Lig<strong>and</strong>–Protein Interactions. 272:<br />

51–82<br />

Houk RJT, Tobey SL, Anslyn EV (2005) Abiotic Guanidinium Receptors for Anion Molecular<br />

Recognition <strong>and</strong> Sensing. 255: 199–229<br />

Huc I, see Brizard A (2005) 256: 167–218<br />

Ihmels H, Otto D (2005) Intercalation <strong>of</strong> Organic Dye Molecules into Double-Str<strong>and</strong>ed DNA<br />

– General Principles <strong>and</strong> Recent Developments. 258: 161–204<br />

Iida H, Krische MJ (2007) Catalytic Reductive Coupling <strong>of</strong> Alkenes <strong>and</strong> Alkynes to Carbonyl<br />

<strong>Compounds</strong> <strong>and</strong> Imines Mediated by Hydrogen. 279: 77–104<br />

Imai H (2007) Self-Organized Formation <strong>of</strong> Hierarchical Structures. 270: 43–72<br />

Indelli MT, Chiorboli C, Sc<strong>and</strong>ola F (2007) <strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong><br />

<strong>Compounds</strong>: Rhodium. 280: 215–255<br />

Indelli MT, see Chiorboli C (2005) 257: 63–102<br />

Inoue Y, see Borovkov VV (2006) 265: 89–146<br />

Ishii A, Nakayama J (2005) Carbodithioic Acid Esters. 251: 181–225<br />

Ishii A, Nakayama J (2005) Carboselenothioic <strong>and</strong> Carbodiselenoic Acid Derivatives <strong>and</strong><br />

Related <strong>Compounds</strong>. 251: 227–246<br />

Ishi-i T, Shinkai S (2005) Dye-Based Organogels: Stimuli-Responsive S<strong>of</strong>t Materials Based<br />

on One-Dimensional Self-Assembling Aromatic Dyes. 258: 119–160<br />

James DK, Tour JM (2005) Molecular Wires. 257: 33–62<br />

James TD (2007) Saccharide-Selective Boronic Acid Based Photoinduced Electron Transfer<br />

(PET) Fluorescent Sensors. 277: 107–152


262 Author Index Volumes 251–280<br />

Jelinek R, Kolusheva S (2007) Biomolecular Sensing with Colorimetric Vesicles. 277: 155–180<br />

Jones W, see Trask AV (2005) 254: 41–70<br />

Jung N, Wiehn M, Bräse S (2007) Multifunctional Linkers for Combinatorial Solid Phase<br />

Synthesis. 278:1–88<br />

Justicia J, see Cuerva JM (2006) 264: 63–92<br />

Justicia J, see Gansäuer A (2007) 279: 25–52<br />

Kambe N, see Fujiwara S-i (2005) 251: 87–140<br />

Kane-Maguire NAP (2007) <strong>Photochemistry</strong> <strong>and</strong> <strong>Photophysics</strong> <strong>of</strong> <strong>Coordination</strong> <strong>Compounds</strong>:<br />

Chromium. 280: 37–67<br />

Kann N, see Ljungdahl N (2007) 278: 89–134<br />

Kano N, Kawashima T (2005) Dithiocarboxylic Acid Salts <strong>of</strong> Group 1–17 Elements (Except<br />

for Carbon). 251: 141–180<br />

Kappe CO, see Kremsner JM (2006) 266: 233–278<br />

Kaptein B, see Kellogg RM (2007) 269: 159–197<br />

Karlen SD, Garcia-Garibay MA (2005) Amphidynamic Crystals: Structural Blueprints for<br />

Molecular Machines. 262: 179–227<br />

Kato S, Niyomura O (2005) Group 1–17 Element (Except Carbon) Derivatives <strong>of</strong> Thio-,<br />

Seleno- <strong>and</strong> Telluro-Carboxylic Acids. 251: 19–85<br />

Kato S, see Niyomura O (2005) 251:1–12<br />

Kato T, Mizoshita N, Moriyama M, Kitamura T (2005) Gelation <strong>of</strong> Liquid Crystals with<br />

Self-Assembled Fibers. 256: 219–236<br />

Kaul M, see Pilch DS (2005) 253: 179–204<br />

Kaupp G (2005) Organic Solid-State Reactions with 100% Yield. 254: 95–183<br />

Kawasaki T, see Okahata Y (2005) 260: 57–75<br />

Kawashima T, see Kano N (2005) 251: 141–180<br />

Kay ER, Leigh DA (2005) Hydrogen Bond-Assembled Synthetic Molecular Motors <strong>and</strong><br />

Machines. 262: 133–177<br />

Kellogg RM, Kaptein B, Vries TR (2007) Dutch Resolution <strong>of</strong> Racemates <strong>and</strong> the Roles <strong>of</strong><br />

Solid Solution Formation <strong>and</strong> Nucleation Inhibition. 269: 159–197<br />

Kessler H, see Weide T (2007) 272:1–50<br />

Kimura M, Tamaru Y (2007) Nickel-Catalyzed Reductive Coupling <strong>of</strong> Dienes <strong>and</strong> Carbonyl<br />

<strong>Compounds</strong>. 279: 173–207<br />

King GC, see Di Giusto DA (2005) 261: 131–168<br />

Kirchner B, see Thar J (2007) 268: 133–171<br />

Kita Y, see Akai S (2007) 274: 35–76<br />

Kitamura T, see Kato T (2005) 256: 219–236<br />

Kniep R, Simon P (2007) Fluorapatite-Gelatine-Nanocomposites: Self-Organized Morphogenesis,<br />

Real Structure <strong>and</strong> Relations to Natural Hard Materials. 270: 73–125<br />

Koenig BW (2007) Residual Dipolar Couplings Report on the Active Conformation <strong>of</strong><br />

Rhodopsin-Bound Protein Fragments. 272: 187–216<br />

Kolusheva S, see Jelinek R (2007) 277: 155–180<br />

Komatsu K (2005) The Mechanochemical Solid-State Reaction <strong>of</strong> Fullerenes. 254: 185–206<br />

Kremsner JM, Stadler A, Kappe CO (2006) The Scale-Up <strong>of</strong> Microwave-Assisted Organic<br />

Synthesis. 266: 233–278<br />

Kriegisch V, Lambert C (2005) Self-Assembled Monolayers <strong>of</strong> Chromophores on Gold Surfaces.<br />

258: 257–313<br />

Krische MJ, see Iida H (2007) 279: 77–104<br />

Kuhn LT, Bargon J (2007) Transfer <strong>of</strong> Parahydrogen-Induced Hyperpolarization to Heteronuclei.<br />

276: 25–68


Author Index Volumes 251–280 263<br />

Kuhn LT, Bargon J (2007) Exploiting Nuclear Spin Polarization to Investigate Free Radical<br />

Reactions via in situ NMR. 276: 125–154<br />

Lacôte E, see Albert M (2006) 264:1–62<br />

Lahav M, see Weissbuch I (2005) 259: 123–165<br />

Lambert C, see Kriegisch V (2005) 258: 257–313<br />

Lansiaux A, see Dias N (2005) 253: 89–108<br />

LaPlante SR (2007) Exploiting Lig<strong>and</strong> <strong>and</strong> Receptor Adaptability in Rational Drug Design<br />

Using Dynamics <strong>and</strong> Structure-Based Strategies. 272: 259–296<br />

Larhed M, see Nilsson P (2006) 266: 103–144<br />

Larhed M, see Wannberg J (2006) 266: 167–197<br />

Larsson O, see Du Q (2005) 261: 45–61<br />

Laue S, see Greiner L (2007) 276: 111–124<br />

Leigh DA, Pérez EM (2006) Dynamic Chirality: Molecular Shuttles <strong>and</strong> Motors. 265: 185–208<br />

Leigh DA, see Kay ER (2005) 262: 133–177<br />

Leiserowitz L, see Weissbuch I (2005) 259: 123–165<br />

Lhoták P (2005) Anion Receptors Based on Calixarenes. 255: 65–95<br />

Li WP, Meyer LA, Anderson CJ (2005) Radiopharmaceuticals for Positron Emission Tomography<br />

Imaging <strong>of</strong> Somatostatin Receptor Positive Tumors. 252: 179–192<br />

Liang Z, see Du Q (2005) 261: 45–61<br />

Liese A, see Greiner L (2007) 276: 111–124<br />

Lingerfelt L, see Guiseppi-Elie A (2005) 260: 161–186<br />

Listorti A, see Armaroli N (2007) 280: 69–115<br />

Litvinchuk S, see Matile S (2007) 277: 219–250<br />

Liu S (2005) 6-Hydrazinonicotinamide Derivatives as Bifunctional Coupling Agents for<br />

99m Tc-Labeling <strong>of</strong> Small Biomolecules. 252: 117–153<br />

Liu S, Robinson SP, Edwards DS (2005) Radiolabeled Integrin αvβ3 Antagonists as Radiopharmaceuticals<br />

for Tumor Radiotherapy. 252: 193–216<br />

Liu XY (2005) Gelation with Small Molecules: from Formation Mechanism to Nanostructure<br />

Architecture. 256:1–37<br />

Ljungdahl N, Bromfield K, Kann N (2007) Solid Phase Organometallic Chemistry. 278:<br />

89–134<br />

De Lucchi O, see Zonta C (2007) 275: 131–161<br />

Luderer F, Walschus U (2005) Immobilization <strong>of</strong> Oligonucleotides for Biochemical Sensing<br />

by Self-Assembled Monolayers: Thiol-Organic Bonding on Gold <strong>and</strong> Silanization on<br />

Silica Surfaces. 260: 37–56<br />

Maeda K, Yashima E (2006) Dynamic Helical Structures: Detection <strong>and</strong> Amplification <strong>of</strong><br />

Chirality. 265: 47–88<br />

Magnera TF, Michl J (2005) Altitudinal Surface-Mounted Molecular Rotors. 262: 63–97<br />

Maini L, see Braga D (2005) 254: 71–94<br />

Malacria M, see Albert M (2006) 264:1–62<br />

Marquette CA, Blum LJ (2005) Beads Arraying <strong>and</strong> Beads Used in DNA Chips. 261: 113–129<br />

Mascini M, see Palchetti I (2005) 261: 27–43<br />

Matile S, Tanaka H, Litvinchuk S (2007) Analyte Sensing Across Membranes with Artificial<br />

Pores. 277: 219–250<br />

Matsumoto A (2005) Reactions <strong>of</strong> 1,3-Diene <strong>Compounds</strong> in the Crystalline State. 254: 263–<br />

305<br />

McGhee AM, Procter DJ (2006) Radical Chemistry on Solid Support. 264: 93–134


264 Author Index Volumes 251–280<br />

Mentel M, Breinbauer R (2007) Combinatorial Solid-Phase Natural Product Chemistry. 278:<br />

209–241<br />

Meyer B, Möller H (2007) Conformation <strong>of</strong> Glycopeptides <strong>and</strong> Glycoproteins. 267: 187–251<br />

Meyer LA, see Li WP (2005) 252: 179–192<br />

Michl J, see Magnera TF (2005) 262: 63–97<br />

Milea JS, see Smith CL (2005) 261: 63–90<br />

Mizoshita N, see Kato T (2005) 256: 219–236<br />

Modlinger A, see Weide T (2007) 272:1–50<br />

Möller H, see Meyer B (2007) 267: 187–251<br />

Montgomery J, Sormunen GJ (2007) Nickel-Catalyzed Reductive Couplings <strong>of</strong> Aldehydes<br />

<strong>and</strong> Alkynes. 279:1–23<br />

Moonen NNP, Flood AH, Fernández JM, Stoddart JF (2005) Towards a Rational Design <strong>of</strong><br />

Molecular Switches <strong>and</strong> Sensors from their Basic Building Blocks. 262: 99–132<br />

Moriyama M, see Kato T (2005) 256: 219–236<br />

Moussa A, see Coleman AW (2007) 277: 31–88<br />

Murai T (2005) Thio-, Seleno-, Telluro-Amides. 251: 247–272<br />

Murakami H (2007) From Racemates to Single Enantiomers – Chiral Synthetic Drugs over<br />

the last 20 Years. 269: 273–299<br />

Mutule I, see Suna E (2006) 266: 49–101<br />

Naka K (2007) Delayed Action <strong>of</strong> Synthetic Polymers for Controlled Mineralization <strong>of</strong><br />

Calcium Carbonate. 271: 119–154<br />

Nakayama J, see Ishii A (2005) 251: 181–225<br />

Nakayama J, see Ishii A (2005) 251: 227–246<br />

Narayanan S, see Reif B (2007) 272: 117–168<br />

Nastasi F, see Campagna S (2007) 280: 117–214<br />

Neese F, see Sinnecker S (2007) 268: 47–83<br />

Nguyen GH, see Smith CL (2005) 261: 63–90<br />

Nicolau DV, Sawant PD (2005) Scanning Probe Microscopy Studies <strong>of</strong> Surface-Immobilised<br />

DNA/Oligonucleotide Molecules. 260: 113–160<br />

Niessen HG, Woelk K (2007) Investigations in Supercritical Fluids. 276: 69–110<br />

Nilsson P, Ol<strong>of</strong>sson K, Larhed M (2006) Microwave-Assisted <strong>and</strong> Metal-Catalyzed Coupling<br />

Reactions. 266: 103–144<br />

Nishiyama H, Shiomi T (2007) Reductive Aldol, Michael, <strong>and</strong> Mannich Reactions. 279:<br />

105–137<br />

Niyomura O, Kato S (2005) Chalcogenocarboxylic Acids. 251:1–12<br />

Niyomura O, see Kato S (2005) 251: 19–85<br />

Nohira H, see Sakai K (2007) 269: 199–231<br />

Oda R, see Brizard A (2005) 256: 167–218<br />

Okahata Y, Kawasaki T (2005) Preparation <strong>and</strong> Electron Conductivity <strong>of</strong> DNA-Aligned Cast<br />

<strong>and</strong> LB Films from DNA-Lipid Complexes. 260: 57–75<br />

Okamura T, see Ueyama N (2007) 271: 155–193<br />

Oller-López JL, see Cuerva JM (2006) 264: 63–92<br />

Ol<strong>of</strong>sson K, see Nilsson P (2006) 266: 103–144<br />

Oltra JE, see Cuerva JM (2006) 264: 63–92<br />

Onoda A, see Ueyama N (2007) 271: 155–193<br />

Otto D, see Ihmels H (2005) 258: 161–204<br />

Otto S, Severin K (2007) Dynamic Combinatorial Libraries for the Development <strong>of</strong> Synthetic<br />

Receptors <strong>and</strong> Sensors. 277: 267–288


Author Index Volumes 251–280 265<br />

Palchetti I, Mascini M (2005) Electrochemical Adsorption Technique for Immobilization <strong>of</strong><br />

Single-Str<strong>and</strong>ed Oligonucleotides onto Carbon Screen-Printed Electrodes. 261: 27–43<br />

Pascal R, Boiteau L, Commeyras A (2005) From the Prebiotic Synthesis <strong>of</strong> α-Amino Acids<br />

Towards a Primitive Translation Apparatus for the Synthesis <strong>of</strong> Peptides. 259: 69–122<br />

Paulo A, see Santos I (2005) 252: 45–84<br />

Pérez EM, see Leigh DA (2006) 265: 185–208<br />

Perret F, see Coleman AW (2007) 277: 31–88<br />

Perron H, see Coleman AW (2007) 277: 31–88<br />

Pianowski Z, see Winssinger N (2007) 278: 311–342<br />

Piestert F, see Gansäuer A (2007) 279: 25–52<br />

Pilch DS, Kaul M, Barbieri CM (2005) Ribosomal RNA Recognition by Aminoglycoside<br />

Antibiotics. 253: 179–204<br />

Pividori MI, Alegret S (2005) DNA Adsorption on Carbonaceous Materials. 260:1–36<br />

Piwnica-Worms D, see Sharma V (2005) 252: 155–178<br />

Plesniak K, Zarecki A, Wicha J (2007) The Smiles Rearrangement <strong>and</strong> the Julia–Kocienski<br />

Olefination Reaction. 275: 163–250<br />

Polito M, see Braga D (2005) 254: 71–94<br />

Poulin-Kerstien AT, see Dervan PB (2005) 253:1–31<br />

de la Pradilla RF, Tortosa M, Viso A (2007) Sulfur Participation in [3,3]-Sigmatropic Rearrangements.<br />

275: 103–129<br />

Procter DJ, see McGhee AM (2006) 264: 93–134<br />

Puntoriero F, see Balzani V (2007) 280:1–36<br />

Puntoriero F, see Campagna S (2007) 280: 117–214<br />

Quiclet-Sire B, Zard SZ (2006) The Degenerative Radical Transfer <strong>of</strong> Xanthates <strong>and</strong> Related<br />

Derivatives: An Unusually Powerful Tool for the Creation <strong>of</strong> Carbon–Carbon Bonds. 264:<br />

201–236<br />

Ratner MA, see Weiss EA (2005) 257: 103–133<br />

Raymond KN, see Seeber G (2006) 265: 147–184<br />

Rebek Jr J, see Scarso A (2006) 265:1–46<br />

Reckien W, see Thar J (2007) 268: 133–171<br />

Reggelin M (2007) [2,3]-Sigmatropic Rearrangements <strong>of</strong> Allylic Sulfur <strong>Compounds</strong>. 275:<br />

1–65<br />

Reif B, Narayanan S (2007) Characterization <strong>of</strong> Interactions Between Misfolding Proteins<br />

<strong>and</strong> Molecular Chaperones by NMR Spectroscopy. 272: 117–168<br />

Reiher M, see Herrmann C (2007) 268: 85–132<br />

Renaud P, see Darmency V (2006) 263: 71–106<br />

Revell JD, Wennemers H (2007) Identification <strong>of</strong> Catalysts in Combinatorial Libraries. 277:<br />

251–266<br />

Robinson SP, see Liu S (2005) 252: 193–216<br />

Saha-Möller CR, see You C-C (2005) 258: 39–82<br />

Sakai K, Sakurai R, Hirayama N (2007) Molecular Mechanisms <strong>of</strong> Dielectrically Controlled<br />

Resolution (DCR). 269: 233–271<br />

Sakai K, Sakurai R, Nohira H (2007) New Resolution Technologies Controlled by Chiral<br />

Discrimination Mechanisms. 269: 199–231<br />

Sakamoto M (2005) Photochemical Aspects <strong>of</strong> Thiocarbonyl <strong>Compounds</strong> in the Solid-State.<br />

254: 207–232<br />

Sakurai R, see Sakai K (2007) 269: 199–231


266 Author Index Volumes 251–280<br />

Sakurai R, see Sakai K (2007) 269: 233–271<br />

Saladino R, Crestini C, Costanzo G, DiMauro E (2005) On the Prebiotic Synthesis <strong>of</strong> Nucleobases,<br />

Nucleotides, Oligonucleotides, Pre-RNA <strong>and</strong> Pre-DNA Molecules. 259: 29–68<br />

Santos I, Paulo A, Correia JDG (2005) Rhenium <strong>and</strong> Technetium Complexes Anchored by<br />

Phosphines <strong>and</strong> Scorpionates for Radiopharmaceutical Applications. 252: 45–84<br />

Santos M, see Szathmáry E (2005) 259: 167–211<br />

Sato K (2007) Inorganic-Organic Interfacial Interactions in Hydroxyapatite Mineralization<br />

Processes. 270: 127–153<br />

Sauvage J-P, see Collin J-P (2005) 262: 29–62<br />

Sawant PD, see Nicolau DV (2005) 260: 113–160<br />

Sc<strong>and</strong>ola F, see Chiorboli C (2005) 257: 63–102<br />

Scarso A, Rebek Jr J (2006) Chiral Spaces in Supramolecular Assemblies. 265:1–46<br />

Schaumann E (2007) Sulfur is More Than the Fat Brother <strong>of</strong> Oxygen. An Overview <strong>of</strong><br />

Organosulfur Chemistry. 274:1–34<br />

Scheffer JR, Xia W (2005) Asymmetric Induction in Organic <strong>Photochemistry</strong> via the Solid-<br />

State Ionic Chiral Auxiliary Approach. 254: 233–262<br />

Schenning APHJ, see Ajayaghosh A (2005) 258: 83–118<br />

Schepers U, see Hahn F (2007) 278: 135–208<br />

Schindler J, see Faigl F (2007) 269: 133–157<br />

Schmidtchen FP (2005) Artificial Host Molecules for the Sensing <strong>of</strong> Anions. 255:1–29Author<br />

Index Volumes 251–255<br />

Sc<strong>and</strong>ola F, see Indelli MT (2007) 280: 215–255<br />

Schmuck C, Wich P (2007) The Development <strong>of</strong> Artificial Receptors for Small Peptides Using<br />

Combinatorial Approaches. 277:3–30<br />

Scho<strong>of</strong> S, see Wolter F (2007) 267: 143–185<br />

De Schryver F, see De Feyter S (2005) 258: 205–255<br />

Schulten K, see Dittrich M (2007) 268: 319–347<br />

Schuman B, Alfaro JA, Evans SV (2007) Glycosyltransferase Structure <strong>and</strong> Function. 272:<br />

217–258<br />

Seeber G, Tiedemann BEF, Raymond KN (2006) Supramolecular Chirality in <strong>Coordination</strong><br />

Chemistry. 265: 147–184<br />

Seeberger PH, see Castagner B (2007) 278: 289–309<br />

Seitz O, see Haase C (2007) 267:1–36<br />

Senn HM, Thiel W (2007) QM/MM Methods for Biological Systems. 268: 173–289<br />

Severin K, see Otto S (2007) 277: 267–288<br />

Sewell SL, see Carney CK (2007) 270: 155–185<br />

Sharma V, Piwnica-Worms D (2005) Monitoring Multidrug Resistance P-Glycoprotein Drug<br />

Transport Activity with Single-Photon-Emission Computed Tomography <strong>and</strong> Positron<br />

Emission Tomography Radiopharmaceuticals. 252: 155–178<br />

Shinkai S, see Ishi-i T (2005) 258: 119–160<br />

Shiomi T, see Nishiyama H (2007) 279: 105–137<br />

Sibi MP, see Zimmerman J (2006) 263: 107–162<br />

Siebbeles LDA, see Grozema FC (2005) 257: 135–164<br />

Silvi S, see Balzani V (2005) 262:1–27<br />

Simon P, see Kniep R (2007) 270: 73–125<br />

Sinnecker S, Neese F (2007) Theoretical Bioinorganic Spectroscopy. 268: 47–83<br />

Skrydstrup T, see Hansen SG (2006) 264: 135–162<br />

Smith CL, Milea JS, Nguyen GH (2005) Immobilization <strong>of</strong> Nucleic Acids Using Biotin-<br />

Strept(avidin) Systems. 261: 63–90<br />

Smith DK, see Hirst AR (2005) 256: 237–273


Author Index Volumes 251–280 267<br />

Sormunen GJ, see Montgomery J (2007) 279:1–23<br />

Specker D, Wittmann V (2007) Synthesis <strong>and</strong> Application <strong>of</strong> Glycopeptide <strong>and</strong> Glycoprotein<br />

Mimetics. 267: 65–107<br />

Sromek AW, Gevorgyan V (2007) 1,2-Sulfur Migrations. 274: 77–124<br />

Stadler A, see Kremsner JM (2006) 266: 233–278<br />

Stibor I, Zlatuˇsková P (2005) Chiral Recognition <strong>of</strong> Anions. 255: 31–63<br />

Stoddart JF, see Moonen NNP (2005) 262: 99–132<br />

Strauss CR, Varma RS (2006) Microwaves in Green <strong>and</strong> Sustainable Chemistry. 266: 199–231<br />

Suk D-H, see Crich D (2006) 263:1–38<br />

Suksai C, Tuntulani T (2005) Chromogenetic Anion Sensors. 255: 163–198<br />

Sun J-S, see Escudé C (2005) 253: 109–148<br />

Suna E, Mutule I (2006) Microwave-assisted Heterocyclic Chemistry. 266: 49–101<br />

Süssmuth RD, see Wolter F (2007) 267: 143–185<br />

Svith H, see Daasbjerg K (2006) 263: 39–70<br />

Swerdlow H, see Du Q (2005) 261: 45–61<br />

Szathmáry E, Santos M, Fern<strong>and</strong>o C (2005) Evolutionary Potential <strong>and</strong> Requirements for<br />

Minimal Protocells. 259: 167–211<br />

Taira S, see Yokoyama K (2005) 261: 91–112<br />

Takahashi H, see Tamura R (2007) 269: 53–82<br />

Takahashi K, see Ueyama N (2007) 271: 155–193<br />

Tamiaki H, see Balaban TS (2005) 258:1–38<br />

Tamaru Y, see Kimura M (2007) 279: 173–207<br />

Tamura R, Takahashi H, Fujimoto D, Ushio T (2007) Mechanism <strong>and</strong> Scope <strong>of</strong> Preferential<br />

Enrichment, a Symmetry-Breaking Enantiomeric Resolution Phenomenon. 269: 53–82<br />

Tanaka H, see Matile S (2007) 277: 219–250<br />

Thar J, Reckien W, Kirchner B (2007) Car–Parrinello Molecular Dynamics Simulations <strong>and</strong><br />

Biological Systems. 268: 133–171<br />

Thayer DA, Wong C-H (2007) Enzymatic Synthesis <strong>of</strong> Glycopeptides <strong>and</strong> Glycoproteins.<br />

267: 37–63<br />

Thiel W, see Senn HM (2007) 268: 173–289<br />

Tiedemann BEF, see Seeber G (2006) 265: 147–184<br />

Tobey SL, see Houk RJT (2005) 255: 199–229<br />

Toda F (2005) Thermal <strong>and</strong> Photochemical Reactions in the Solid-State. 254:1–40<br />

Tortosa M, see de la Pradilla RF (2007) 275: 103–129<br />

Tour JM, see James DK (2005) 257: 33–62<br />

Trask AV, Jones W (2005) Crystal Engineering <strong>of</strong> Organic Cocrystals by the Solid-State<br />

Grinding Approach. 254: 41–70<br />

Tuntulani T, see Suksai C (2005) 255: 163–198<br />

Uccelli L, see Boschi A (2005) 252: 85–115<br />

Ueyama N, Takahashi K, Onoda A, Okamura T, Yamamoto H (2007) Inorganic–Organic<br />

Calcium Carbonate Composite <strong>of</strong> Synthetic Polymer Lig<strong>and</strong>s with an Intramolecular<br />

NH···OHydrogenBond.271: 155–193<br />

Ushio T, see Tamura R (2007) 269: 53–82<br />

Varma RS, see Strauss CR (2006) 266: 199–231<br />

Veciana J, see Amabilino DB (2006) 265: 253–302<br />

Venturi M, see Balzani V (2005) 262:1–27<br />

Vezin H, see Dias N (2005) 253: 89–108


268 Author Index Volumes 251–280<br />

Viso A, see de la Pradilla RF (2007) 275: 103–129<br />

Vögtle F, see Fages F (2005) 256: 77–131<br />

Vögtle M, see Žinić M (2005) 256: 39–76<br />

Volkmer D, see Fricke M (2007) 270:1–41<br />

Volpicelli R, see Zonta C (2007) 275: 131–161<br />

Vries TR, see Kellogg RM (2007) 269: 159–197<br />

Walschus U, see Luderer F (2005) 260: 37–56<br />

Walton JC (2006) Unusual Radical Cyclisations. 264: 163–200<br />

Wannberg J, Ersmark K, Larhed M (2006) Microwave-Accelerated Synthesis <strong>of</strong> Protease<br />

Inhibitors. 266: 167–197<br />

Warman JM, see Grozema FC (2005) 257: 135–164<br />

Warren JD, Geng X, Danishefsky SJ (2007) Synthetic Glycopeptide-Based Vaccines. 267:<br />

109–141<br />

Wasielewski MR, see Weiss EA (2005) 257: 103–133<br />

Weide T, Modlinger A, Kessler H (2007) Spatial Screening for the Identification <strong>of</strong> the<br />

Bioactive Conformation <strong>of</strong> Integrin Lig<strong>and</strong>s. 272:1–50<br />

Weiss EA, Wasielewski MR, Ratner MA (2005) Molecules as Wires: Molecule-Assisted Movement<br />

<strong>of</strong> Charge <strong>and</strong> Energy. 257: 103–133<br />

Weissbuch I, Leiserowitz L, Lahav M (2005) Stochastic “Mirror Symmetry Breaking” via Self-<br />

Assembly, Reactivity <strong>and</strong> Amplification <strong>of</strong> Chirality: Relevance to Abiotic Conditions.<br />

259: 123–165<br />

Wennemers H, see Revell JD (2007) 277: 251–266<br />

Wich P, see Schmuck C (2007) 277:3–30<br />

Wicha J, see Plesniak K (2007) 275: 163–250<br />

Wiehn M, see Jung N (2007) 278:1–88<br />

Williams LD (2005) Between Objectivity <strong>and</strong> Whim: Nucleic Acid Structural Biology. 253:<br />

77–88<br />

Winssinger N, Pianowski Z, Debaene F (2007) Probing Biology with Small Molecule Microarrays<br />

(SMM). 278: 311–342<br />

Wittmann V, see Specker D (2007) 267: 65–107<br />

Wright DW, see Carney CK (2007) 270: 155–185<br />

Woelk K, see Niessen HG (2007) 276: 69–110<br />

Wolter F, Scho<strong>of</strong> S, Süssmuth RD (2007) Synopsis <strong>of</strong> Structural, Biosynthetic, <strong>and</strong> Chemical<br />

Aspects <strong>of</strong> Glycopeptide Antibiotics. 267: 143–185<br />

Wöltinger J, see Greiner L (2007) 276: 111–124<br />

Wong C-H, see Thayer DA (2007) 267: 37–63<br />

Wong KM-C, see Yam VW-W (2005) 257:1–32<br />

Worgull D, see Gansäuer A (2007) 279: 25–52<br />

Wright AT, see Collins BE (2007) 277: 181–218<br />

Würthner F, see You C-C (2005) 258: 39–82<br />

Xia W, see Scheffer JR (2005) 254: 233–262<br />

Yam VW-W, Wong KM-C (2005) Luminescent Molecular Rods – Transition-Metal Alkynyl<br />

Complexes. 257:1–32<br />

Yamamoto H, see Ueyama N (2007) 271: 155–193<br />

Yashima E, see Maeda K (2006) 265: 47–88<br />

Yokoyama K, Taira S (2005) Self-Assembly DNA-Conjugated Polymer for DNA Immobilization<br />

on Chip. 261: 91–112


Author Index Volumes 251–280 269<br />

Yoshikawa I, see Araki K (2005) 256: 133–165<br />

Yoshioka R (2007) Racemization, Optical Resolution <strong>and</strong> Crystallization-Induced Asymmetric<br />

Transformation <strong>of</strong> Amino Acids <strong>and</strong> Pharmaceutical Intermediates. 269: 83–132<br />

You C-C, Dobrawa R, Saha-Möller CR, Würthner F (2005) Metallosupramolecular Dye<br />

Assemblies. 258: 39–82<br />

Yu J, see Dittrich M (2007) 268: 319–347<br />

Yu S-H (2007) Bio-inspired Crystal Growth by Synthetic Templates. 271: 79–118<br />

Zampella G, see Bertini L (2007) 268:1–46<br />

Zard SZ, see Quiclet-Sire B (2006) 264: 201–236<br />

Zarecki A, see Plesniak K (2007) 275: 163–250<br />

Zhang W (2006) Microwave-Enhanced High-Speed Fluorous Synthesis. 266: 145–166<br />

Zhang X-E, Deng J-Y (2005) Detection <strong>of</strong> Mutations in Rifampin-Resistant Mycobacterium<br />

Tuberculosis by Short Oligonucleotide Ligation Assay on DNA Chips (SOLAC). 261:<br />

169–190<br />

Zimmerman J, Sibi MP (2006) Enantioselective Radical Reactions. 263: 107–162<br />

ˇZinić M, see Fages F (2005) 256: 77–131<br />

Žinić M, Vögtle F, Fages F (2005) Cholesterol-Based Gelators. 256: 39–76<br />

Zipse H (2006) Radical Stability—A Theoretical Perspective. 263: 163–190<br />

Zlatuˇsková P, see Stibor I (2005) 255: 31–63<br />

Zonta C,DeLucchi O,Volpicelli R,Cotarca L(2007)Thione–Thiol Rearrangement:Miyazaki–<br />

Newman–Kwart Rearrangement <strong>and</strong> Others. 275: 131–161


Subject Index<br />

Alkyne bridges 148<br />

Angular overlap model (AOM) 41<br />

Antenna system 26<br />

Antennae, artificial light-harvesting<br />

119<br />

Azido-Cr(III) 62<br />

Back-intersystem crossing (BISC) 51<br />

Bimolecular processes 10<br />

–, metal complexes 11<br />

–, quenching 94<br />

Bis(alkyl)naphthalene 149<br />

Bis[2-(diphenylphosphino)phenyl] ether)<br />

95<br />

Bisphenanthroline Cu(I) complexes 78<br />

trans-Chalcone 29<br />

Chemiluminescence 131<br />

Chromium 37<br />

–, coordination compounds,<br />

photochemistry/photophysics 37<br />

Cluster centered (CC) character 70<br />

Clusters 69<br />

<strong>Coordination</strong> compounds, chromium,<br />

photochemistry/photophysics 37<br />

–, photochemical molecular<br />

devices/machines 24<br />

Copper 70<br />

–, biology 73<br />

Coulombic mechanism 22<br />

Cr(acac)3 42<br />

[Cr[18]aneN6] 3+ 49<br />

[Cr(diimine)3] 3+ systems, photoredox<br />

behavior 54<br />

[Cr(N4)(CN)2] + 51<br />

[Cu(NN)2] + complexes 92<br />

[Cr(phen)3] 3+<br />

photoracemization/hydrolysis 43<br />

[Cr(sen)3] 3+ 49<br />

Cr(III) lig<strong>and</strong> field excited states, ultrafast<br />

dynamics 41<br />

Cr(III) porphyrins, axial lig<strong>and</strong><br />

photodissociation 45<br />

Cu(I) 71<br />

–, luminescent complexes 107<br />

–, supramolecular chemistry 78<br />

Cu(I)-bisphenanthroline 79, 81<br />

Cu(II) 71<br />

Cuprous halide clusters 101<br />

Cyclam 47<br />

Cytochrome c oxidase 78<br />

Dendrimers 26, 155<br />

Diimine/diphosphine [Cu(NN)(PP)] +<br />

complexes 95<br />

Diphosphine 95<br />

DNA binding, rhodium complexes,<br />

photocleavage 241<br />

DNA damage, long-range oxidative, excited<br />

Rh(III) complexes 248<br />

DNA interactions 56<br />

DNA intercalators 215, 241<br />

DNA photocleavage 242<br />

Donor–chromophore–acceptor triads 164<br />

Dyads 227<br />

Dye-sensitized solar cells 117<br />

Electrochemiluminescence 131<br />

Electron collection, photoinduced 239<br />

Electron transfer 16, 69<br />

Emissive excited state(s), luminescence<br />

spectra 88<br />

Energy transfer 21, 37, 53, 69, 117, 215<br />

–, self-exchange between identical<br />

chromophores 53<br />

Eu(III) complexes 95<br />

Exchange mechanism 23<br />

Excimers 15


272 Subject Index<br />

Exciplexes 15<br />

Excited-state decay, intramolecular 8<br />

Excited-state distortion 86<br />

Extension cable 28<br />

Fe(III) porphyrin 13<br />

Formaldehyde 3,4<br />

Grids 153<br />

Halide-to-metal charge transfer (XMCT)<br />

70, 102<br />

Intralig<strong>and</strong> charge transfer (ILCT) 221<br />

Jablonski diagram, [Cr(acac)3] 42<br />

–, light absorption 5<br />

Lanthanide ions, long-lived luminescence<br />

26<br />

LEC devices 99, 100<br />

Lig<strong>and</strong>-centered (LC) transitions 6, 120,<br />

218<br />

Lig<strong>and</strong>-to-metal charge-transfer (LMCT)<br />

transitions 6, 76<br />

Light absorption 8<br />

Light-initiated time-resolved X-ray<br />

absorption spectroscopy (LITR-XAS)<br />

70<br />

Light-powered molecular machines 117<br />

Luminescence 69, 117<br />

Machines, light-powered molecular<br />

117<br />

Marcus inverted region 12<br />

Marcus theory 16<br />

Metal complexes 5<br />

Metal-centered (MC) transitions 6, 120,<br />

217<br />

Metalloproteins, copper 75<br />

Metal-to-lig<strong>and</strong> charge-transfer (MLCT)<br />

transitions 6, 70, 73, 119<br />

4 ′ -Methoxyflavylium ion 29<br />

MLCT excited states 128<br />

–, multiple low-lying, polypyridine lig<strong>and</strong><br />

138<br />

Molecular machines, light-powered 117<br />

Molecular wires 24<br />

Multihole storage, photoinduced, mixed<br />

Ru–Mn 177<br />

Nanomotor, sunlight-powered 30<br />

Naphthalene 10<br />

Nitric oxide (NO) 60<br />

Nitrido complexes, Cr(III) coordinated<br />

azide, photogeneration 61<br />

Nitrido-Cr(V) 62<br />

NO, Cr(III)-coordinated nitrite,<br />

photolabilization 60<br />

Nuclear motions, photoactive molecular<br />

machines 183<br />

OLED devices 69, 99, 100, 133<br />

Oligophenylene bridges 145<br />

Optical electron transfer 20<br />

Os(II) bipyridine-type complexes 11<br />

9,10-Phenanthrenequinonediimine 241<br />

Phenanthroline 69<br />

Phosphorescence microwave double<br />

resonance (PMDR) 221<br />

Photocatalytic processes, supramolecular<br />

species 180<br />

<strong>Photochemistry</strong>, molecular 3<br />

–, supramolecular 12<br />

Photogeneration, hydrogen 180<br />

Photoinduced processes, supramolecular<br />

systems 15<br />

Photonuclease 63<br />

Photoracemization, [Ru(bpy)3] 2+ 127<br />

Photoredox 37<br />

Photosubstitution 37, 43<br />

Photosynthesis, Z-scheme 77<br />

Plastocyanin, blue copper site 75<br />

Polyacetylenic bridges 148<br />

Polyads, oligoproline assemblies 170<br />

Polynuclear complexes 215<br />

Polystyrene, multi-ruthenium assemblies<br />

172<br />

POP 95<br />

Porphyrin-Rh(III) conjugates,<br />

photoinduced electron transfer 234<br />

Pseudorotaxanes 28<br />

Quantum mechanical theory 19<br />

Racks, Ru(II) 153<br />

Rh(III) complexes, as acceptors in electron<br />

transfer reactions 245<br />

–, DNA-mediated long-range electron<br />

transfer 245


Subject Index 273<br />

Rh(III) cyclometalated complexes 223<br />

Rh-DNA 248<br />

Rhodium 215<br />

–, complexes, DNA intercalators 241<br />

–, cyclometalated complexes 223<br />

–, homobinuclear complexes 226<br />

–, mononuclear species 218<br />

–, polynuclear/supramolecular species<br />

226<br />

Rhodium polypyridine complexes 215,<br />

218<br />

[Ru(bpy)3] 2+ 120, 123<br />

Ru(II) complexes, tridentate polypyridine<br />

lig<strong>and</strong>s 136<br />

Ru(II) dendrimers, luminescent 155<br />

Ru(II) polypyridine complexes 117, 119<br />

–, nonradiative decay 133<br />

Ru(II) racks 154<br />

Ru(II)-Rh(III) polypyridine dyads,<br />

photoinduced electron transfer 228<br />

Ru–Os dyads, tridentate lig<strong>and</strong>s 145<br />

Ruthenium 117<br />

–, complexes, biological systems 185<br />

–, species, photoactive multinuclear 153<br />

–, supramolecular photochemistry 141<br />

Solar cells, photoelectrochemical,<br />

dye-sensitized 188<br />

–, photoelectrochemical,<br />

ruthenium-sensitized 191<br />

State energy levels/conversion 38, 117<br />

Sunlight-powered nanomotor 30<br />

Supramolecular species/sensitizers 12,<br />

180, 193<br />

Tb(III) complexes 95<br />

Thermal excited state relaxation 37<br />

Ultrafast dynamics 37<br />

Wires, molecular 24<br />

XMCT 70, 102<br />

XOR logic gate 29<br />

Zn(II) porphyrin 13

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!