Вы находитесь на странице: 1из 247

by

Dipl.-Ing. Georg Wolfbauer

Carried out at the

Department of Chemistry, Monash University Clayton 3168, Melbourne, Australia


from

February 1997 to July 1999


under the supervision of

Professor Alan Maxwell Bond and Professor Douglas Robert MacFarlane

in order to fulfill the legal requirements to obtain the academic title

Doctor of Philosophy in Science

Table of Contents

page II

Table of Contents
Table of Contents .........................................................................................................................................II Plain Word Summary................................................................................................................................ VI Abstract......................................................................................................................................................VII Zusammenfassung ..................................................................................................................................... IX Acknowledgement .....................................................................................................................................XII 1. Introduction to Solar Energy Conversion..............................................................................................1 1.1. Renewable Energy.........................................................................................................................2 1.2. Types of Renewable Energy .........................................................................................................2 1.3. Types of Solar Energy...................................................................................................................3
(a) Thermal Solar Energy Conversion.........................................................................................................4 (b) Thermoelectric Solar Energy Conversion .............................................................................................4 (c) Photoelectric Solar Energy Conversion .................................................................................................4 (d) Chemical Solar Energy Conversion.......................................................................................................4

1.4. Energy Storage ..............................................................................................................................5 1.5. Solar Radiation and Air Mass .......................................................................................................6 1.6. Solid State Solar Cells...................................................................................................................8
1.6.1. Introduction ............................................................................................................................................8 1.6.2. Semiconductor ........................................................................................................................................9 1.6.3. Light and Semiconductors ....................................................................................................................11 1.6.4. The p-n Junction ...................................................................................................................................12 1.6.5. Photovoltaic Processes in a Solar Cell .................................................................................................13 (a) Short Circuit Current, Isc ......................................................................................................................14 (b) Open Circuit Potential, Voc ..................................................................................................................14 (c) Fill Factor.............................................................................................................................................15 (d) Efficiency.............................................................................................................................................15 (e) IPCE-Value (Incident Monochromatic Photon-to-Current Conversion Efficiency) ...........................16

1.7. Dye Sensitized Photoelectrochemcial Solar Cells ......................................................................16


1.7.1. History ..................................................................................................................................................16 1.7.2. SemiconductorSolution Interface .......................................................................................................18 1.7.3. The Dye Sensitized Semiconductor Interface.......................................................................................19 1.7.4. Nano-porous Titanium Dioxide............................................................................................................20 1.7.5. The Dye Sensitized Solar Cell..............................................................................................................22 (a) Operating Principle ..............................................................................................................................22 (b) Side Reactions .....................................................................................................................................23 1.7.6. Comparison to Solid State Cells ...........................................................................................................23

1.8. The Dyes: Octahedral Ruthenium(II) Complexes ......................................................................24


1.8.1. History ..................................................................................................................................................24 1.8.2. Octahedral Complexes..........................................................................................................................26 (a) Valence Bond Theory ..........................................................................................................................26 (b) Molecular Orbital and Ligand Field Theory........................................................................................26 (c) back-bonding.....................................................................................................................................27

1.9. Aim of the Work .........................................................................................................................28 2. Experimental and Theoretical Details..................................................................................................30 2.1. History.........................................................................................................................................31 2.2. Definitions and Basic Units ........................................................................................................32
2.2.1. Current and Potential ............................................................................................................................32 2.2.2. Ohm's Law............................................................................................................................................32 2.2.3. Capacitance...........................................................................................................................................33 2.2.4. Redox Reactions ...................................................................................................................................33 2.2.5. Electrochemical Redox Reactions ........................................................................................................34 2.2.6. Three Electrode Experimental Arrangements ......................................................................................35

2.3. Theory .........................................................................................................................................36


2.3.1. Thermodynamics: The Nernst Equation...............................................................................................36 2.3.2. Mass Transport: Fick's law ...................................................................................................................38 (a) Migration..............................................................................................................................................39 (b) Convection...........................................................................................................................................39 (c) Diffusion ..............................................................................................................................................39

2.4. Linear Sweep and Cyclic Voltammetry ......................................................................................39


2.4.1. Electrodes Used in Voltammetric Experiments....................................................................................40 2.4.2. Solvents ................................................................................................................................................41 2.4.3. Supporting Electrolyte ..........................................................................................................................41

Table of Contents

page III

2.4.4. Linear Sweep Voltammetry..................................................................................................................41 2.4.5. Cyclic Voltammetry .............................................................................................................................42 (a) Chemical Irreversibility .......................................................................................................................44 (b) Heterogeneous Irreversibility ..............................................................................................................44

2.5. Steady State Techniques .............................................................................................................45


2.5.1. Theory...................................................................................................................................................45 (a) "Log-plot" ............................................................................................................................................46 2.5.2. Microdisk Electrode .............................................................................................................................47 2.5.3. Rotating Disk Electrode........................................................................................................................48 2.5.4. Channel Flow Electrode .......................................................................................................................50

2.6. Potential Step and Pulse Techniques ..........................................................................................50


2.6.1. Double Step Chronocoulometry ...........................................................................................................50 2.6.2. Differential Pulse Voltammetry............................................................................................................52 2.6.3. Controlled Potential Bulk Electrolysis .................................................................................................53

2.7. Spectroscopic Techniques...........................................................................................................54


2.7.1. UV/VIS Spectroscopy and Optically Transparent Thin Layer Electrolysis (OTTLE).........................54 2.7.2. Electron Spin Resonance Spectroscopy ...............................................................................................55 2.7.3. Electrospray Mass Spectrometry ..........................................................................................................56 2.7.4. Nuclear Magnetic Resonance Spectroscopy.........................................................................................56

2.8. Electrochemical Quartz Crystal Microbalance ...........................................................................57 3. Studies on Model Esters Compounds...................................................................................................59 3.1. Introduction .................................................................................................................................60 3.2. Experimental ...............................................................................................................................62 3.3. Synthesis......................................................................................................................................64
3.3.1. Synthetic Remarks:...............................................................................................................................65 3.3.2. Characterization by NMR Spectroscopy ..............................................................................................68

3.4. Oxidation.....................................................................................................................................68
3.4.1. L2RuCl2 .................................................................................................................................................68 (a) Voltammetry and Bulk Electrolysis Experiments in DMF..................................................................68 (b) Voltammetric Studies in other Solvents ..............................................................................................70 (c) Spectroelectrochemical Studies ...........................................................................................................71 (d) ESR Spectrum of [L2RuCl2]+...............................................................................................................72 (e) Electrospray-MS Studies on Oxidized Solutions of L2RuCl2 ..............................................................73 (f) Voltammetry of Solid L2RuCl2 Attached to an Electrode Surface ......................................................73 3.4.2. L2RuI2 ...................................................................................................................................................74 (a) Voltammetry in DMF ..........................................................................................................................74 (b) Bulk Electrolysis in DMF....................................................................................................................77 (c) ES-Mass Spectrometric Studies on Bulk Electrolyzed Solutions in DMF ..........................................78 (d) Studies in Other Solvents.....................................................................................................................79 3.4.3. L2Ru(CN)2 ............................................................................................................................................80 (a) Voltammetry and Bulk Electrolysis Experiments in DMF..................................................................80 (b) ES-Mass Spectrometric Studies in DMF.............................................................................................82 (c) Studies in other Solvents......................................................................................................................82 3.4.4. L2Ru(NCS)2 ..........................................................................................................................................83 (a) Voltammetry and Bulk Electrolysis Experiments in DMF..................................................................83 (b) Studies in other Solvents .....................................................................................................................84 3.4.5. Spectroscopic Considerations...............................................................................................................86

3.5. Reduction ....................................................................................................................................87


3.5.1. Et2-dcbpy ..............................................................................................................................................87 3.5.2. L2Ru(CN)2 ............................................................................................................................................88 (a) Voltammetry in DMF ..........................................................................................................................88 (b) Temperature Dependence ....................................................................................................................91 (c) Bulk Electrolysis at 22C and Examination of Reduction Products by ES-MS ..................................93 (d) ESR Measurements on L2Ru(CN)2 Reduced at 55C .......................................................................94 (e) Spectroelectrochemical Studies on Reduced L2Ru(CN)2 ....................................................................96 3.5.3. L2Ru(NCS)2 ..........................................................................................................................................99 (a) Voltammetry in DMF ..........................................................................................................................99 (b) Bulk Electrolysis and ESR Spectra....................................................................................................101 3.5.4. L2RuI2 and L2RuCl2 ............................................................................................................................101 (a) Voltammetry and Reductively Induced Halide Ligand Elimination .................................................101 (b) Bulk Reductive Electrolysis of L2RuI2 in DMF and Identification of Products by ES-MS..............103 3.5.5. Electronic Spectra...............................................................................................................................104

3.6. Conclusion.................................................................................................................................105 3.7. Appendix: Temperature Dependence of Ligand Based Reductions .........................................106
3.7.1. Introduction ........................................................................................................................................106 3.7.2. Experimental: Compounds .................................................................................................................107 3.7.3. Results and Discussion .......................................................................................................................107 3.7.4. Conclusion ..........................................................................................................................................110

Table of Contents

page IV

4. Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 .....................................................................111 4.1. Introduction ...............................................................................................................................112


4.1.1. General................................................................................................................................................112 4.1.2. Aim of the Work.................................................................................................................................113

4.2. Experimental .............................................................................................................................113


4.2.1. Materials and Equipment....................................................................................................................113 4.2.2. Synthesis .............................................................................................................................................114 (a) cis-(H2-dcbpy)2RuCl2 .........................................................................................................................114 (b) cis-(H2-dcbpy)2Ru(NCS)2 ..................................................................................................................115

4.3. Results and Discussion..............................................................................................................117


4.3.1. Cyclic Voltammetry of (H2-dcbpy)2Ru(NCS)2 in Acetone ................................................................117 4.3.2. Voltammetric Oxidation of (H2-dcbpy)2Ru(NCS)2 in Acetone at a Microelectrode .........................120 4.3.3. Rotating Disk Voltammetry of (H2-dcbpy)2Ru(NCS)2 in Acetone ....................................................121 4.3.4. EQCM Studies in Acetone .................................................................................................................122 4.3.5. Cyclic Voltammetry in Acetone in the Presence of SS-bpy...............................................................124 4.3.6. Voltammetry of (H2-dcbpy)2Ru(NCS)2 in THF, Acetonitrile and DMF............................................126 4.3.7. Voltammetry in Acetonitrile of Solid (H2-dcbpy)2Ru(NCS)2 Attached to an Electrode Surface ......128 4.3.8. The Influence of Isomers ....................................................................................................................129 4.3.9. The Influence of Water.......................................................................................................................130

4.4. Conclusions ...............................................................................................................................131 5. Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN)...............................................132 5.1. Introduction ...............................................................................................................................133 5.2. Experimental .............................................................................................................................134
5.2.1. Instrumentation and Electrodes for Voltammetry ..............................................................................134 5.2.2. Instrumentation for other Techniques.................................................................................................134 5.2.3. Reagents and Synthesis of Compounds..............................................................................................135

5.3. Results and Discussion..............................................................................................................135


5.3.1. Electrochemical Reduction of (H2-dcbpy)2Ru(NCS)2 in DMF ..........................................................135 (a) Voltammetry in DMF: The Initial Two One-electron Reduction Processes .....................................135 (b) Bulk Reductive Electrolysis at Platinum Electrodes in DMF ...........................................................140 (c) Bulk Electrolysis at Glassy Carbon Electrodes .................................................................................145 (d) Molecular Orbital Calculations..........................................................................................................145 (e) Voltammetry at Positive Potentials as a Function of Deprotonation.................................................148 (f) Studies on Bu4N-salts of Deprotonated (H2-dcbpy)2Ru(NCS)2 .........................................................151 (g) In Situ Reductive OTTLE Experiments ............................................................................................152 5.3.2. Electrochemical Reduction of (H2-dcbpy)2Ru(CN)2 in DMF ............................................................153 (a) Voltammetry in DMF ........................................................................................................................153 (b) Bulk Reductive Electrolysis in DMF.................................................................................................154 (c) The oxidation Process as a Function of Deprotonation .....................................................................155 (d) Reductive OTTLE Experiments ........................................................................................................155 (e) Bu4N+ "Salts" of (H2-dcbpy)2Ru(CN)2 ..............................................................................................157 5.3.3. Reduction of (H2-dcbpy)2Ru(NCS)2 in Acetone ................................................................................157 (a) Voltammetry, Adsorption and Bulk Electrolysis in Acetone ............................................................157 (b) Adsorption Studies of (H2-dcbpy)2Ru(NCS)2 in Acetone in the Negative Potential Range .............158

5.4. Conclusions ...............................................................................................................................160


5.4.1. Summary.............................................................................................................................................160 5.4.2. Relationship of Results to Photovoltaic Cells ....................................................................................161

6. A Novel Sensitizer: The Black Dye .....................................................................................................163 6.1. Introduction ...............................................................................................................................164 6.2. Experimental .............................................................................................................................165
6.2.1. Instrumentation ...................................................................................................................................165 6.2.2. Reagents and Synthesis of Compounds..............................................................................................165

6.3. Oxidation: Results and Discussion ...........................................................................................167


6.3.1. Studies on the Oxidation Process in Acetonitrile ...............................................................................168 (a) The Metal Centered Oxidation Process .............................................................................................168 (b) Processes at More Positive Potentials................................................................................................171 (c) EQCM Studies of Surface Based Processes ......................................................................................172 6.3.2. Oxidation Studies in Acetone .............................................................................................................173 (a) The Metal Centered Oxidation Process .............................................................................................173 (b) Processes at More Positive Potentials................................................................................................173 (c) Cyclic Voltammetry in the Presence of SS-bpy ................................................................................174 6.3.3. Oxidation Studies in DMF..................................................................................................................174 6.3.4. Voltammetric Oxidation of Mixtures of Thiocyanate Linkage Isomers ............................................176

6.4. Reduction: Results and Discussion ...........................................................................................178


6.4.1. Studies on the Reduction Process in DMF .........................................................................................178 6.4.2. Controlled Potential Reduction and Deprotonation of [(H3-tctpy)Ru(NCS)3] in DMF...................179 (a) Voltammetric Oxidation of [(H3xtctpyx)Ru(NCS)3](1+x) ................................................................182 (b) Voltammetric Reduction of [(tctpyx)Ru(NCS)3](1+x) ......................................................................183 (c) Molecular Orbital Calculations..........................................................................................................183

Table of Contents

page V

6.4.3. Spectroelectrochemical Measurements ..............................................................................................184 6.4.4. Reduction and Deprotonation of the H3-tctpy ligand in DMF ...........................................................187 6.4.5. In Situ Reductive OTTLE Experiments on the H3-tctpy Ligand........................................................188

6.5. Conclusions ...............................................................................................................................188 7. A Channel Flow Cell System Specifically Designed to Test the Efficiency of Redox Shuttles in Dye Sensitized Solar Cells. ..........................................................................................................................190 7.1. Introduction ...............................................................................................................................191 7.2. Aim of the Work .......................................................................................................................192 7.3. Experimental .............................................................................................................................192 7.4. TiO2 Plates.................................................................................................................................193 7.5. Flow Cell Design.......................................................................................................................195 7.6. Results and Discussion..............................................................................................................198
7.6.1. Electrochemistry at TiO2 and SnO2 Electrodes ..................................................................................198 7.6.2. Calibration and Mass Transport in the Channel Flow Cell ................................................................202 7.6.3. Phototransients and Double Layer Capacitance .................................................................................204 7.6.4. The Effect of Cation Size ...................................................................................................................206 7.6.5. Efficiency of Redox Shuttles ..............................................................................................................209 (a) Thermodynamic Requirements and other desirable Properties of Redox Shuttles............................209 Redox Potential ...............................................................................................................................209 High Solubility ................................................................................................................................209 High Diffusion Coefficient .............................................................................................................209 No spectral Characteristics in the Visible Region ..........................................................................209 High Stability of both Reduced and Oxidized Form of the Couple................................................209 Highly Reversible Couple ...............................................................................................................209 (b) Results................................................................................................................................................209 Iodide ..............................................................................................................................................209 Triiodide..........................................................................................................................................211 Bromide, Thiocyanate and Ferrocene .............................................................................................212 Chromium and Molybdenum Complexes .......................................................................................213

7.7. Conclusions ...............................................................................................................................214 8. Conclusions ...........................................................................................................................................216 8.1. The Thermodynamic Importance of the Oxidation Potential ...................................................217 8.2. Surface Based Processes During the Voltammetric Oxidation.................................................217 8.3. Electrochemical Reduction and the Effect of Deprotonation ...................................................218 8.4. Redox Shuttles...........................................................................................................................219 8.5. Future Outlook ..........................................................................................................................219 Reference List............................................................................................................................................221 Curriculum Vitae......................................................................................................................................233

Plain Word Summary

page: VI

Plain Word Summary


A novel type of solar cell, which harvests light with the aid of dyes, has recently gained much attention by having advantages over silicon type solar cells of improved efficiencies under "real" (diffuse) daylight conditions and lower material and production costs. However, fundamental knowledge concerning these so called "dye sensitized solar cells" is very limited. In this dissertation, the electrochemical and photophysical properties of dyes employed in such solar cells have been studied in depth. This knowledge is intrinsically important in the understanding of the operating principles of the solar cell. Additionally, a solar cell test system has been developed specifically for this kind of cell which allows rapid probing of cell efficiencies under different conditions. Thus, for the first time these dyes, as well as charge transporting substances used in these solar cells, have been extensively characterized by a range of electrochemical and spectroscopic techniques.

Abstract

page: VII

Abstract
The work described in this dissertation deals predominantly with the solution phase electrochemical and photophysical properties of sensitizers (dyes) employed in novel dye sensitized solar cells. Additionally, the efficiencies of redox mediators, which are also an integral part of this type of solar cell, have been studied. Wherever appropriate, implications of the importance of the findings are related to the operation of photovoltaic cells containing these compounds. Firstly, a series of cis-(Et2-dcbpy)2RuIIX2 (Et2-dcbpy = 2,2'-bipyridine-4,4'-diethoxydicarboxylic acid, X = Cl, I, NCS and CN) sensitizer ethylester analogues has been synthesized. The voltammetric reduction and oxidation processes of these compounds have been studied in detail. Where applicable, their reduced or oxidized forms and decomposition products were isolated and characterized by a range of spectroscopic techniques such as electron spin resonance spectroscopy, electrospray mass spectrometry and electronic spectroscopy. Reductively induced ligand elimination is the preferred decomposition pathway for the reduced complexes, if the ligands were not able to maintain a strong -back bond, such as for X = Cl and I. A further decomposition pathway for the reduced complexes is de-esterification, yielding the deprotonated acid. The oxidized complex [(Et2-dcbpy)2RuCl2]+ exhibits very high stability allowing its chemical isolation. Oxidation processes for the other complexes are considerably more complicated. In the case of (Et2dcbpy)2RuI2, an oxidatively induced ligand elimination process was observed to occur after formation of [(Et2-dcbpy)2RuI2]+. The rate constants for this and succeeding reactions were estimated from digital simulation of voltammetric data. Oxidation of (Et2-dcbpy)2Ru(NCS)2 leads to elimination of sulfur from the thiocyanate ligand and the formation of (Et2-dcbpy)2Ru(CN)2, which may be further oxidized to mono and poly-nuclear ruthenium compounds. The electrochemical oxidation of the ruthenium photosensitizer cis-(H2-dcbpy)2RuII(NCS)2 (H2dcbpy = 2,2'-bipyridine-4,4'-dicarboxylic acid) was studied in a variety of solvents. Complex surface based reactions as well as chemical reactions are coupled to the charge transfer process under conditions of cyclic voltammetry at macrodisk electrodes. The extent of surface activity was found to depend on solvent and electrode material. An electrochemical quartz crystal microbalance study demonstrates that attachment of solid occurs under open circuit conditions, but that much of the material is lost from the surface at potentials prior to the reversible value. Techniques used to minimize the influence of the surface based processes and chemical reactions coupled to the charge transfer process, such as chemical modification of the electrode surface, steady-state, fast-scan and hydrodynamic voltammetry enable reversible potentials to be measured for the mass transport controlled, solution phase [(H2-dcbpy)2Ru(NCS)2]+/0 redox couple in acetone, acetonitrile, tetrahydrofuran and dimethylformamide. The voltammetric reduction of (H2-dcbpy)2Ru(NCS)2 was even more complicated than its oxidation. Short time domains, reduced temperatures and use of glassy carbon electrodes lead to detection of transiently stable ligand reduced forms of (H2-dcbpy)2Ru(NCS)2. The major decompo-

Abstract

page: VIII

sition pathway for reduced forms of the sensitizer was found to be deprotonation and consequent formation of hydrogen gas. Thus, under certain experimental conditions, [(H-dcbpy)2Ru(NCS)2]2 is believed to be the major product formed by bulk electrolysis. The reversible half-wave potentials for the ligand based reductions of electrochemically generated deprotonated (H2-dcbpy)2Ru(NCS)2 are 0.65V more negative than their protonated counterparts whereas the metal based oxidation process shifted in the negative direction by only 0.3V. Molecular orbital calculations were employed to provide theoretical insights into the effect of deprotonation on reversible potentials and electronic spectra. The voltammetry of the most efficient sensitizer presently available in photoelectrochemical solar cell applications, [(H3-tctpy)RuII(NCS)3] (H3-tctpy = L = 2,2':6',2''-terpyridine-4,4',4''-tricarboxylic acid), was also studied and the results compared to those of the bipyridine analogue. Studies on mixtures containing S- bonded linkage isomers show a shift in reversible oxidation potential to less positive values and a decrease in contribution of surface based processes. The latter fact suggests that surface attachment of these thiocyanate ruthenium complexes predominately occurs via the terminal sulfur of the isothiocyanate ligands. The last part of the thesis is aimed at establishing the efficiency of redox mediators, which shuttle the charge between the dye-sensitized TiO2 electrode and the counter electrode. For this purpose, a dye sensitized solar flow cell test system has been successfully constructed and tested. The new approach represents a significant improvement over current cell assembly procedures for measurements of photocurrents under different solvent/electrolyte conditions. This study confirms the dominant role of the iodide/triiodide system, even though this redox shuttle system possesses many undesirable properties. All alternative redox shuttles tested, including bromide and ferrocene, did not perform satisfactorily and were more than two orders of magnitude less efficient than iodide. However, interestingly, some heavy metal complexes were found to significantly outperform iodide in respect of efficiency under certain conditions. Importantly, the studies show that the performance of redox shuttles in solar cells is not dominated by their thermodynamically properties.

Zusammenfassung

page: IX

Zusammenfassung
Die Arbeit in dieser Dissertation beschreibt die bislang weitgehend unerforschten elektrochemischen und photophysikalischen Eigenschaften von Sensitizern (Farbstoffen) wie sie in farbstoffsensitierten Solarzellen verwendet werden. Weiters wird noch die Effizienz von Redoxmediatoren untersucht, welche auch einen integralen Teil dieser Art von Solarzellen darstellen. Wo immer mglich, werden die Resultate in Relation zu tatschlichen Solarzellen gesetzt und ihre Bedeutungen diskutiert. Zuerst werden die Untersuchungen an der Serie von cis-(Et2-dcbpy)2RuIIX2 (Et2-dcbpy = 2,2'-bipyridine-4,4'-diethoxydicarboxylic acid, X = Cl, I, NCS und CN) Verbindungen beschrieben, welche Esterderivate von hocheffizienten Sensitizern sind. Die voltametrischen Reduktions- und Oxidationsprozesse dieser Verbindungen wurden detailliert studiert. Wenn mglich erfolgte Isolation der reduzierten bzw. oxidierten Produkte und Charakterisierung mittels einer Reihe von spektroskopischen Methoden, wie z.B. Elektronen Spin Resonanz Spektroskopie, Elektrospray Massenspektrometrie und elektronischer Spektroskopie. Reduktiv induzierte Ligandeneliminierung ist der Hauptzersetzungsweg der reduzierten Komplexe, sofern die Liganden nicht in der Lage sind eine starke -Rckbindung mit dem Rutheniumzentrum einzugehen (z.B. X = Cl und I). Ein weiterer Zersetzungsweg der reduzierten Verbindungen ist Deesterfizierung, welche zu den deprotonierten Karbonsureanionen fhrt. Der oxidierte Komplex [(Et 2-dcbpy)2RuCl2]+ ist sehr stabil was seine chemische Isolierung erlaubt, whrend die Oxidationsprozesse der anderen Komplexe sehr kompliziert sind. Im Falle von (Et2-dcbpy)2RuI2 wurde eine oxidativ induzierte Ligandenelimnierung nach der Bildung von [(Et2-dcbpy)2RuI2]+ festgestellt. Die Geschwindigkeitskonstanten dieser und der nachfolgenden Reaktionen wurde mit Hilfe digitaler Simulation errechnet. Oxidation von (Et2-dcbpy)2Ru(NCS)2 fhrt zur Elimnierung des Schwefels im Thiozyanatliganden und zur Bildung von (Et2-dcbpy)2Ru(CN)2. Weiterfolgende Oxidation von (Et2-dcbpy)2Ru(CN)2 fhrt zu dessen Zersetzung zu nicht identifizierbaren Produkten, die wahrscheinlich durch die Reaktion mit dem oxidierten Liganden, Zyanogen (CN)2, entstehen oder seinen Derivaten. Die elektrochemische Oxidation des Rutheniumsensitizers cis-(H2-dcbpy)2RuII(NCS)2 (H2dcbpy = 2,2'-bipyridine-4,4'-dicarboxylic acid) wurde in einer Reihe von Lsungsmitteln studiert. Komplexe Oberflchenreaktionen und chemische Reaktionen begleiten den Ladungstransfer wenn zyklische Voltametrie an Makroelektroden angewendet wird. Das Ausma dieser Obflchenaktivitt hngt vom Lsungsmittel und Elektrodenmaterial ab. Eine Untersuchung mit der elektrochemischen Quarzkristallmikrowaage zeigt, da bereits unter Open Circuit Konditionen Adsorption von Material auftritt, aber da das meiste davon wieder desorbiert vor dem reversiblen Oxidationspotential. Die Anwendung von chemischer Modifikation der Elektrodenoberflchen und von Steady-State, schnell scannender voltametrischer und hydrodynamischer Techniken auf die Oberflchenprozesse und chemischen Reaktionen die den Ladungstransfer begleiten werden

Zusammenfassung

page: X

diskutiert und Werte fr das reversible Oxidationspotential des [(H2-dcbpy)2Ru(NCS)2]0/+ Prozesses in Aceton, Acetonitril, Tetrahydrofuran und Dimethlyformamid bereitgestellt. Weitaus komplizierter als die Oxidation ist die voltametrische Reduktion von (H2-dcbpy)2Ru(NCS)2. Kurze Zeitdomnen, reduzierte Temperaturen und Verwendung von Glaskohlenstoffelektroden fhren zur Detektion von Liganden reduzierter Formen von (H2-dcbpy)2Ru(NCS)2. Der Hauptzersetzungsweg der reduzierten Formen ist Deprotonation mit darauffolgender Bildung von Wasserstoffgas. Das heit, unter bestimmten experimentellen Bedingungen wird durch reduzierende Elektrolyse hauptschlich [(H-dcbpy)2Ru(NCS)2]2 gebildet. Die reversiblen Halbstufenpotentiale der Ligandenreduktionen von elektrochemisch generiertem [(H-dcbpy)2Ru(NCS)2]2 sind 0.65V negativer als die des protonierten Komplexes, whrend die des Rutheniumoxidationsprozesses nur um 0.3V negativ verschoben werden. In Analogie zu Untersuchungen die an (H2-dcbpy)2Ru(NCS)2 durchgefhrt wurden, wurde der zur Zeit effizienteste Farbstoff der fr photoelektrochemische Solarzellanwendungen verfgbar ist, [(H3-tctpy)RuII(NCS)3] (H3-tctpy = L = 2,2':6',2''-terpyridine-4,4',4''-tricarboxylic acid), untersucht und die Resultate mit denen des gemeinhin verwendeten (H2-dcbpy)2Ru(NCS)2 Sensitizers verglichen. Untersuchungen von Mischungen die Isomere mit groem Gehalt an schwefelkoordinierten Thiozyanat enthalten zeigen eine Verschiebung des reversiblen Oxidationspotentials in negative Richtung an sowie eine Verminderung von Strungen durch Oberflchenreaktionen. Der letztere Punkt gibt Anlass zur Spekulation, da wahrscheinlich die Oberflchenwechselwirkungen dieser Rutheniumthiozyanatkomplexe durch den Schwefel der Isothiozyanatliganden erfolgt. Der letzte Teil dieser Dissertation versucht die Effizienz von Redoxmediatoren zu erforschen, welche die Ladung zwischen der farbstoffsensitierten TiO2 Elektrode und der Gegenelektrode transportieren. Fr dieses Vorhaben wurde ein Solarzellen Flow System entwickelt und getestet um damit Photostrme unter oft wechselnden Lsungsmittel/Elektrolytkonditionen zu messen. Dieses System stellt einen wesentlichen Fortschritt zu herkmmlichen Zusammenbauvorgangsweisen von Solarzellen dar. Die Studie besttigt die dominante Rolle des Iodid/Triiodid Redoxmediator Systems, obwohl gezeigt werden kann, da dieses System viele unerwnschte Eigenschaften aufweist. Alle potentiellen Redoxmediatoren die getestet wurden zeigen eine sehr schwache Leistung und sind mehr als zwei Grenordnungen weniger effizient als Iodid. uerst interessant ist, da einige Schwermetallkomplexe unter ganz bestimmten Bedingungen wesentlich besser arbeiten als Iodid. Aus den Resultaten ergibt sich, da die Leistung von Redoxmediatoren nicht entscheidend von ihren thermodynamischen Eigenschaften bestimmt wird.

Declaration

page: XI

Declaration

This work has been not been accepted for the award of any other degree by any other institution and due reference is given to any previously published work.

Acknowledgement

page: XII

Acknowledgement

Adrian van den Bergen ........................................... for synthesis of Bu4NPF6. Alan Bond & Douglas MacFarlane ........................ for their guidance throughout my doctoral studies. Axel, Glenda, Graeme, Melissa, Peter, Steven, Vanda, Vicky and Yvonne ......................... for their friendship. Darren Coomber ..................................................... The only person who has ever read my complete thesis. Thanks! David Humphrey .................................................... for valuable assistance in the interpretation of electronic spectra and for general discussion of inorganic complex chemistry. Glen Deacon & Leone Spiccia ............................... for their valuable discussions and donation of some compounds used in this study. Lai Yoong Goh & Sue Jenkins................................ for their valuable discussion on the syntheses of some of the ruthenium complexes. Julia Howitt ............................................................ for her work on the Red and Black Dye, which she had undertaken during her Honors project. Georgii Lazarev ...................................................... for recording and simulating the ESR spectra. John Eklund ............................................................ for his excellent scientific advise and support in constructing the channel flow solar cell system. Stephen Fletcher & Mike Horne ............................ for their excellent scientific support, generous loan of equipment and for giving me the possibility to conduct part of the work at CSIRO Division of Minerals. Trc V .................................................................. for her companionship and friendship throughout my work in the Bond group. Wujian Miao ........................................................... for his discussion on Chinese foreign policy and for being the first person who has offered me accommodation in Australia. Last but not least, thanks to my wife Claudia who has given up her life in Austria to follow me to the continent on the other side of the world. She has made this stay truly very special.

Chapter 1

Renewable Energy, Solar Cells, Photoelectrochemical Devices, Sensitizers

Chapter 1: Introduction to Solar Energy Conversion Renewable Energy

page: 2

1.1. Renewable Energy


Since the industrial revolution in the 18th century, the demand for energy has increased exponentially. Its presence and availability, which have increased concurrently with the development of technology, have been taken for granted, except for shortages that arose during local crises.1 This was not recognized until the 1970's when the oil crisis struck the world and almost over night the price for crude oil tripled, one of the world's major energy resources. Although this was a political decision rather than a natural catastrophe that caused this incident, it stimulated people's thoughts about energy resources, though, this was also partly driven by government decisions to find solutions to the energy resources problem. Considerable effort was undertaken to identify the limits of current energy resources. At that stage, the availability of crude oil and other fossil energy carriers was estimated to last only shortly beyond the turn of the millenium. However, recent advances in exploration technology have extended this threshold. Nevertheless, the oil crisis was the first time the public was confronted with a possible step back in comfort, the natural driving force of technology and science. The first task was to identify alternative forms of energy which do not posses such finite limits of availability. Of course, nuclear fission was one of the first energy carriers identified and strongly sponsored. However, today we know the problems associated with this kind of energy and its large scale and long-term dangers which are inherent to its technology. It took two major accidents (Three Mile Island 1979 and Chernobyl 1986) to have reinforced these views.

1.2. Types of Renewable Energy


At present, human world energy consumption is made up of about 76% fossil fuels, 14% biomass (wood etc. a form of solar energy), 6% hydroelectricity (also a form of solar energy), 4% nuclear and tiny fraction from other solar energy sources [1]. Renewable energy has become a fashionable term and is unalterably attached to the phrase solar energy. Firstly, we shall try to find a satisfactory definition of the phrase "Renewable Energy". The term "Renewable Energy Resource" is used for energy flows which are replenished at rates comparable to which they are used [2]. Fossil energy (petrol and the like) cannot be a renewable energy resource. The crude oil found on our planet was formed over periods of several tens of thousands of years. However, we have used almost all of it within only a few decades and it becomes clear that this does not meet the requirements of our definition of a renewable energy source. Nuclear energy obtained from nuclear

For example: During World War II, shortages of petrol in Germany resulted in peculiar inventions, such as the wood fired car, which was propelled by wood gases, obtained through distillation of wood.

Chapter 1: Introduction to Solar Energy Conversion Types of Renewable Energy

page: 3

fission is also clearly not a renewable energy because its main resources, e.g. uranium, were formed during the birth of our planet and are therefore limited. Similarly, nuclear fusion is not a renewable energy resource, since it depends on the presence of deuterium. However, these fusion reactions are similar to those that occur in the sun and other stars, and provide a much greater energy potential than nuclear fission. Overall, in a fusion process deuterium, a hydrogen isotope, reacts to give helium and hydrogen and simultaneously releases heat. Although the quantity of starting material is finite, it is so abundant in seawater, that the potential nuclear energy that could be released by nuclear fusion of deuterium contained in one cubic meter of seawater is 1013Jm3 [2]!2 Thus, man's energy expenditure might be sustained for a long time. Nevertheless, radioactive substances are formed as side products and the technology will not be available in time for this generation to make fusion an important energy resource. The quantity of solar energy absorbed by the Earth is about equal to the amount of heat reradiated back into space. Utilization of solar energy by man usually means temporary storage in a usable form (e.g. battery or another natural process like photosynthesis). However, solar energy is eventually reconverted into heat upon utilization, so that the renewable cycle is completed. The same renewable cycle is true for wind, tidal and hydrodynamic energy resources. Wood as an energy resource is more complicated to categorize and demonstrates that the definition of renewable energy cannot be sufficiently rigid to allow the adoption of a universal black and white classification scheme. Timber needs on average 6070 years for regrowth, although some species grow to full size within 1020 years. In the wood energy cycle, a tree is felled and burnt for energy conversion within a maximum of a few months, and this cycle occurs on a timescale which is significantly shorter than its "energy flow rate". However, if proper forestry management techniques are employed so that sufficient forest regrowth is ensured, wood can be, and is at the moment considered to be, a renewable energy resource. The above discussion does not take into account the environmental impact of the energy cycle, such as the greenhouse effect and the release of toxic fumes, undesirable chemicals or dangerous radiation into the atmosphere. Ideally, the environment must include the biosphere of the Earth as well as all other resources directly required for our existence. In order to survive we have to make sure not to adversely alter it. Consequently, a positive attitude toward renewable energy resources is essential and ultimately only the use of this kind of energy will be acceptable.

1.3. Types of Solar Energy


The amount of solar energy intercepted by the Earth and hence the amount of energy flowing in the "solar energy cycle" is about 5.41024J per year [2]. The total worldwide demand for end-use

For comparison: One ton of coal provides 2.9311010J. Also compare the Norbornadiene cycle on page 4 where quadricyclene can store energy in a density of 109Jm3.

Chapter 1: Introduction to Solar Energy Conversion Types of Solar Energy

page: 4

energy has been estimated to be 8109tce per year (tce = tons of coal equivalent). Since one tce = 2.9311010J (1988) [3], this equals 1.091020J and allows us to reach the following conclusion: The energy of less than 10 minutes of sunshine on the planet Earth is equal to the total yearly human energy consumption. Hence, if only a fraction of the solar energy reaching the Earth could be utilized, many energy problems would be solved. The above statement should provide sufficient justification and motivation for anybody to undertake research in the area of solar energy. Solar energy emitted by the sun and reaching the Earth's surface is a form of electromagnetic radiation that is available over a wide spectral range (3002100nm). In order to be used, the radiation needs to be converted into an energy form suitable for our needs. Four different types of solar energy conversion methods are currently available for this purpose. (a) Thermal Solar Energy Conversion The most commonly encountered domestic type of solar energy usage involves the use of solar radiation to directly heat water. In Australia, this method is observed as black patches on roofs of houses and is only used to heat water for domestic purposes, for which otherwise electricity or natural gas would be used. (b) Thermoelectric Solar Energy Conversion Solar energy is converted into heat which in turn is converted into electricity. In this method, conventional fuels, such as coal or natural gas, commonly used to heat water and to generate the steam required to rotate electricity producing turbines in power plants, are replaced by solar energy. This technology requires large sun collection systems and has been applied to construct small to medium sized power plants. (c) Photoelectric Solar Energy Conversion The direct conversion of solar energy into electricity is the best known form of solar energy conversion. This method requires devices such as solar cells (see chapter 1.6). (d) Chemical Solar Energy Conversion This path represents the conversion of solar energy into chemical energy and is important because of its potential to overcome problems with long term storage and transport of energy. Use of endergonic reactions (with a positive change in Gibb's free energy) are especially suited for this purpose. Two examples of photochemical reactions of this kind are the concerted Paterno-Bchi rearrangement reaction (see below) and photocatalytic splitting of water into hydrogen and oxygen.

Chapter 1: Introduction to Solar Energy Conversion Types of Solar Energy

page: 5

storage

quadricyclene

LIGHT + 1000Jcm3 heat

activated complex Eb* product (quadricyclene) catalyst

2+2 photoisomerization

energy

h = Ef * E > 0

HEAT

reactant (norbornadiene) reaction ordinate

reuse

norbornadiene

Figure 1.1. Schematic diagram of the norbornadiene/quadricyclene energy cycle. Right: energy diagram of reaction, Ef* = formation energy for activated complex = minimum solar energy needed for conversion, E = energy gained by reaction reversal, Eb* = energy barrier.

A classical example for the former type of reaction is the norbornadiene cycle [4, 5] shown in Figure 1.1. Norbornadiene is photoisomerized via an endergonic 2+2 Paterno-Bchi cycloaddition to quadricyclene. The energy of one photon is thereby stored in one molecule of quadricyclene, which is an energetically higher state than the starting material. Thus, by choosing an appropriate catalyst [6] the reaction barrier (Eb*, see Figure 1.1 right) can be lowered to allow the exothermic back reaction to take place. Via use of this mechanism, the system can store energy at a high volumetric energy density of approximately 1000Jcm3. This very elegant reaction combines organic photochemistry, catalysis and solar technology. The simplicity of the scheme is striking and it is unfortunate that large scale applications have never been realized.

1.4. Energy Storage


Although solar energy is abundant on this planet, its local availability is limited and is negatively related to the demand of energy. When the demand for energy is largest, such as during winter and in the evenings, the availability is lowest. Since solar energy cannot be utilized on demand, unlike the situation prevailing in coal or nuclear power plants, the energy must be stored. This task of storage is currently one of the biggest challenges in the field of solar energy conversion technology. In the case of thermoelectric power plants, thermal storage of heat occurs via a heat transfer medium like silicon oil. Thermoelectric plants operate at temperatures of 200300C [7] and exhibit round trip efficiencies of over 70%. A much more powerful energy storage medium is hydrogen. The advantage of hydrogen is that it can be used both as a temporary energy storage medium and as a fuel. Hydrogen is a clean fuel which is transportable and storable on a large scale

Chapter 1: Introduction to Solar Energy Conversion Energy Storage

page: 6

in gaseous or liquid form. It is capable of serving most, if not all, of the fuel functions served by today's natural gas, petroleum and coal products. In principle, hydrogen can be produced in almost unlimited amounts from the sole material feedstock: water. During the energy producing step hydrogen is burnt with oxygen and reacts to give water, and hence it is a renewable energy source. The key resource needed for production is a suitable non-fossil primary energy source for splitting water into its elements, hydrogen and oxygen. Solar energy, wind and hydro power converted into process heat and electricity are good candidates for being inexhaustible clean primary energy sources for electrolytic hydrogen production. The direct splitting of water by means of photocatalysis has attracted many researchers in the second half of this century [8-10]. However, work on this type of energy conversion appears to have been abandoned in the past 510 years, probably due to inadequate technological improvements that have occurred during that time. The reaction H2O O2 + H2 requires an energy of 1.23eV, which corresponds to a wavelength of 1000nm. Thus, in theory most of the solar energy available in the visible spectrum (<1000nm) could be utilized for this process. In direct photoelectrochemical water splitting, an n-type semiconductor (e.g. TiO2) is often used as the photoanode. An electron is exited by a photon from the valence band and excited into the conduction band (see Figure 1.2). The resulting electron hole is then filled by water which leads to formation of oxygen. The reduction of water to hydrogen by the conduction band electron occurs at the cathode (not shown in Figure 1.2).
e conduction band Fermi level H+/H2 1.23eV h = Eg H2O/O2 Eg

energy

valence band 2H2O + 4h+ O2 + 4H+

h+

semiconductor

solution

Figure 1.2. Energy level diagram of semiconductor assisted water splitting

1.5. Solar Radiation and Air Mass


The energy of the sun is created by the nuclear fusion reaction of hydrogen and helium which occurs inside the sun at several million degrees. The mass difference that occurs in this process is converted into energy. The hot sun is in radiation equilibrium with the cold universe, and this gives rise to its surface temperature of 5800K. Because all elements are ionized to some degree at this temperature, their spectral lines are strongly broadened so that the gaseous surface of the sun radiates like a black body. The solar energy that reaches Earth is determined by the radiation of the sun and the distance between the Earth and the sun. The solar radiation power just outside the Earth's atmosphere is 1.367kWm2 [11] and this value is know as the solar constant.

Chapter 1: Introduction to Solar Energy Conversion Solar Radiation and Air Mass

page: 7

spectral irradiance [Wm m ]

1500

-2

-1

1000

500

400-800nm visible spectrum

0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

wavelength [m]
Figure 1.3. Spectral distribution of solar radiation. Air mass 1.5 global spectrum [12].

Radiation is partly absorbed and scattered during the course of its journey through the atmosphere. Water and carbon dioxide absorb in the infrared radiation and ozone absorbs ultraviolet radiation. Scattering of radiation is caused by Rayleigh scattering by molecules, Mie scattering by aerosols and cirrus scattering by clouds. For example, in central Europe (Germany) on a cloudless summer day with the sun at the zenith, up to 70% of the total solar radiation can be due to diffuse (scattered and reflected) light and only 30% of light actually hits the surface perpendicular to the sun's path. Furthermore, due to the nature of scattering, the diffuse light component contains a higher fraction of the higher energy ultra violet radiation. Thus, solar cell devices depending on the light arriving at a certain angle will be significantly limited in efficiency so that cells that can collect solar radiation over a large angle and are tuned for optimal performance under diffuse light conditions will be preferred. Specific solar radiation conditions are defined by the Air Mass (AM) value. The spectral distribution and total flux of radiation just outside the Earth's atmosphere, similar to the radiation of a black body of 5800K, has been defined as AM-0. In passing through the atmosphere the radiation becomes attenuated by complex and varying extinction processes mentioned above. At the equator at sea level at noon when the incidence of sunlight is vertical (=90, sun in zenith) and the light travels the shortest distance through the atmosphere and air ("air-mass") to the surface, the spectral solar radiance and flux (1.07kWm2) is defined as AM-1. However, if the angle of light incidence is smaller then 90, the light has to travel through more air-mass than under AM-1 conditions. The relative pathlength through the atmosphere by the shortest geometrical path is given by:
AM = 1 sin

(1.1)

Chapter 1: Introduction to Solar Energy Conversion Solar Radiation and Air Mass

page: 8

The relationship described in Equation 1.1 is only valid for a planar atmosphere, but only introduces an error of less than 1% from the curved atmosphere of the Earth, if >20. For accurate measurements, the local atmospheric pressure also has to be considered. This problem has lead to introduction of the relative air mass AM* given by (p0=1.013bar):
AM* = p AM p0

(1.2)

The so-called AM-1.5 conditions are achieved when the sun is at an angle of 41.8 above the horizon and results in the spectral distribution shown in Figure 1.3 and a solar flux of 963Wm 2. This angle of incidence is commonly encountered in western countries and hence AM-1.5 is taken as a standard condition for solar cell testing and referencing. In the last few years, the AM-1.5 spectrum has been standardized by both the International Organization of Standardization (ISO 9845-1:1992) and by the American Society for Testing and Materials (ASTM E892-87:1992), although the latter standard is more commonly referred to in respect of solar cell testing. For convenience, the flux of the standardized AM-1.5 spectrum has been corrected to 1000Wm2. However, despite availability of standards, great care has to be taken when results reported in the literature are compared. For example, often AM-1.5 conditions are reported, but this may only mean that a radiance of 1000Wm2 has been used, with the proper spectral distribution being neglected.

1.6. Solid State Solar Cells


Data from the references [3, 13-18]3 have been compiled to generate this section. The thermodynamic limits of solar energy conversion have been described by Adamson and co-workers [19]. 1.6.1. Introduction Photovoltaic effects were firstly observed more than 150 years ago. In 1839, Becquerel detected a photovoltage when sunlight was allowed to shine on one of two electrodes he had placed in an electrolytic solution [20]. In 1954, researchers at the Bell Telephone Laboratories demonstrated the first practical conversion of solar radiation into electric energy via use of a p-n junction type solar cell with 6% efficiency [21]. With the advent of the space program, photovoltaic cells made from semiconductor-grade silicon quickly became the power source of choice for use on satellites. The systems were very reliable, and cost was of little concern. In the early 1970s, the disruption of oil supplies to the industrialized world led to serious consideration of photovoltaics as a terrestrial power source. This work focused research attention on improving performance, lowering costs and increasing reliability. These three issues remain important today, even though extraordinary progress has been made over the years.

Part of the text may contain quotations from these references without specific citation.

Chapter 1: Introduction to Solar Energy Conversion Solid State Solar Cells

page: 9

Photovoltaic represents a high-technology approach to converting sunlight directly into electrical energy. The electricity obtained is direct current and can be used directly to operate direct current devices, converted to an alternating current or stored for later use. Conceptually, in its simplest form, a photovoltaic device is a solar-powered battery whose only consumable is the light that fuels it. Furthermore, there are no moving parts, the operation is environmentally benign and if the device is correctly insulated against harm from the environment, there is nothing to wear out. Because sunlight is universally available, photovoltaic devices provide many benefits that make them usable and acceptable to all inhabitants of our planet. Photovoltaic systems are modular, and so their electrical power output can be engineered for virtually any application, from low-powered consumer uses (e.g. wristwatches or calculators) to high capacity solar power plant stations. Moreover, incremental power additions are easily accommodated in photovoltaic systems, unlike more conventional approaches based on use of fossil or nuclear fuel, which require multimegawatt plants to be economically feasible. In order to illustrate the potential of solar cells, it has been estimated [22] that a photovoltaic power plant of an area of 140km2 located in an average part of the US could generate all the electricity needed by the country, assuming realizable conditions and a system efficiency of 10%. To understand the many facets associated with generation of photovoltaic power, one must understand the fundamentals of how the devices work. Although photovoltaic cells come in many forms, the most common structure is based on the use of a semiconductor material into which a large-area diode, or p-n junction, has been formed. The fabrication processes tend to be based on traditional semiconductor approaches diffusion, ion implantation, etc. Electrical current is taken from the device through a grid contact structure on the front that allows the sunlight to enter the solar cell, a contact on the back that completes the circuit, and an antireflection coating that minimizes the amount of sunlight reflecting from the device. 1.6.2. Semiconductor In a semiconductor, the electrons reside within two energetic values or bands, viz. the valence band and the conduction band. Separating these two bands is a band of forbidden states, the energy gap (Eg). According to band theory, a solid body is characterized by a unified electron spectrum in which each electron belongs to the entire body but not to an individual bond. The energy band diagram of that part of the spectrum which determines the electrical and optical characteristics of a semiconductor in the range of interest is shown in Figure 1.4.

Chapter 1: Introduction to Solar Energy Conversion Solid State Solar Cells

page: 10

e EC energy ED
Fermi level

EF

Eg=EV-EC

EV
intrinsic

EA
n-type

h+

p-type

Figure 1.4. Energy diagram of semiconductors with intrinsic, n-type and p-type properties. EC = energy of conduction band, EV = energy of valence band, EF = energy of Fermi level, ED and EA = energy of donor or acceptor dopant level, respectively

In contrast, no forbidden states exist in a metal, with the valence and the conduction bands overlapping. An insulator is usually identified4 as a material having a band gap of Eg>3eV (e.g. diamond: 5.4eV). However, this definition is not rigid, since, for example TiO2 is still considered to be a semiconductor even though it has a band gap of Eg3.2eV. At all temperatures T>0K there are always electrons with a kinetic energy larger than Eg. That is, some electrons are excited into the conduction band, which leave positive holes in the valence band. In the simplest case, the equilibrium concentrations of electrons and holes are equal, n0=p0. Both, the electrons in the conduction band and the holes in the valence band are responsible for transmission of electric current. Such an ideal semiconductor is called intrinsic. The equilibrium behavior of the electron system of a solid body is described by the electrochemical potential of electrons known as the Fermi level. In this context, the electrochemical potential represents the increase in the system's free energy when an electron is added. The thermal excitation of electrons into higher energetic levels is given by the Boltzmann distribution and allows the equilibrium concentrations of electrons and holes to be calculated by Equation 1.3:
F EC kT EV F kT

n 0 = N ce

p0 = N v e

(1.3)

where NC and NV are the effective density of states in the conduction and valence bands (and can be treated as constants), k is the Boltzmann constant (1.381023JK1) and T is the absolute temperature. The Fermi energy of an intrinsic semiconductor lies half way between the conduction and valence band energy at a temperature of T=0K. Similarly, at higher temperatures, excited electrons must be taken into account resulting in Equation 1.4:
EF = 1 N E v + E c kT ln C 2 NV

(1.4)

Additional electrons in the conduction band and holes in the valence band can be generated by deliberately adding (doping) small amounts of acceptor or donor substances (called dopants) into
4

Another classification is the level of conductivity [14]: ` conductors: > 1041cm1 ` semiconductors: 104 > > 1081cm1 ` insulators: > 1081cm1

Chapter 1: Introduction to Solar Energy Conversion Solid State Solar Cells

page: 11

the semiconductor to produce extrinsic semiconductors. Arsenic, a group V element, behaves as an electron donor when introduced into the intrinsic silicon (group IV element) lattice, since it provides an additional electron which is loosely bound and hence can be more easily excited into the conduction band than in case of intrinsic silicon. This results in a donor level near the conduction band as shown in Figure 1.4. At room temperature, most of the donor atoms will be ionized, each one yielding a conduction band electron and leaving behind isolated positively charged nuclei, whereas the number of holes in the valence band will be unaffected. Clearly, in such a material most of the electrical conductivity can be attributed to the conduction band electrons, which are thus the majority carriers. The holes, which only make a small contribution to the conductivity, are called the minority carriers. A material doped with donor atoms is called a n-type (negative-type) semiconductor. Conversely, introducing an acceptor element (e.g. gallium, a group III element) to intrinsic silicon creates an acceptor level (EA) near the valence band (Figure 1.4). In this case electrons are thermally excited from the valence band into these acceptor sites, leaving behind mobile holes which thus become the majority carriers. Electrons are now the minority carriers and the material is called a p-type (positive-type) semiconductor. For doped material, the location of the Fermi level (EF) depends strongly on the doping level, NA or ND. For moderately or heavily doped n-type solids (ND>1017cm3), EF lies slightly below the conduction band edge and similarly for p-type materials, EF lies just above the valence band edge (see Figure 1.4). 1.6.3. Light and Semiconductors When an intrinsic semiconductor is irradiated with light of energy larger than the band gap energy (hEg), an electron from the valence band can be promoted into the conduction band (see Figure 1.5).
EC energy

Eg EF

EV

+ absorption

Figure 1.5. Light absorption by an intrinsic semiconductor. The impinging photon induces an electron transition from the valence band into the conduction band, if hEg

This process creates non-equilibrium carriers in pairs, so that n=p, where n and p are the excess concentrations of electrons and holes, compared to the equilibrium concentrations (n0 and p0), and is called the intrinsic or fundamental absorption. The amount of light absorbed by a semiconductor is given by Lambert's law of extinction:

Chapter 1: Introduction to Solar Energy Conversion Solid State Solar Cells

page: 12

I = I0 ex

(1.5)

where x is the coordinate, I0 is the intensity of non-reflected light incident on the surface of the semiconductor at x=0 and is the linear coefficient of light absorption. The dimension of this coefficient is the inverse of length, thus, 1 is often arbitrarily called the depth or length of penetration and is a measure of distance over which the incident light attenuates e times. To be able to obtain electricity from a solar semiconductor device, the charged hole-electron pairs that are generated have to be separated. The natural force opposing any charge separation is charge recombination, that is the electron in the conduction band drops back into the valence band and recombines with an electron hole either by radiative, Auger or defect level (impurity) recombination. The carrier pair can be separated either by diffusion or by an electric field (migration). In a field free doped semiconductor, the recombination process is pseudo first order because there is a large excess of one type of carriers, assuming that the illumination does not change the majority carrier concentration. Thus, the recombination process can be characterized by a first order minority lifetime min (in the order of 105 to 107s), which is sensitive to crystal purity and doping density. Minority carriers are also characterized by their diffusion coefficient (D), and when considering the problem of carrier collection, it is convenient to define the majority carrier diffusion length [18] Lmin = (Dmin) (1.6)

as a measure of the distance that a photogenerated minority carrier travels in the field-free (quasineutral) region before recombining. Once the minority carriers have reached the surface, a region characterized by an abrupt ending of the semiconductor crystal structure and the presence of foreign atoms on or in the surface (2 to 10nm in thickness), they are free to participate in electron transfer reactions, either with redox species in solution (see page 18) or with the semiconductor lattice. Alternatively, they can be collected at a metal back contact, which acts as an electron sink. 1.6.4. The p-n Junction The purpose of a solar cell is to convert the energy in sunlight to electrical energy and to deliver that electrical energy to an external load. This implies that a current has to flow within the semiconductor composed of the positively charged electron holes and negatively charged electrons which the light has generated. In the preceding section, it was observed that these holes and electrons will recombine unless separated by some means. Separation can be achieved by applying an electric external field.5 However, this would defeat the purpose of the cell, since the energy delivered to the external load would come from the electricity source and not from photons. 6 An electric field needs to be generated within the semiconductor, which is independent of any external
5

Separation by diffusion is not applicable, since diffusion follows a concentration gradient, which in this case is similar for holes and electrons. 6 Semiconductor devices biased this way with an externally powered reverse bias are used as light detection devices.

Chapter 1: Introduction to Solar Energy Conversion Solid State Solar Cells

page: 13

energy source. Optically generated hole-electron pairs will be separated by this electric field and then be used to transport charge (current). Various methods exist to generate internal electric fields in a semiconductor [23]. The most common method employed in solid state solar cells is the creation of a p-n junction. A p-n junction might be visualized either theoretically as one piece of a semiconductor with one side p-doped and the other side n-doped, or practically as two semiconductor samples, one p-doped and the other ndoped, metallurgically joined to one uniform sample.

EC energy EF EC EF

EV
n-type p-type

+ ++ ++ ++ ++ ++ ++ +
n-type

EV
p-n junction
space charge region

p-type

Figure 1.6. The charge dipole and space charge region produced at an p-n junction.

When an n-doped and a p-doped semiconductor are joined into a single semiconductor crystal, the Fermi levels (chemical potential) of the n- and p-type regions must be aligned to the same level, leading to the configuration shown in Figure 1.6. At the moment of joining a diffusion gradient is established and carriers will diffuse due to the difference in concentration, that is the holes from the p-region will move into the n-region, and the electrons from the n-area will move into the p-region. This is an uphill reaction, energetically unfavorable, only driven by the concentration gradient. Diffusion currents will arise. The ionized acceptor and donor atoms, which are no longer electrically compensated since they cannot diffuse, remain behind as fixed space charges. As demonstrated in Figure 1.6, negative space charges will arise on the left hand side in the p-region and positive space charges arise on the right hand side in the n-region. Correspondingly, as occurs in a plate capacitor, an electric field is generated at the p-n junction, which is directed so that it drives the diffusing charge carriers in the opposite direction of the diffusion. Hence, the diffusion continues until an equilibrium is established, in other words until the diffusional flow is compensated by a field current of equal magnitude. The resulting (extremely large) internal field exists, even if both sides of the semiconductor are grounded. 1.6.5. Photovoltaic Processes in a Solar Cell Illumination of a semiconductor containing a p-n junction generates a hole-electron pair as shown in Figure 1.5 for an intrinsic semiconductor. However, if the hole-electron generation occurs within the carrier diffusional length (see Equation 1.6) of the p-n junction, the charges will be separated. The electron will be drawn to the n-side (the n-side is the side of the junction with

Chapter 1: Introduction to Solar Energy Conversion Solid State Solar Cells

page: 14

minimum energy for electrons) and the hole to the p-side (this side exhibits minimum energy for holes) of the p-n junction.

EB EC energy EF EC EF

EB - EP

EP
+

+ + + +

EV
n-type p-n junction p-type

EV
n-type p-n junction p-type

Figure 1.7. The p-n junction under illumination. Left: A photon induced hole-electron pair is separated by the local field of the junction. Right: The origin of the photovoltage EP. EB = Energy barrier created by the pn junction, EP = Energy equivalent of the photovoltage.

As illustrated in Figure 1.7, continuous irradiation leads to a reduction of the energy barrier (E B) created by the p-n junction which generates the desired photovoltage EP. Eventually a steady-state condition will be reached between holes and electrons generated by light and their recombination. The band gap energy determines the maximum photovoltage that can be generated, which is called the open circuit potential, Eoc or Voc. The current versus voltage characteristic of an ideal p-n junction is quantitatively given by [23] Equation 1.7:
q EP e kT

I = Is

(1.7)

where q is the elemental charge, k is the Boltzmann constant, T is the absolute temperature and Is is the saturation current of the junction. The qualitative description of a p-n junction under illumination is illustrated in Figure 1.8. (a) Short Circuit Current, Isc This is the current obtained if the solar cell is short circuited, that is there is no potential across the cell. Furthermore, the short circuit current is equal to the absolute number of photons converted to hole-electron pairs. (b) Open Circuit Potential, Voc The open circuit potential is obtained when no current is drawn from the solar cell. High values are achieved when diffusion length of charge carriers is as high as possible. doping concentration ND and NA is as high as possible. crystal volume is as small as possible, i.e. thin wafer.

Chapter 1: Introduction to Solar Energy Conversion Solid State Solar Cells

page: 15

VOC

illuminated
ISC Im Pm

-1 dark -1 1
Vm

potential [V] current [A]


1

Figure 1.8. Voltagecurrent characteristic of a solar cell. Voc and Isc are the open circuit potential and the short circuit current, respectively. Vm and Im are the voltage and current at the optimal operating point and Pm is the maximum achievable power output. Shaded area represents the maximum power output achievable.

(c) Fill Factor Optimal power output requires a suitable resistor, which corresponds to the ratio Vm/Im (Figure 1.8). Vm and Im are the voltage and current at the optimum operating point and Pm=VmIm is the maximum achievable power output. The ratio of peak output VmIm to VocIsc is called the fill factor (FF) of a solar cell and is an important quality criterion.
FF = Vm I m Voc I sc

(1.8)

The name, fill factor, is derived from its graphical representation. It indicates how much area underneath the IV characteristic curve is filled by the rectangle described by V mIm (shaded area in Figure 1.8) in relation to the rectangle VocIsc. The theoretically maximum obtainable FF is a function of the open circuit potential, the higher the Voc the higher FF. Fill factors for optimized solar cells are typically within the range of 0.60.75. (d) Efficiency The efficiency of a solar cell () is defined as the ratio of the photovoltaically generated electric output and the energy of the incident irradiant solar light.
= I m Vm FF I sc Voc = Plight Plight

(1.9)

Plight is the energy of the light shining on the solar cell and is obtained when the light intensities of the whole spectral range (compare Figure 1.3) are integrated. The present record (1998) for efficiency is almost 25% achieved by the University of New South Wales Photovoltaic Special Research Center team led by Prof. Martin Green [24]. However, commercially available solar cell modules usually operate with efficiencies of 1216%.

Chapter 1: Introduction to Solar Energy Conversion Solid State Solar Cells

page: 16

(e) IPCE-Value (Incident Monochromatic Photon-to-Current Conversion Efficiency) The incident monochromatic photon-to-current conversion efficiency is defined as the ratio between the number of collected electrons (current) and the number of incident light photons. As the name suggests, the value is commonly only measured for a specific light wavelength. Plots of IPCE versus wavelength illustrate the spectral operation range of a specific solar cell. For high performance solar cells, the IPCE value can reach unity over a large spectral section. The discrepancy between a low conversion efficiency and a high IPCE value is due to energetic losses, for example a 3eV photon absorbed by a 1eV bandgap semiconductor solar cell already will lose 67% of its initial energy at the charge generation step. The IPCE value can be calculated by the following equation:
IPCE = 125 photocurrent density [A/cm 2 ] wavelength [nm] photon flux [mW/cm 2 ]

(1.10)

1.7. Dye Sensitized Photoelectrochemcial Solar Cells


In the previous section the basic principle behind semiconductor based light-to-electricity conversion has been demonstrated. Two energetically distinctive electronic states (valence and conduction band) allow an electron to directly acquire the energy of an incident light photon and thereby transit into an energetically higher state, and thus charges are separated and a potential created. Therefore light absorption and charge separation are directly coupled and both occur within the semiconductor. In dye sensitized photoelectrochemical cells these two processes are separated. Light adsorption is achieved by the dye, also called the sensitizer. Charge separation occurs when the acquired light energy of the now excited dye is passed to the semiconductor in the form of an electron. 1.7.1. History The first reports on enhancements of photovoltaic effects by dyes can be dated back to the 19th century (see reference [25] for a review). In June 1887, about half a century after Becquerel's discovery [20] of the photovoltaic effect, Dr. Moser from the physical-chemical laboratories at the Vienna University reported [26] the first dye sensitized photoelectric effect:

Note about strengthening of photoelectric currents by optical sensibilization.


I wish to report that I could significantly increase the photoelectric currents discovered by E. Becquerel, by bathing both chlorinated, iodited or bromited silver plates in a dye solution, e.g. erythrosin.

Chapter 1: Introduction to Solar Energy Conversion Dye Sensitized Photoelectrochemcial Solar Cells

page: 17

For example, between two chlorinated silver plates the electromotoric force was 0.02, between two similarly treated but bathed plates it was 0.04 Volt. ...... using iodited plates electromotoric forces of up to ally be measured. I could achieve erythrosin. ..... The importance of these results were rapidly employed by researchers working in the area of photography, and eventually lead to development of color photography. However, it took almost a century after Moser's discovery before these effects were investigated in the context of solar energy conversion and photovoltaics. In the 1960's, the first experiments were carried out using single crystal semiconductor electrodes immersed into dye solutions. Theoretical and practical studies showed that only molecules directly adsorbed onto the electrode surface cause photovoltaic effects and that a closed packed mono-layer surface coverage was advantageous, since thicker layers prevent electron transfer from the excited dye into the semiconductor and eventually block adsorption of light. Disappointingly, photoelectric devices constructed in this way only exhibited conversion efficiencies of less than 0.5% and exhibited poor long-term stability. This was partly caused by the low absorbance that a single mono-layer of dye on a plane surface can exhibit. These devices were constructed for applications in water splitting systems (see page 4) rather than in direct electricity producing photovoltaic systems. A very valuable discussion on this topic is available in the conference proceedings of the Third International Conference on Photochemical Conversion and Storage of Solar Energy (1980) [27]. A breakthrough in conversion efficiency was reported by Tshubomura and co-workers in 1976 in Nature [28] when, instead of a single crystal semiconductor, they used powdered high porosity multi-crystalline ZnO, thereby significantly increasing the surface area of the electrode. Although the dye, rose bengal, was still adsorbed in mono-layer coverage, light absorbance was significantly increased. The cell had an energy conversion of 1.5%, which was about one order of magnitude higher than previously published efficiencies for similar types of cells. It is noteworthy Tshubomura and co-workers have already found that the iodide/triiodide redox shuttle system (see Section 1.7.5 on page 22) was outperforming other systems and was beneficial for obtaining high conversion efficiencies. Since the mid 1980's Professor Grtzel's research group at the Ecole Polytechnique Fdrale (EPFL) in Lausanne, Switzerland, has been the main driving force behind the development of dye sensitized solar cells. The cells utilize low-cost nano-porous TiO2 as the semiconductor material, which is deposited onto a conducting glass substrate. The Grtzel group also developed a solar cell based on a thin layer arrangement. Conversion efficiencies of over 10% were first published in 1993 [29]. Dye sensitized solar cells are now informally called "Grtzel cells" in many laboratories.
1 4 Volt

1 15

Volt can usu-

by bathing the iodited plates in

Chapter 1: Introduction to Solar Energy Conversion Dye Sensitized Photoelectrochemcial Solar Cells

page: 18

1.7.2. SemiconductorSolution Interface In this section, the basic theory required for understanding the functioning of dye sensitized solar cells is presented [17]. In contrast to solid state solar cells, the semiconductor employed in dye sensitized cells is in contact with a solution containing an electrolyte and redox couple, which results in the need to use a different theory [30] to the previously discussed theory which is valid for solid-solid interfaces (p-n junction, page 12). The Fermi level in the solution phase can be identified by the chemical potential and is calculated in terms of reversible potential or E0-values. While for most electrochemical purposes it is convenient to refer E0-values to a reference electrode, for semiconductors it is more intuitive to reference them with respect to the vacuum level. When the semiconductor and the solution are brought into contact, the Fermi level of both systems must become equal (see Figure 1.9) and this can occur by charge transfer between both phases. In the case illustrated in Figure 1.9, where the Fermi level of the semiconductor EF lies above that in solution (E0Ox/Red), electrons will flow from the semiconductor (which becomes positively charged) to the solution phase (which becomes negatively charged). However, such reversible behavior of a semiconductor electrode is rarely found. This lack of equilibration can be ascribed to corrosion of the semiconductor surface ( e.g. oxide formation) or inherently slow electron transfer across the interface.

ener gy

EC EF Ox Red EV
n-type semiconductor

EC EF

Esc Ox Red

pot ent ial

E0Ox/Red

EV
solution

n-type semiconductor space charge region

solution

Figure 1.9: Schematic representation of energy levels of the formation of a junction between an n-type semiconductor and a solution containing a redox couple (Ox/Red). Left: Before contact. Right: After contact and equilibration.

The excess charge is distributed over a layer near the surface, the space charge region of thickness:
kT L D = 2 0 sc q (n 0 + p 0 )
1 2

(1.11)

Here 0 and sc are the electrical permittivity of vacuum and semiconductor, respectively and q is the charge of the free electron. LD is called the Debye screening length. The charge distribution is analogous to that found in the diffuse double layer that forms in solution. The resulting electric field that forms in the space charge region is represented by a bending of the bands (Figure 1.9 right hand side). The space charge region usually has a length of 5 to 1000nm, depending on the level of doping and applied potential. The bands are bent upward with respect to the level in the

Chapter 1: Introduction to Solar Energy Conversion Dye Sensitized Photoelectrochemcial Solar Cells

page: 19

bulk semiconductor, when the semiconductor charge is positive with respect to the solution. A created hole-electron pair within the space charge region would then be separated in a similar way to the separation that occurs within the space charge region of a p-n junction. In the case of the illustrated example in Figure 1.9, an excess electron would thus be drawn toward the inside of the semiconductor, conversely, an excess hole would move toward the semiconductor surface and would be available for charge transfer reactions (compare Figure 1.2 on page 5). When the concentration of majority carriers in the space charge region near the surface is less than in the bulk of the semiconductor, the bands bend upward in the case of an n-type semiconductor (Figure 1.9). This region is called the depletion layer. If, for example, a potential would be applied across this interface so that the space charge region is composed of majority carriers at concentrations higher than in the bulk semiconductor, the bands would bend the other way to the previous case (downward in Figure 1.9). This is called the accumulation layer. A third case exists, when in the region near the surface the concentration of minority carriers exceeds the concentration of majority carriers in the bulk material. This is known as an inversion layer and is an extreme case of the depletion regime and leads to very strong band bending. The potential at which no excess charge exists in the semiconductor (EF = E0Ox/Red) is the potential of zero charge. Under these conditions no electric field and hence no space charge region exists and the bands are flat (not bent). This electrode potential is called the flat band potential, Efb. Various methods are available to determine this important characteristic value of a semiconductor. The flat band potential is commonly obtained from measurements of the differential capacity, which is then plotted in a Mott-Schottky plot (capacitance2 versus potential). Extrapolation of the linear relationship to zero capacitance2 allows the extraction of Efb-values. 1.7.3. The Dye Sensitized Semiconductor Interface The previous sections have discussed light induced carrier generation by creation of a holeelectron pair within the semiconductor via light absorption by the semiconductor. These two steps (absorption and charge separation) can be separated by employing a sensitizer adsorbed onto the semiconductor surface (see Figure 1.10 left) with the following advantages. Since both processes are now spatially separated, adverse charge recombination will be minimized. As illustrated in Figure 1.10, semiconductors with a very large band gap can be employed, the only requirement being that the excited state of the sensitizer (Dye+/Dye*) is energetically higher than the conduction band level of the semiconductor. The spectral response of such a device is thus determined by the dye. This is of particular interest, since the number of easily available semiconductors with band gaps suitable for operation in the visible spectrum (remember: hEg) is very limited. In contrast, the number of dyes (dyes intrinsically absorb light in the visible spectrum) available is substantial and by appropriate synthetic strategies these can be tuned to produce the desired properties.

Chapter 1: Introduction to Solar Energy Conversion Dye Sensitized Photoelectrochemcial Solar Cells

page: 20

e Dye+/Dye* energy EC EF Dye+/Dye

Dye+/Dye

EV

n-type semiconductor

solution n-type semiconductor

h+
solution

Figure 1.10: Electron transitions at a dye sensitized semiconductor surface. Left: Light absorption by the dye, followed by an electron injection into the conduction band of the semiconductor (charge separation). Right: Light absorption by the semiconductor results in an undesirable reduction of dye (hole injection into the dye).

The schematics of the dyesemiconductor interface are shown above in Figure 1.10. The dye absorbs a light photon and is promoted into an excited state. In case of commonly employed transition metal complexes, this originates from a metal-to-ligand charge transfer (MLCT). From this excited state7 the dye is now able to inject an electron into the conduction band, thereby becoming oxidized (Dye+). The injected electron is drawn into the semiconductor by the electric field of the space charge region. In the absence of a suitable redox mediator in solution (not shown in Figure 1.10), this photoprocess would cease when all of the dye (Dye) is consumed. Once more, a hole-electron pair has been created, the electron being in the bulk semiconductor and the hole being the oxidized dye (Dye+) isolated on the semiconductor surface. In Figure 1.10 the right hand side shows the case when highly energetic light passes through the absorbed layer of dye and directly excites the semiconductor. The semiconductor now acts like a p-n junction, a hole-electron pair is generated within the semiconductor, the electron moves into the bulk whereas the hole moves toward the electrode surface where it can reduce the dye. This normally is an unwanted reaction, since it can result in decomposition of the dye. 1.7.4. Nano-porous Titanium Dioxide As outlined above, one of the requirements of dye sensitized solar cells is the use of a large surface area semiconductor to provide sufficient light absorption with only one adsorbed monolayer of dye. The material of choice was found in titanium dioxide (TiO2, anatase). TiO2 is cheap, abundant, chemically inert and its properties have been extensively studied [31-47]. The band gap energy of anatase is 3.2eV [48], which makes it unsuitable for solid state solar cell applications (3.2eV=387nm). However, for dye sensitized solar cells this fact is rather an advantage (see above). To obtain a large surface area, nano-sized TiO2 particles are used (see Figure 1.11), which usually have diameters of 1530nm. These particles are packed into a layer of 210m thickness. The layer thickness has to be optimized in regard to light absorption of the film (where thick is advanta-

It has emerged from laser spectroscopic studies [77, 86, 93] that in the case of the Grtzel cell electron injection occurs ultrafast from "hot" excited states rather than from vibrationally relaxed states as assumed in Figure 1.10.

Chapter 1: Introduction to Solar Energy Conversion Dye Sensitized Photoelectrochemcial Solar Cells

page: 21

geous) and charge carriers average diffusion length to avoid charge recombination (where thin is advantageous).

Figure 1.11: IFESEM micrograph of nano-porous TiO2 layer. Side and top view. Layer thickness 10m, particle diameter 22nm.

For synthesis [38] of nano-sized TiO2, usually titanium alkyloxides are hydrolyzed under acidic conditions, deposited as a thin film on conducting glass substrate and sintered at temperatures of 450C for several minutes to give a final film thickness of 810m. Material obtained according to this procedure [38] exhibits very high surface areas of up to 293m2g1 [38, 49]. Particles of semiconductors of the size described above (<100nm) have different properties than macroparticles. For example, they give clear suspensions and show intriguing quantum sized effects in their absorption and luminescence properties [48]. Another effect arises when nano-sized semiconductor junctions (e.g. semiconductor/solution) are formed. Commonly band bending of the semiconductor bands relative to the bulk material is expected. As pointed out on page 18, the space charge region, where the band bending occurs, can be several orders of magnitude thicker than the actual size of the nano-sized semiconductor itself, so that no "bulk" material with uniform spatial band distribution exists to which the band bending can be related. The question arises, does band bending also occur in this case and can it be treated similarly to the bulk material? At present this important distinction appears to be often neglected in the literature and nano-sized semiconductors are treated as bulk material as shown in Figure 1.9, although it has been shown that an electrical potential gradient, which may form in nano-sized particles, is in the submillivolt range [50]. Thus, band bending is negligible in this case and no, or only a very small, electric field gradient exists in a nano-sized semiconductor to aid charge separation. The small particle size allows the bands to "float" and rapidly change their energetic position by several hundred meV, triggered by small changes in the solution phase environment [51]. This fact might be crucial to understanding the properties of dye sensitized nano-structured solar cells [52].

Chapter 1: Introduction to Solar Energy Conversion Dye Sensitized Photoelectrochemcial Solar Cells

page: 22

1.7.5. The Dye Sensitized Solar Cell In the previous section it has been shown how dyes sensitize semiconductors and why nanostructured semiconductor material is preferable. This information is combined to give a cell of the kind shown in Figure 1.12. The cell is constructed in a sandwich configuration. The working electrode is nano-porous TiO2 placed on a conducting glass slide support, and only separated by a 50100m thick layer of electrolyte solution (acetonitrile) from the counter electrode. The dye is adsorbed (usually chemisorbed) onto the TiO2 surface. The counter electrode is also made of conducting glass with a thin transparent layer of platinum sputtered onto it to reduce overpotentials. Since both electrode materials are transparent, illumination might occur from either the front (through counter electrode) or from the back (through the working electrode). The latter case is preferred, as it reduces the diffusion length of the generated carrier to the back contact.

A
e
conduction band

kinj

ker2
k0

fermi level

energy

ker1
semiconductor

Eoc h R /R

e
counter electrode


valence band

electrolyte solution

2-10m

50-100m

Figure 1.12: Schematic representation of electron transitions in a dye sensitized solar cell.k0 = rate constant for decay of excited state (Dye*Dye). kinj = rate constant for electron injection into TiO2 (Dye*eTiO2), ker1 = rate constant for electron recombination with oxidized dye (e + Dye+), ker2 = rate constant for electron recombination with redox mediator (e + R), R/R = redox mediator system, commonly I/I3.

(a) Operating Principle A light photon enters the cell and transverses it until it is absorbed by a dye molecule. The dye will then be promoted into its excited state (Dye*) from where it is now energetically able to inject an electron into the TiO2 conduction band. The electron can flow into an external circuit through a load (resistor) in order that the energy can be utilized. After this, the electron, which now carries less energy, enters the cell via the counter electrode. From the counter electrode/electrolyte interface it is transported via a charge mediator (commonly the iodide/triiodide couple, present in very high concentrations of about 0.5M/0.05M) to the working electrode. This process is diffusion controlled. The remaining oxidized dye (Dye+) on the TiO2 surface is then reduced back to its original state by the redox mediator (R + Dye+ R + Dye), which completes the cycle.

Chapter 1: Introduction to Solar Energy Conversion Dye Sensitized Photoelectrochemcial Solar Cells

page: 23

(b) Side Reactions Several reactions need to be considered which may lower the performance of dye sensitized solar cells. These reactions are: Excited state decay: Electron injection (kinj) or regeneration of Dye+ by R needs to be faster than the decay rate (k0). Electron recombination with dye+: Electron in the TiO2 conduction band recombines (ker1) with the oxidized dye (Dye+) on the TiO2 electrode surface. Electron recombination with R: Electron in the TiO2 conduction band recombines (ker2) with oxidized form of redox mediator (R). Bleaching of dye: If the oxidized form of dye (Dye+) is not stable on the time scale over which it is reduced back to its ground state (Dye), it will decompose and hence, the cell will bleach. A more detailed discussion of these side reactions will be found in the following chapters. 1.7.6. Comparison to Solid State Cells As is the case of solid state solar cells, a semiconductor is used for charge separation in dye sensitized solar cells. In solid state solar cells, an electric field is necessary for efficient charge separation, which is created by a p-n junction, whereas in dye sensitized solar cells the charge is spatially separated by utilizing a sensitizer compound. In solid state cells the charges (holes and electrons) need to diffuse in the semiconductor to the back contact, in wet cells only the majority carriers diffuse in the TiO2 semiconductor and the holes (oxidized dye) remain as fixed charges on the surface. However, the solution diffusion of the charge mediator needs to be taken into account. The most important advantages of wet solar cells are: Low cost material: TiO2 is very cheap and abundant. Solid state solar cells, especially high performance solar cells require high purity semiconductor material which is inherently expensive. The amount of sensitizer needed is low. In case of polypyridyl ruthenium complexes less than 10mg ruthenium/m2 is needed. Diffuse light performance: Solid state cells have one significant draw back, they only operate well under direct (perpendicular) irradiation, unless significant effort is made to alter the surface structure. This is partly due to the perfect materials needed, resulting in total reflection phenomena. In case of the rough TiO2 surfaces used in wet solar cells, reflection is minimized and, as a matter of fact, these cells operate best under diffuse conditions.

Chapter 1: Introduction to Solar Energy Conversion The Dyes: Octahedral Ruthenium(II) Complexes

page: 24

1.8. The Dyes: Octahedral Ruthenium(II) Complexes


1.8.1. History Most sensitizers currently used in solar cell applications are ruthenium polypyridyl complexes. Ruthenium is a silver-grayish metal first discovered by Russian chemist Karl Karlovich Klaus in 1844, after having been predicted and named by his countryman Gottfried Wilhelm Osann in 1828. Elemental ruthenium occurs in native alloys of iridium and osmium: up to 14.1 percent in iridosmine and 18.3 percent in siserskite. It also occurs in sulfide and other ores (e.g., in pentlandite of the Sudbury, a nickel-mining region of Canada) in very small quantities that are commercially recovered. The natural abundance in the Earth's crust is 1ppb and its cost (crude) is 1US$/100g. Ruthenium has a [Kr]4d75s1 electron configuration, belongs to the group of the platinum metals (Ru, Rh, Pd, Os, Ir and Pt) and has very similar chemistry to osmium, but also shares many characteristics with iron and technetium. Importantly, ruthenium also can form many organometallic and coordination complex compounds. The development of sensitizers over the past two decades is summarized in Figure 1.13. Over this period, more than 900 sensitizers have specifically been synthesized and tested for solar cell applications [53], but only those few displayed in Figure 1.13 performed satisfactorily. Most of these sensitizers were ruthenium complexes, some were osmium [54] complexes, pure organic dyes [28] or other compounds. So far, it is not understood, why only a very small fraction of dyes perform well as photosensitizers, whilst the majority of dyes, with very similar chemical and physical properties, do not. The lack of structure/performance understanding was one of the major motivations for undertaking studies described in this dissertation.
HOOC COOH HOOC COOH

N N N Ru

N N N HOOC N

N Ru N

N N N Cl COOH HOOC N

N Ru

N N OH Cl N Ru N N N C Ru C N N N Ru N N C N COOH N COOH COOH N N COOH N C

HO

COOH

1
[RuII(bpy)3]2+

HOOC

COOH

3
RuII(H2-dcbpy)2X2

[RuII(H2-dcbpy)3]2+

COOH S C N S C N N C S Ru N N N COOH HOOC

HOOC

COOH

HOOC

COOH

N N N C S COOH Ru

N N N C S N COOH HOOC N

N Ru C

N N C N COOH

[NC-RuII(bpy)2-CN-RuII(H2-dcbpy)2-NC-RuII(bpy)2-CN]2+

6
RuII(H2-dcbpy)2NCS2

5
RuII(H2-dcbpy)2CN2

[RuII(H3-tctpy)NCS3]-

Figure 1.13: Historic development of sensitizers for solar cells.

Chapter 1: Introduction to Solar Energy Conversion The Dyes: Octahedral Ruthenium(II) Complexes

page: 25

Details on the development of the sensitizer in dye sensitized solar cells (Figure 1.13) are as follows: (1) [Ru(bpy)3]2+:8 [Ru(bpy)3]2+ is one of the best studied photosensitizers, partly due to its successful application in homogeneous and heterogeneous water splitting systems [4, 55-58] (Figure 1.2). It has never been routinely used as a sensitizer in solar cells, which might be attributable to its ability to only physisorb onto the semiconductor surface [59]. However, since it is a convenient model compound, literature on investigations of [Ru(bpy)3]2+ in solar cell like applications are available [59-61]. An excellent review on this complex and its photochemistry can be found in reference [62]. (2) [Ru(H2-dcbpy)3]2+ [50, 63]: Introducing carboxylate groups onto the ligand results in a favorable red shift of the absorption spectrum. Complexes, where one or more of the bpy ligands of [Ru(bpy)3]2+ have been replaced, were first synthesized by the group of Wolfgang Sasse in Australia in an attempt to obtain a more efficient sensitizer in water splitting systems [64-66]. In the case of nano-porous TiO2 electrodes, this compound is able to form ester linkages with the surface hydroxy groups of the TiO2, thus allowing very efficient attachment onto the surface via formation of chemical bonds [67-69]. (3) Ru(H2-dcbpy)2(OH)2 [70]: The replacement of one H2-dcbpy unit by two negatively charged hydroxyls or chlorides shifts the absorbance bands of the complex by over 100nm toward the red, thereby increasing the spectral response. (4) [NC-Ru(bpy)2-CN-Ru(H2-dcbpy)2-NC-Ru(bpy)2-CN]2+ [71, 72]: With this trimeric complex, a new area of sensitizers started, where two bipyridine derivatives and two pseudo halides, which are able to maintain a strong -back bond with the metal center, are coordinated to ruthenium. This is also the sensitizer used by O'Regan and Grtzel in their work published 1991 in Nature [72]. Achieved conversion efficiencies with use of this dye are around 7.5%. (5) Ru(H2-dcbpy)2(CN)2 [54, 71, 73-75]: Developed at about the same time as the latter trimeric sensitizer and exhibits very similar efficiencies. (6) Ru(H2-dcbpy)2(NCS)2 [29, 38, 72, 76-87]: First reported in 1993 [29] and currently (1999) still the most commonly used sensitizer. Provides a solar conversion efficiency over 10%. Use of the ambidentate thiocyanate (NCS) ligand gives rise to three different thiocyanate linkage isomers [80], which cannot be completely separated. The synthetically easily available double N- bond isothiocyanato complex is mainly used. The majority of work reported in this thesis is concentrated on this complex.

One of the very first dyes employed as a sensitizer in nano-porous solar cells was of course rose bengal used by Tsubomura and co-workers [28].

Chapter 1: Introduction to Solar Energy Conversion The Dyes: Octahedral Ruthenium(II) Complexes

page: 26

(7) [Ru(H3-tctpy)(NCS)3] [88]: A recently reported dye that outperforms Ru(H2-dcbpy)2(NCS)2. Several attempts may be found in the literature [89, 90] to vary the structure of the bipyridine ligand and to replace one or two bipyridines by terpyridine. Chapter 6 gives details of this compound. 1.8.2. Octahedral Complexes (a) Valence Bond Theory In general, a molecular orbital is constructed from atomic orbitals of the appropriate symmetry. However, instead of describing an orbital in terms of a linear combination of all possible atomic orbitals, it is sometimes convenient to form a mixture of orbitals on one atom and then use the hybrid orbitals to construct localized orbitals [91, 92]. In CH4 for instance, each CH bond can be regarded as formed by the overlap of an H1s orbital and a hybrid orbital composed of C2s and C2p orbitals, thus forming four equivalent sp3-hybrid orbitals. To form hexacyano ruthenium [Ru(CN)6]4, the first three of the five d-orbitals will be doubly occupied, so that two d-orbitals will become vacant. Subsequently these two 4d-orbitals will be hybridized with the 5s- and 5p-orbitals. Each cyanide molecule donates an electron pair for bonding to the coordination center, thus resulting in six equivalent d2sp3-hybrid orbitals (Figure 1.14).
4d Ru2+ [Ru(CN)6]45s 5p

d2sp3 (Octahedron)

Figure 1.14. Orbital hybridization in case of an octahedral ruthenium complex (VB-theory, adopted from [91]).

(b) Molecular Orbital and Ligand Field Theory Ligand field theory, which is a particular application of molecular orbital theory [91], provides a simple conceptual model to interpret electronic spectra [92]. In an octahedral complex each ligand has a single valence orbital directed toward the central metal atom with local symmetry (rotational symmetry) with respect to the ML axis. As illustrated in Figure 1.15, in an octahedral d-metal complex, the dz2 and dx2y2 orbitals (degenerated or hybridized into two eg orbitals) point along the axes directly toward the ligands and thus are repelled more strongly by the negative charge of ligands than electrons in the dxy, dzx and dyz orbitals, which are triply degenerated into three equal t2g orbitals. Due to the energetically unfavorable configuration of the eg subset, it is increased in energy, concurrently the energy of the t2g subset is lowered, resulting in zero net change. Both subsets are split by 0, called the ligand field splitting parameter.

Chapter 1: Introduction to Solar Energy Conversion The Dyes: Octahedral Ruthenium(II) Complexes

page: 27

ligand

z
ligand ligand

ligand

complex

ligand

y x
ligand ligand ligand

eg

5p 5s antibonding

eg LUMO

dz2

dx2-y2
4d

eg + t2g

t2g HOMO

nonbonding

t2g
eg
bonding

dxy

dzx

dyz

Figure 1.15. Splitting of the d-orbitals in the octahedral ligand field into two energetically non-equivalent subsets, t2g and eg. 0 = octahedral ligand field splitting parameter. Shaded area (right) shows frontier orbitals (adopted from [92]).

(c) back-bonding If the ligands in a complex have orbitals with local symmetry with respect to the ML axis, then they may form molecular orbitals with the t2g metal orbitals [92]. A -acceptor ligand means a ligand that (usually) has filled -orbitals at lower energies than the metal t2g orbitals and low energy empty orbitals that are available for occupation. Typically, the acceptor orbitals are vacant antibonding orbitals on the ligand, as in CO and bipyridine. If these vacant * ligand orbitals are close in energy to the metal t2g orbitals and the metalligand overlap is strong, some electron density will be delocalized from the metal to the ligand.
ligand complex ligand

dxy (Metal)

py (Ligand)
x

5p 5s antibonding

eg LUMO
0

(empty)

eg + t2g
4d

t2g HOMO
(filled)

eg
(filled)

bonding

Figure 1.16. -back bonding in an octahedral metal complex.

As illustrated in Figure 1.16, this type of bonding results in a decrease in energy of the metal character t2g orbital and as a net result 0 is increased by the -acceptor interaction. Conversely, if a ligand has filled orbitals and the metal t2g orbitals are predominantly empty, bonding between the

Chapter 1: Introduction to Solar Energy Conversion The Dyes: Octahedral Ruthenium(II) Complexes

page: 28

filled and partly empty t2g orbitals can be established, resulting in a decrease of the 0 value. These kind of ligands are called -donor ligands (such as Cl, I and H2O). Thus, ligands which are able to form strong bonds and are good acceptors form very stable complexes when coordinated to a metal center. Schematic molecular orbital diagrams, as shown in Figure 1.15 and 1.16 (right), are a very valuable tool for interpretation of electronic spectra. And, vice versa, electronic spectra can provide much insight into orbital structure and bonding of complexes. For instance, the ruthenium bipyridine complexes used in dye sensitized solar cells are dominated by very strong MLCT transitions in the visible spectrum. Such transitions involve an electron from the metal t2g orbital to be promoted into the empty * orbital of the bipyridine ligand and results in a momentary spatial charge separation. Hence, its name "metal-to-ligand charge transfer" is derived. The intensity of this band in the spectrum reflects the quality of orbital overlap between the metal and ligand orbitals (see Figure 1.16 left), whereas the position of the transition is a measure of the energy between both orbital levels.9

1.9. Aim of the Work


The sensitizer compound forms an essential part of a dye sensitized solar cell. It determines its absorption characteristics and spectral response. As illustrated in Figure 1.12, the value for the formal potential of the dye/dye+ process is an essential thermodynamic value for describing and understanding the properties of the cell. Photophysical studies [68, 77, 86, 93, 94] on these kinds of sensitizers have been extensive. However, general electrochemical studies characterizing the nature of the oxidized states of those complexes, are yet to appear. This is surprising, since much effort has been put into the syntheses of new sensitizers and the reasons for success of only a very few (see Figure 1.13) is not understood. Similarly, the role of the redox mediator is also not understood. Since Tsubomura and co-workers' first report [28] of a nano-crystalline dye sensitized solar cell in 1976, the iodide/triiodide redox system has been used. Since then, no other redox system has ever been reported to be used as a charge mediator in these kinds of solar cells. Similarly, no studies may be found in the literature about alternative systems. The above limitations in the knowledge base on dyes and electrolytes prompted me to conduct electrochemical and spectroscopic studies on the sensitizer and to investigate the role of the redox shuttle, in order to gain a better understanding of the impact of these compounds on the overall performance of wet solar cells and enable predictions to be made about the preferable physical/chemical properties of novel sensitizers and redox shuttles. The following is a brief research outline of the content of the thesis:

For a true measure of orbital separation the E0-0 transition has to be taken, which can be estimated from emission measurements.

Chapter 1: Introduction to Solar Energy Conversion Aim of the Work

page: 29

Model Compounds: Firstly, model compounds were found, which are chemically and electronically similar to the types of sensitizers used in photoelectrochemical solar cells, but which exhibit simple electrochemical behavior. That is, their oxidized and reduced states are sufficiently stable to allow detailed electrochemical and spectroscopic characterization.

Sensitizers: The currently best performing sensitizers, Ru(dcbpy)2(CN)2, Ru(dcbpy)2(NCS)2 and [Ru(tctpy)(NCS)3], have been electrochemically and spectroscopically characterized. A major task has been to establish an accurate value for the formal potential of the redox processes, specifically for the important oxidation processes of the sensitizers. Data obtained from studies on the model compounds can then be conveniently used for comparison or to predict physical properties of the sensitizers, which are not as readily accessible.

Redox Shuttles: Several possible redox shuttles, such as the ferrocene/ferrocenium system, have been tested for performance and compared to the established iodide/triiodide system under cell-like conditions. An apparatus has been developed, which allows rapid and uncomplicated measurement of cell performance under varying electrolyte conditions. This required the construction of a flow system with a thin-layer optically transparent flow cell consisting of a TiO2 working electrode, in order to mimic actual solar cell conditions.

Chapter 2

Chapter 2: Experimental and Theoretical Details History

page: 31

The theory and application of electrochemical techniques are well documented. For example, detailed overviews can be found in books by Bard and Faulkner [95], Brett and Brett [96], Hamann and Vielstich [97, 98], and Rieger [99]. Oldham and Myland [100] provide an excellent introduction to theoretical aspects of electrochemistry and Kissinger and Heinemann [101] discuss practical applications of electroanalytical techniques. Since a range of electroanalytical techniques form an essential part of results reported in this thesis, a short introduction is given in this chapter to the theoretical foundations and practical applications.

2.1. History
The first electric effects were discovered by Luigi Galvani (17371798), who was a physiologist at the University of Bologna [99, 102]. In the late 18th century he found that animal muscles contracted when they came in contact with copper and iron. The field of "animal electricity" attracted the attention of Alessandro Volta (17451827), Professor of Physics at the University of Pavia. In 1793 he established the series of standard electrode potentials for metals. Sir Humphry Davy (17781829) was a well educated man who became in 1801 at the age of twenty three Lecturer of the Royal Institution, and later in 1812 Professor of Chemistry and Director of the Chemical Laboratories of the Royal Institution. At this time the Royal Institution was the home, and almost the only center in England, of systematic scientific research. He was fascinated by the discoveries of Volta and the accidental observation in 1800 of the decomposition of water with the aid of a battery by Nicholson and Carlisle. In 1807, after several years of research in the field of electrochemistry, Davy was led to the discovery of the metallic elements potassium and sodium obtained by electrolysis of their carbonate salts. In 1813 before retiring from active work, he appointed Michael Faraday (17911867) as a laboratory assistant. Faraday was the son of a blacksmith and he describes his education as [103]: "My education was of the most ordinary description, consisting of little more than rudiments of reading, writing and arithmetic at a common day school. My hours out of school were passed at home and in the streets"

Figure 2.1. Michael Faraday and his laboratory at the Royal Institution [103].

Chapter 2: Experimental and Theoretical Details History

page: 32

Faraday used to work as a book binder which allowed him to study books, such as Marcet's Conversations in Chemistry, the Encyclopdia Britannica and Watt's On the Mind. His attention was turned to science by the article Electricity in an encyclopedia he was employed to bind. Faraday quickly advanced in the Institution becoming a lecturer. His work was much more fundamental than Davy's, he defined the term electrode, anode, cathode and electrolysis and in 1831 he discovered electro-magnetic induction. Today his name is unalterably attached to electrochemistry; one of the fundamental laws of electrochemistry, the relation of current to mass, bears his name as does its fundamental constant.

2.2. Definitions and Basic Units


2.2.1. Current and Potential "At the heart of electrochemistry lies the coupling of chemical changes to the passage of electricity" [100]. Electricity is often treated as if it were a fluid which is able to flow through substances [100], called conductors. The fluid consists of electrical charges, which are related directly to electrons.10 Often, comparison with the flow of water has been used to explain the phenomenon of electricity. Thus, as is the case with water where water molecules flow11, in the case of electricity electrical charges or electrons flow. Water which is pumped through a hose will flow at a certain flow rate (amount of water per time unit) and will have a certain pressure at the outlet. In an electrical generator, the electrons flow through a circuit. Two leads are necessary to establish a flow, one for the outlet, one for the inlet. Unlike the water analogy, due to charge neutrality constraints, for each electron that leaves the generator one has to return. The amount of electrons flowing in the leads is called the current and its units are coulomb per second or ampere [1Cs1 = 1A]. The pressure when applying the water analogy corresponds to the voltage or potential difference measured between both leads. The unit for an electrical potential is the volt or joule per coulomb [1V = 1JC 1] and therefore the product of current and potential gives the energy flow or power [1AV = 1Js 1 = 1W]. 2.2.2. Ohm's Law The fundamental relationship governing the conduction of electricity is Ohm's law. This law states, that when electricity flows through most electrical conductors, that the current passing through a sample of conducting material is proportional to the potential difference across it. = RI (2.1)

10 11

One electron is the carrier of one elemental charge unit, q=1.602191019C [242]. A flow is defined as the change of something per time unit. In this case the flow of water would be correctly expressed as a mole-flow, mols1. However, the more familiar volume-flow [liter s1] can be easily calculated from the mole-flow.

Chapter 2: Experimental and Theoretical Details Definitions and Basic Units

page: 33

In Equation 2.1, stands for potential [V], I is the current [A] and R is called the resistance [1 = 1VA1]. As we shall see later, the resistance, whether Ohmic or not, plays a crucial role in electrochemical measurements. 2.2.3. Capacitance If an electrical circuit includes a pair of parallel conducting plates separated by a narrow gap containing air or an other insulator, electricity will flow across the plates as illustrated in Figure 2.2 when the current source is switched on.
c urre nt s o urc e

I
+++++++++++

Figure 2.2. A capacitor in an electrical circuit.

In this case, charge accumulates on the surfaces of the plates, so that the plates develop a potential difference E. The amount of stored (passed) charge is proportional to potential. Q = CE (2.2)

The proportionality constant linking Q and E is negative and its absolute value is known as the capacitance. The capacitance unit is farad [1F = 1s]. The capacitance of a certain gapplategas unit depends on its width (l), area of the plates (A) and on the permittivity () of the filling gas or medium. The ratio of the permittivity of a material to the permittivity 0 of the vacuum is known as the dielectric constant of the material.
Q E =C= A l

(2.3)

The relevance of this model to electrochemical experiments will be revealed later in this chapter. 2.2.4. Redox Reactions Redox reactions involve electron transfer from one compound to another. That is, in a redox reaction one compound will be oxidized whereas the second one will be reduced. The name redox is derived from: Redox = Reduction/Oxidation. The reaction of a sodium atom with a chlorine atom is a classical example of a redox reaction:

e
Na Cl Na+ + Cl (2.4)

Chapter 2: Experimental and Theoretical Details Definitions and Basic Units

page: 34

Sodium is oxidized (loss of electrons) and acts as an electron donor or reductant, chlorine is reduced (uptake of electrons) and acts as an electron acceptor or oxidizer. Many reactions are redox reactions, although this fact may not always be as clear cut as the example shown in Equation 2.4 Redox reactions are identified by a change in oxidation number in at least two of the participating atoms or moieties that participate in the reaction. Other major types of reactions encountered in chemistry are coordination reactions or acid-base reactions. In these types of reaction, no distinctive electron transfer occurs between two centers, rather only a change of electron densities. These reactions are characterized by the electron pair donor (also called LewisBase or nucleophile) and the electron pair acceptor (also called LewisAcid or electrophile) and no formal change in oxidation number of any of the participating atoms takes place. 2.2.5. Electrochemical Redox Reactions Electrochemistry is the field of chemistry that deals with redox reactions, in which the electron donor or the electron acceptor is an electrode. In the case of the reaction of sodium with chlorine (Equation 2.4), the electron for the reduction of chlorine to chloride is provided by the sodium atom. In case of an electrochemical reduction of a chlorine atom, the reaction could be rewritten as: eelectrode + Cl Cl (2.5)

The electrode in an electrochemical reaction is the part of an electrical circuit where physical contact is made with the chemicals of interest and where charge transfer reactions take place. (There are electrochemical processes where no chemical charge transfer takes places, e.g. adsorption, but these will be considered later.) Thus, in a "normal" electrochemical reaction, a charge transfer reaction takes place between an electrode and a chemical species. The basic arrangement required for an electrochemical reaction to occur is shown in Figure 2.3.

+ + Na Na+ + Cl Na+ + Cl Cl Na+ + Na+ Cl + Cl Na+ Cl Na+ Na+ + Cl Na+ Cl Cl


Cl

Figure 2.3. An example of a basic electrochemical experimental arrangement.

where  symbolizes the device to measure the current, called an ammeter and

is the

symbol for a voltmeter, which measures the potential between two defined points. In Figure 2.3

Chapter 2: Experimental and Theoretical Details Definitions and Basic Units

page: 35

represents the current source. In the experiment shown in Figure 2.3, a solution of NaCl is used, which dissociates into positively charged Na+ and negatively charged Cl ions. When the two electrodes are immersed into the solution and the current source is switched off (interrupted), no reaction will occur. However, immediately when the current source is switched on and a potential is applied, electrons will flow to generate negatively and positively charged electrodes. This charge separation creates a potential distribution between the two electrodes. Consequently, ions will be attracted to the oppositely charged electrode to compensate for the build up charge in order to achieve charge neutrality. The movement of the ions is called migration and occurs when an electric field gradient is present. Under the conditions of Figure 2.3, the solution acts like the capacitor shown in Figure 2.2. Thus, if the potential difference is sufficiently small so that no charge transfer occurs, capacitive current will cease to flow when the capacitor has accepted all the charge it can store (Equation 2.2). When a sufficiently large potential difference is reached (see chapter 2.3), electrons will be transmitted from the negatively charged electrode to the sodium ions and sodium metal will be formed according to Equation 2.4. Similarly, at the positively charged electrode, chloride ions will be oxidized to chlorine. The current that flows during these charge transfer reactions is called Faradaic current and the processes are called Faradaic reactions, since they are governed by Faraday's laws (see chapter 2.3). The electrode where an oxidation process occurs (electrons transferred from a substance to the electrode) is called the anode. Conversely, a reduction process (electrons transferred from the electrode to a substance) occurs at the cathode. Electrode materials commonly are inert conducting metals, such as gold, platinum or carbon (glassy carbon, carbon paste, etc.). The redox active substances of interest are usually present in the solution phase but electrochemical reactions may be performed with electrodes and substances in the gas, liquid or solid state. The only requirement to perform an electrochemical experiment is adequate conductance, so that the electrical circuit is complete at all points and charge can flow. 2.2.6. Three Electrode Experimental Arrangements The experimental arrangement shown in Figure 2.3, contains two large electrodes and is very frequently used for coulometric analysis in electrolysis, electrosynthesis or galvanization experiments when the charge passed or the number of electrons transferred is of more interest than the potential applied. However, in many electrochemical experiments employed in this thesis, the value potential is of great significance. Furthermore, commonly only processes occurring at one electrode are of interest. In this case, the electrode where the reaction of interest occurs is called the working electrode. In order to be able to measure potential at the working electrode with high precision, the three electrode arrangement, schematically shown in Figure 2.4, is used.

Chapter 2: Experimental and Theoretical Details Definitions and Basic Units

page: 36

(diaphragm) counter electrode


no current flow output (e.g. X-Y recorder)

V
feedback loop

electricity source

reference electrode

working electrode

potentiostat

Figure 2.4. An electrochemical experiment under potentiostatic control.

The potential between the reference electrode, and the working electrode is measured. In order to minimize ohmic drop, the reference electrode is positioned in close vicinity to the working electrode. The reference electrode is used to reference the applied potential against the working electrode and ideally no current flows between these two electrodes. The working electrode has a known geometrical shape (i.e. disk, hemisphere, etc.) and its area is relatively small, in order to minimize the amount of substance electrolyzed at its surface compared to that remaining in bulk solution. Under these conditions no appreciable bulk solution concentration change occurs during the course of an electrochemical experiment. The current in the arrangement in Figure 2.4 flows between the working and the third electrode, the counter or auxiliary electrode. The area of the counter electrode is sufficiently large to avoid limitations in charge transfer reactions. Reactions taking place at the counter electrode are not of interest nor is the actual potential between the counter and working electrodes. Therefore, the counter electrode can be separated from the bulk solution by a diaphragm (if required), to prevent mixing of the unknown products formed at the counter electrode and species in the bulk solution. A potentiostat is used to generate a specified potential between the working and reference electrodes. This is achieved via a feedback loop where the potential is measured and if not matching the selected value, the current is changed until the desired potential is reached.

2.3. Theory
2.3.1. Thermodynamics: The Nernst Equation General reading can be found in references [104] and [99]. Consider again the redox reaction between sodium with chlorine gas Na + Cl2 Na+ + Cl (2.6)

If the sodium and chlorine molecules are separated into separate compartments, then each compartment forms a half-cell element. If current allowed to flow the reaction would proceed as

Chapter 2: Experimental and Theoretical Details Theory

page: 37

written in Equation 2.6. Moreover, the overall reaction can be subdivided into the individual halfcell reactions. Na e Na+ + e (2.7) (2.8)

+ Cl2 Cl

More generally a reaction of this kind may be written as Red1 + Ox2 Ox1 + Red2 (2.9)

where in Equation 2.9 the index "1" refers to sodium and "2" to chlorine. The change in Gibb's energy is given by G = H TS (2.10)

(H = change in enthalpy, T = temperature and S = change in entropy) and assuming that the reaction is reversible and that an electrical work, welec, is necessary to move the electron from the sodium to the chlorine gives G = welect. If a potential difference is required to drive the reaction and an electron carries the charge q, then welect = nFE (F = qN is the Faraday constant, the charge of one mole of electrons, 96485C and N is the Avogadro constant.) so that G = welect = nFE (2.11)

According to the laws of chemical thermodynamics, the Gibb's Free energy for the reaction in Equation 2.9 is
G = G 0 + RT ln c Ox1c Re d 2 = RT ln K eq c Re d1c Ox 2

(2.12)

where R is the gas constant, T the absolute temperature, c are the concentrations of the participating species under ideal conditions and Keq is the equilibrium constant. If we take now Equations 2.11 and 2.12 we get
E= G 0 RT c Ox1c Re d 2 RT c Ox1c Re d 2 ln = E0 ln nF nF c Re d1c Ox 2 nF c Re d1c Ox 2

(2.13)

where
E0 = G 0 nF

(2.14)

and E0 is called the standard potential. Equation 2.13 is called the Nernst Equation and applies the standard temperature of 25C so that it can be transformed into the more commonly used form where the logarithm to the base 10 is used:

Chapter 2: Experimental and Theoretical Details Theory

page: 38

E = E0

c c 0.0592 log Ox1 Re d 2 n c Re d1c Ox 2

(2.15)

The Nernst Equation is a fundamental equation electrochemistry which shows that when a certain potential is applied, how the concentration quotient will adjust or how, when a certain concentration quotient is achieved, the potential will adjust under equilibrium conditions. The standard potential, E0, is an important thermodynamic value and electroanalytical techniques can often be employed to extract this value. The absolute E0 value for a half cell reaction is often not accessible and therefore values are often reported relative to the potential of a known reaction such as the oxidation of gaseous dihydrogen in acid aqueous solution (H2 2H+ + 2e) under standard conditions. This particular half-cell element is called the standard hydrogen electrode (SHE) which can be used as a reference electrode (Figure 2.4). Reference [105] demonstrates how to convert the potential of the SHE to the absolute electrode potential. However, for convenience, other half-cell reactions are used (e.g. Ag Ag+ + e) in experiments where the potential against the SHE is accurately known. The main requirement for a reference electrode is that the concentration quotient of the half-cell reaction is known so that the reference potential is defined by the Nernst equation. Use of standard potentials requires the use of activities rather than concentrations. As a result, the formal potential, E0', has been introduced, which is obtained under conditions where the concentration rather than activities quotient of the Nernst Equation equals 1. The formal potential is used to approximate the standard potential. In most voltammetric experiments, the reversible halfwave potential, Er1/2, rather than the E0' or E0 is quoted. The Er1/2 is measured from the voltammetric currentpotential curve. Reversible means, the rate of electron transfer between the species and the electrode is very fast, so that measurements are made under equilibrium conditions [95]. Er1/2 is also an approximation of the formal and standard potential. 2.3.2. Mass Transport: Fick's law Let us assume that species A+ is present in solution and that it can be electrochemically reduced in a reversible manner to A. If a potential more negative than E0 is applied so that A+ is reduced to A, the appropriate concentration quotient will be formed according to the Nernst Equation. However, under voltammetric conditions the equilibrium concentration will only be formed instantaneously at the electrode surface. All A+ ions other than those initially present at the electrode surface need to travel to the electrode surface in order to participate in the equilibrium reaction. Thus, the rate of mass transport determines the actual current (charge passed per unit time) measured in an experiment. Species can be transported by three different kind of mass transport mechanisms:

Chapter 2: Experimental and Theoretical Details Theory

page: 39

(a) Migration If charged species are present in a solution containing two electrodes, then because of electrostatic forces, positively charged particles will be drawn to the cathode and negatively charged particles to the anode (Figure 2.3). The impact of this type of mass transport is suppressed in many electrochemical experiments by addition of a large excess of an inert supporting electrolyte, which itself consists of ions, so that any contributions from migration of the analyte can be neglected [101]. (b) Convection Convection is the term used to describe mass transport by mechanical movement and can be achieved by stirring, vibration, flow, heat induced motion, etc. Convection represents the faster rate of mass transport and is used to study rapid processes. When well defined convective mass transport conditions can be achieved, mathematical modeling is possible. Convective mass transport is also used in wall-jet electrodes [106], thin layer electrodes [107], rotating disk electrodes [108], etc. (c) Diffusion Diffusion occurs in response to a concentration gradient and is expressed by Fick's first law:
J = D c x

(2.16)

J is the flux of species [mol cm2s1],

c is the concentration gradient and D is the diffusion x

coefficient [cm2s1]. The variation of concentration with time for a one-dimensional system is given by Fick's second law:
c 2c = D 2 t x

(2.17)

In the absence of migration and convection, these equations describe the mass flux to and from the electrode and hence for example allow a quantitative interpretation of the current response to be obtained after a change is made in potential. Fick's laws are the starting point for the mathematical derivation of many electroanalytical responses since many electroanalytical experiments are conducted under diffusional conditions.

2.4. Linear Sweep and Cyclic Voltammetry


The term voltammetry is used to describe electroanalytical methods in which the current is recorded as a function of applied potential. The potential may be changed with time, according to a specified waveform.

Chapter 2: Experimental and Theoretical Details Linear Sweep and Cyclic Voltammetry

page: 40

2.4.1. Electrodes Used in Voltammetric Experiments Working Electrode: The dropping mercury electrode is based on use of a stream of mercury that flows under gravity through a fine glass capillary and forms a well defined small drop at the outlet of the capillary, until it drops off, after which a new drop is formed [98]. This electrode has the advantages of providing a renewable and highly reproducible electrode surface, but the disadvantages of a limited anodic potential range (oxidation of the electrode material) and the toxicity associated with mercury. Over the past decades, the dropping mercury electrode has been replaced by solid electrodes, made from inert conducting materials such as glassy carbon, gold or platinum disks which can give well defined planar surfaces. Electrodes made from these materials have a wider positive potential range than Hg. However, they require frequent cleaning of the surface by polishing to remove adsorbed species, in order to obtain reproducible results. For transient techniques, disk electrodes with diameters between 1 and 5mm are commonly used. In contrast, microdisk electrodes usually have diameters between 0.5 and 100m. Reference Electrode [109]: A stable and reversible half-cell process is used for the reference electrode. The half-cell element is separated from the bulk solution by a glass frit, or membrane, to avoid mixing of both compartments, but still to allow charge transport. The saturated calomel electrode is employed commonly in studies in aqueous solutions and utilizes the Hg2Cl2 + 2e 2Hg + 2Cl reaction. In organic solvents, commonly the Ag/Ag+ redox couple is employed, which is achieved by dipping a silver wire into a Ag+ solution of known concentration using the same solvent as the bulk solution. If no silver salts are soluble in the specific solvent (e.g. CH2Cl2), a Ag/AgClwire in a solution with known chloride concentration might be taken or alternatively, a reference electrode consisting of a different solvent (e.g. CH3CN) can be used. Some solvents, like dimethylformamide, react with Ag+, which restricts the construction of a stable Ag/Ag+ half-cell element in the particular solvent, so that a reference electrode in a different solvent needs to be utilized. It has become common to quote potentials relative to well known processes [110], such as the ferrocene/ferrocenium (Fc/Fc+) redox couple. The potential of this couple is measured under identical conditions to that used for the analyte of interest. Under this scenario, the reference electrode needs only provide a stable potential for the duration of an experiment and the calibration cycle versus the reference substance. This approach avoids many errors (e.g. existence of unknown potentials, such as liquid/liquid potentials arising at the reference electrodebulk solution interface) and has been utilized in this work. Counter Electrode: Reactions at the counter electrode are ignored, as is its potential (see page 36). Inert materials such as platinum are used as counter electrodes. Meshes or similar geometries may be chosen to provide a sufficiently large electrode area, that should be greater in size than the working electrode.

Chapter 2: Experimental and Theoretical Details Linear Sweep and Cyclic Voltammetry

page: 41

2.4.2. Solvents When the technique of polarography or voltammetry was initially developed, the solvent medium of choice was invariably water, since at that time the apparatus could only function properly when the resistance of the solution was very small. Water has one big disadvantage, that is, it has only a theoretical operation range of 1.23V, the energy needed to split water into hydrogen and oxygen (Chapter 1.4). This potential range can be extended slightly by choice of the right electrode material (see above). The limited solubility of many organic compounds in water means that organic solvents are sometimes preferred for electrochemical use, their only major disadvantage being their high specific resistance, which can significantly distort voltammograms from their theoretical response. The use of supporting electrolyte minimizes the ohmic drop as well as migration problems. Non-aqueous solvents provide operational potential ranges of up to six volts [111]. Problems associated with organic solvents are impurity contaminations, e.g. residual water, which can significantly lessen the theoretical solvent potential range, and require tedious purification. Selection criteria for choice of solvent are inertness (wide potential range), solubility of analyte and toxicity. Reviews on non-aqueous solvents for electrochemical use have been published by Mann [111] and Fry [112]. 2.4.3. Supporting Electrolyte The role of the supporting electrolyte is to lower the resistivity of the solvent and thereby to reduce distortion of the voltammetric signals, and to suppress migration of charged electroactive analytes (see above). The supporting electrolyte must be ionic and electrochemically inert. In organic solvents, commonly tetraalkylammonium (R4N+) salts are used. This type of cations is very difficult to oxidize and to reduce and thus can be regarded as inert in the electrochemical sense in most solvents. The length of the alkyl (R) chain determines the solubility in organic solvents. Inert perchlorate (ClO4), hexafluorophosphate (PF6) or tetrafluoroborate (BF4) anions usually form the anion component of the electrolyte. 2.4.4. Linear Sweep Voltammetry In linear sweep voltammetry, the potential is linearly changed (swept) with time from the starting, Estart, to the end potential, Eend. The rate of potential change with time (
dE dt

) is called the scan

rate. Commonly Estart is selected where no Faradaic processes take place and Eend is set to a potential sufficiently past the E1/2 of the investigated redox reaction. In the case of an oxidation (the analyte being present in its reduced state), the Eend>Estart and the scan will proceed to more positive potentials (anodic scan). When Eend<Estart, a reduction process is usually being investigated and the scan will go to more negative potentials (cathodic scan). Linear sweep voltammetry is mainly employed in analytical chemistry. Slow scan rates (<10mVs1) are used and the potentialcurrent curve obtained is similar to steady-state curves (see chapter 2.5). Qualitative information extractable is provided by the E1/2-value of the analyte, and may allow the analyte's identification. The

Chapter 2: Experimental and Theoretical Details Linear Sweep and Cyclic Voltammetry

page: 42

current gives quantitatively valuable information since it is directly proportional to the concentration of the analyte. This technique provides accurate measurement of concentrations down to about 105M, but does not provide any useful information for elucidating reaction mechanisms. 2.4.5. Cyclic Voltammetry The applied potential waveform in cyclic voltammetry (Figure 2.5) is very similar to the waveform applied in linear sweep voltammetry. The linear sweep from Estart to Eswitch (switching potential) is employed initially and then followed by a return sweep from Eswitch to Eend. The basic idea is that each process should be swept at least twice, firstly on the initial (forward) scan and then a second time on the reverse scan. The current response obtained on the forward scan is mainly determined by the concentration and nature of the analyte present, whereas the response on the reverse scan is determined by the reactivity of the oxidized or reduced form of the analyte formed at the electrode surface. Hence, this technique provides information about both the nature of the analyte itself and its oxidized or reduced form. The cyclic technique also allows relatively simple elucidation of reaction mechanisms.
4
Eswitch

0.5

3 2

potential [V]

current [A]

0.0

1 0 -1

Eend Eswitch

Estart

-0.5
Estart Eend

-2 -3 -0.6 -0.4 -0.2 0.0 0.2 0.4 0.6

10

15

20

time [s]

potential [V]

Figure 2.5. Cyclic Voltammetry. Applied waveform (left) and resulting voltammogram (right). Estart = starting potential, Eswitch = switching potential, Eend = end or final potential. In this case Estart=Eend.

Cyclic voltammetry is a very flexible and multipurpose technique. From the appearance of the wave shape preliminary conclusions can be rapidly drawn concerning both the nature and course of the redox reaction being investigated. It is an excellent tool to explore electrochemical/chemically coupled reactions and to extract thermodynamic as well as kinetic data.
E
r 1 /2 ox

Ep E p /2 ip
ox ox

ip / 2 ip E p /2 Ep
- 0 .4 - 0 .2
re d re d re d

ox

0 .0

0 .2

0 .4

- 0 .4

- 0 .2

0 .0

0 .2

0 .4

p o te n tia l [V ]

p o te n tia l [V ]

Figure 2.6. Theoretical cyclic voltammograms. Left hand side shows a fully reversible one-electron transfer process. Right hand side shows an irreversible process.

Chapter 2: Experimental and Theoretical Details Linear Sweep and Cyclic Voltammetry

page: 43

In Figure 2.6. (left), a cyclic voltammogram of a reversible one-electron transfer process and the measured parameters are shown. At 25C the peak current, ip, for the forward peak is given by the Randles-Sevcik Equation [113] ip = 2.69105n3/2AD1/2c01/2 (2.18)

where n is the number of electrons transferred, A is the electrode area [cm2], D is the diffusion coefficient of the analyte [cm2s1], c0 is the concentration of the analyte in the bulk [mol cm3] and is the scan rate [Vs1]. If the diffusion coefficients for the bulk solution species and its oxidized/reduced form are identical (which practically often is the case), then ipox = ipred. For an ideal system under standard conditions E0 = E0' = Er1/2 and the reversible half-wave potential lies half way between the oxidation and reduction peaks (dashed vertical line in Figure 2.6) Er1/2 = Epox Epred (2.19)

At 25C, the oxidation and reduction peaks are separated by 57mV for a one-electron process Epox Epred =
57 mV n

(2.20)

The potentials at half the peak current height are only separated by 56mV
E ox E red = p p
2 2

56 mV n

(2.21)

The cyclic voltammogram displayed in Figure 2.6 (left) corresponds to a reversible one-electron oxidation process A A+ + e (2.22)

and the Er1/2 for this process is Er1/2=0V. The cycle is started at Estart=0.5V, a potential where according to the Nernst Equation (2.15) less than 106% of A will be oxidized to A+, a negligible amount. Thus, when the scan is commenced at Estart no current flow can be detected. The scan will proceed, until a potential value is reached which is sufficiently close to the Er1/2 of the process to allow a detectable amount of current to flow (onset of the forward wave). The current will continue to increase according to the Nernst equation until all of the substance in the vicinity of the electrode surface is used. This is the situation where diffusion is determining the current. As more material in the vicinity of the electrode is oxidized the current will start to decay and the onset of this process forms the forward peak of the cyclic voltammogram. At the switching potential, Eswitch, only A+ will be present at the electrode surface and its vicinity and similarly as A on the forward scan, the reduction of A+ will form the current-potential curve on the reversal.

Chapter 2: Experimental and Theoretical Details Linear Sweep and Cyclic Voltammetry

page: 44

(a) Chemical Irreversibility In Figure 2.6. (right) a cyclic voltammogram of an irreversible oxidation process is shown. This time, the charge transfer is followed by a chemical reaction in which A+ B occurs with a calculable rate constant. A A+ A+ + e B (2.23) (2.24)

The electrochemically formed A+ is removed from the equilibrium. Therefore on the reverse scan, the current response is lower than might otherwise be expected. Under these conditions, the determination of the reversible half-wave potential cannot be done in a straight forward manner as is the case when the process is fully reversible. Due to the kinetics of the follow up reaction the forward peak might be shifted by several hundred millivolts to more negative potentials, thus Er1/2 cannot be calculated from Equation 2.20 by assuming that Er1/2 = Epox 29mV. To be able to extract a value for the reversible half-wave potential from an irreversible cyclic voltammogram, all kinetic parameters must be known or reversible conditions need to be achieved. In many cases, this can be done by increasing the scan rate, thereby reducing the time frame (duration) of the experiment. If the time frame of a voltammetric experiment becomes fast enough, it can outrun the kinetics of the chemical reactions, so that B is not given enough time to form, and reversible conditions will be achieved which then allow an Er1/2-value to be measured. Data obtained from cyclic voltammograms as a function of scan rate also allow the kinetics of follow up reactions to be determined. For example, Nicholson and Shain [114] have developed working curves which allow estimation of kinetic rate constants for simple irreversible systems as described in Equations 2.23 and 2.24 by investigating the peak current ratio (ipox/ipred) as a function of scan rate. (b) Heterogeneous Irreversibility This term describes kinetic problems that arise when the initial electron transfer between a species, usually in solution, and the electrode surface (usually solid, thus heterogeneous irreversibility). A typical example is the high hydrogen overpotential at glassy carbon and mercury electrode material. This is an effect of the rate of electron transfer between electrode material and protons in solution. If the electron transfer is blocked, the potential for the proton reduction will become thermodynamically more difficult, thus an overpotential is created. Other forms of heterogeneous irreversibility can be due to passivation layers formed on the electrode surface. In corrosion chemistry this type of irreversibility and overpotentials are important factors, the Tafel plot is mentioned here as one of many graphicalmathematical tools available to characterize these effects.

Chapter 2: Experimental and Theoretical Details Steady State Techniques

page: 45

2.5. Steady State Techniques


2.5.1. Theory Cyclic voltammetry at a macrodisk electrode is a transient technique, in which the current is a function of time. In contrast, under steady-state conditions, the current is only a function of applied potential. At potentials sufficiently beyond the reversible half-wave potential, the current approaches a constant maximum value, which is called the limiting current (ilim or iL). The waveshape is governed by the Nernst equation and after substitution in Equation 2.15 the following relation is obtained
E = E0 i i 0.0592 log lim n i

(2.25)

where ilim is the limiting current and i is the measured current. Equation 2.15 describes the voltammetric waveform as a function of potential, the limiting current is again determined by the mass transport toward the electrode surface and is technique dependent. In Figure 2.7 the resulting sigmoidal waveform associated with the occurrence of a reversible one-electron oxidation process is shown. At the reversible half-wave potential, the concentration of both the analyte and its oxidized form are equal at the electrode surface and the current has the value of half of the limiting current, so that Er1/2 is the potential at
i lim 2

(2.26)

From Equation 2.25 it can be easily calculated that the potential difference between the potentials, where the current is one quarter and three quarters of the limiting current, is
E 14 E 3 4 = 56.5 mV n

(2.27)

which enables the number of electrons transferred in the redox process to be calculated.
1.0

ilim
0.5

slope: 59mV intercept: E1/2

E1/4

E1/2
zero current

1/2 ilim 1/4 ilim

log((ilim-i)/i)

E3/4

3/4 ilim

0.0

-0.5

-200

-100

100

200

-1.0

-40

-20

20

40

potential [mV]

potential [mV]

Figure 2.7. Steady state voltammogram. Right hand side shows "log-plot" which enables the reversible halfwave potential and the number of electrons transferred to be easily and accurately to be determined.

Chapter 2: Experimental and Theoretical Details Steady State Techniques

page: 46

A much better practical tool for measuring the reversible half-wave potential from steady-state voltammograms is the "log-plot" (Figure 2.7 right). When the logarithmic part of Equation 2.25 is plotted versus the potential, a linear plot is obtained for a reversible process. The inverse of the slope is
59.2 n

mV and enables the number of electrons transferred to be calculated. The position

where the linear plot intersects with the zero line (dashed in Figure 2.7 right) gives the value for the reversible potential. (a) "Log-plot" Practical Considerations: Using a "log-plot" to evaluate steady-state responses generally gives reliable results. Furthermore, since a linear plot usually is only obtained when the Nernst equation is obeyed, the linearity tells whether mass transport controlled reversible electron transfer conditions have been achieved. Data points are only calculated within ca. 100mV around the Er1/2-value, as small errors at other potentials introduce large errors in the "log-plot".
1.5

2
1.0

log((ilim-i)/i))
500 1000
+

current [nA]

0.5

ilim

-1

0.0 0

-2 150 200 250 300 350 400


+

450

potential [mV] vs. Fc/Fc

potential [mV] vs. Fc/Fc

1.5

slope: 69mV E1/2 = 353mV log((ilim-i)/i))


ilim
1

current [nA]

1.0

0.5

0.0 0 500 1000


+

-1 150

200

250

300

350

400
+

450

potential [mV] vs. Fc/Fc

potential [mV] vs. Fc/Fc


3 2 1

slope: 59mV E1/2 = 0.0mV

ilim

log((ilim-i)/i))

current

0 -1 -2 -3 -150

-300

-200

-100

100

200

300

-100

-50

50

100

150

potential [mV]

potential [mV]

4 3 2

slope: 67mV E1/2 = 3.3mV

log((ilim-i)/i))

1 0 -1 -2 -3 -150 -100 -50 0 50 100 150

current

ilim

-300

-200

-100

100

200

300

potential [mV]

potential [mV]

Figure 2.8. Distorted steady state voltammograms and "log-plots". a) + b) 1.0mM [(tctpy)Ru(NCS)3]+ in CH3CN (0.1M Bu4NPF6), 11m diameter microdisk platinum electrode, b) baseline corrected c) + d) simulated voltammograms for one-electron process, d) distorted response caused by presence of a sloping baseline.

Chapter 2: Experimental and Theoretical Details Steady State Techniques

page: 47

Figure 2.8 illustrates the use of "log-plots" for commonly encountered types of slightly distorted steady-state voltammograms. The first two examples (Figure 2.8a and b) are taken from work conducted on the Black Dye (see Chapter 6 for details) and show the effect of adsorption on steady-state voltammograms. The oxidation of [(tctpy)Ru(NCS)3]+ in acetonitrile is an almost reversible one-electron process, which is accompanied by adsorption. The steady-state voltammogram for the oxidation of [(tctpy)Ru(NCS)3]+ shown in Figure 2.8a is clearly different to those expected for a sigmoidal waveform for a reversible one-electron process (compare Figure 2.7 and Figure 2.8c). Due to the influence of adsorption, which in this case leads to passivation of the electrode surface, the current decays after having reached a maximum value. A "log-plot" constructed from this voltammogram taking the indicated limiting current value results in a curved plot rather than a straight line as expected for a reversible process. This could lead to incorrect conclusions regarding the degree of reversibility of the process. If a slightly higher value for the limiting current is assumed (Figure 2.8b), extrapolating the sigmoidal wave and additionally performing a baseline correction, results in the "log-plot" shown in Figure 2.8b right. A straight line is obtained for much of the wave and only data points at positive potential values, which are effected by adsorption, deviate from the linear plot. Performing a linear regression on the linear data points results in an inverse slope of 69mV, which is moderately close to the theoretically expected value of 59mV. This "log-plot" also allows the Er1/2-value and the diffusion coefficient to be calculated by using the adjusted limiting current and Equation 2.28 Comparison of the D value with values obtained from other steady-state techniques can confirm the correctness of the limiting current value used. Figure 2.8c shows a simulated steady-state voltammogram for a one-electron oxidation process with Er1/2=0V. As expected, the resulting "log-plot" is linear with an inverse slope of 59mV and Er1/2=0V. In Figure 2.8d the same voltammogram is shown, but with a slightly sloping baseline, as frequently encountered in experimental studies. The "log-plot" for this steady-state voltammogram is clearly not linear, although it is a fully reversible process. The inverse slope, estimated from data points around 0V, is 69mV, almost 20% more than without the sloping baseline. Furthermore, the calculated reversible half-wave potential occurs at a more positive potential and deviates from the true potential. Thus, it is essential that a proper baseline correction is performed before analysis of steady state voltammograms. 2.5.2. Microdisk Electrode To achieve steady-state conditions mass transport needs to be independent of time. In the case of microdisk electrodes this is achieved by radial or edge diffusion to the electrode surface [115, 116]. When the concentration level of a species is decreased at a disk, a concentration gradient will form. Diffusion will then follow via the shortest possible route to the electrode surface. Two situations are possible. Firstly, a particle present in the area described by a cylindrical extension of the electrode surface, will diffuse perpendicularly to the electrode surface. This is called planar

Chapter 2: Experimental and Theoretical Details Steady State Techniques

page: 48

diffusion and is indicated blue in Figure 2.9. Linear diffusion dominates under transient cyclic voltammetric conditions. Secondly, electroactive species present in areas other than those described by the cylindrical extension of the electrode surface, will diffuse to the edge of the electrode surface, which is its closest point to the electrode surface. This is called radial or edge diffusion.

Figure 2.9. Edge or radial diffusion (red) dominates mass transport at a microdisk electrode (left) and planar diffusion (blue) dominates mass transport to a macrodisk electrode (right). = diffusion layer thickness.

Figure 2.9. illustrates both types of diffusion. By choosing the right experimental conditions, one type of diffusion becomes negligible in relation to the other. Flux and current densities at the edge of an electrode are much larger in magnitude than in the center of the electrode. The theories for both types of diffusion differ [117] in that planar diffusion is time dependent whereas radial diffusion can be treated as time independent, as required for achieving steady-state conditions. The proportions of planar and radial diffusion in an experiment is determined by electrode surface size and the diffusion layer thickness, , which is frequently called the Nernst layer. determines how far the solution is perturbed by the experiment and is a function of experimental conditions, such as the scan rate. For instance, as the scan rate is increased the diffusion layer is decreased. Hence, even at microdisk electrodes, planar diffusion can dominate, if the scan rate is large enough to decrease sufficiently. Microdisk electrodes usually have electrode diameters of 0.1100m, so that scan rates below about 20mVs1 need to be used to achieve steady-state conditions. The small electrode size requires rigorous polishing to prevent deactivation from adsorption or other forms of passivation. The steady-state limiting current is given by the following relationship [117]: ilim = 4nFDc0r (2.28)

where r is the electrode radius [cm] and the other symbols are as usual. Currents measured with this technique require very sensitive equipment, since currents in the pA/nA range are typically measured. On the other hand small currents reduce distortions from Ohmic resistance, which can be a considerable advantage. 2.5.3. Rotating Disk Electrode This technique utilizes an electrode, which rotates around its central vertical axis as shown in Figure 2.10. This rotation causes convection or forced mass transport to the electrode surface [118].

Chapter 2: Experimental and Theoretical Details Steady State Techniques

page: 49

Figure 2.10. Mass transport at a rotating disk electrode.

If the disk is rotated fast enough and a sufficiently small scan rate is chosen, the diffusional component becomes negligible and hydrodynamic mass transport dominates, resulting in steadystate behavior. Hydrodynamic mass transport to a variety of different electrode geometries has been studied and mathematical solutions have been provided by Levich [118]. Hence, the equation describing the limiting current in rotating disk voltammograms is called the Levich equation
i L = 0.620nFAD 3 c 0 6
1 1 1 2

(2.29)

where A is the electrode area [cm2], is the viscosity of the solvent [cm2s1] and is the angular velocity of the electrode [s1] ( = 2f, f is the frequency or rotation rate). It should be noted, since this type of mass transport is hydrodynamic where the fluid (solution) is in motion, the viscosity of the solvent, , needs to be incorporated into the equation. Typically rotation rates of 500 to 5000rpm (rpm = revolutions per minute) are employed. Due to the hydrodynamic mass transport to the electrode and the relatively large electrode areas used compared to microdisk electrodes, the resulting currents are sufficiently large to be detected without significant instrumental amplification. However, distortions due to the ohmic drop may result in incorrectly measured Er1/2-values. To obtain a pure steady-state current, the electrode must be operated under the appropriate hydrodynamic mass transport conditions. Verification of this behavior can be obtained by investigating the limiting current as a function of rotation rate and plotting the limiting current versus the square root of rotation rate in a Levich plot. According to the Levich equation (2.29) the plot must result in a linear dependence, if the appropriate hydrodynamic mass transport conditions have been achieved. It has been suggested [96], that it might be desirable to conduct voltammograms at a rotating disk electrode under mixed diffusionalhydrodynamic conditions. Diffusional mass transport will dominate around the potential of the redox process and this will result in formation of peaks, similar as found under cyclic voltammetric conditions, which eases the determination of Er1/2-values. In contrast, at potentials sufficiently past the redox process, the current will approach the hydrodynamic limiting current, which eases the determination of the diffusion coefficient.

Chapter 2: Experimental and Theoretical Details Steady State Techniques

page: 50

2.5.4. Channel Flow Electrode In this electrode geometry hydrodynamic mass transport conditions are achieved by a flow of the solution over the electrode surface.

wall flow diffusion electrode


Figure 2.11. Channel flow electrode. Laminar flow dominates parallel to the electrode surface, diffusion takes place perpendicular to the electrode surface.

height

The rectangular electrode is situated in a thin-layer cell arrangement as shown in Figure 2.11. The cell height and the flow rate are chosen to obtain laminar flow conditions in the channel. The flow profile that develops in the channel is parallel to the electrode surface and is elliptical perpendicular to the electrode surface, that is the flow is largest in the center of the cell and approaches zero near the electrode surface. The limiting current is given by the Levich equation for the channel flow electrode [118]
 x V i L = 0.925nFD c 0 w f2 e h d
2 3 1 3

(2.30)

w is the electrode width [cm], xe is the electrode length (in direction of the flow) [cm], h is half of

 is the volume flow rate [cm3s1] the channel height [cm], d is the width of the channel [cm], V f
and the other symbols have their usual meaning.

2.6. Potential Step and Pulse Techniques


A sudden step in potential applied to an electrochemical system results in an instantaneous alteration of the solution at the electrode surface. Analysis of the current response of the system after this perturbation permits deductions to be made concerning electrode reactions and their rates. 2.6.1. Double Step Chronocoulometry In this experiment, the potential is stepped from a rest potential, Erest, where usually no Faradaic reaction takes place, to a potential significantly past the E1/2-value of the investigated redox reaction, Eforward. After a preset time, the potential is stepped back, usually to the potential of Erest. For a simple one-electron redox system like Red Ox + e, after a potential step past the corresponding Er1/2 the current response due to Faradaic electron transfer processes is expressed by the Cottrell equation, which is shown in red in Figure 2.12.

Chapter 2: Experimental and Theoretical Details Potential Step and Pulse Techniques

page: 51

I f (t ) =

nFAD 2 c 0 2t
1 1 2

(2.31)

However, when the potential is changed, the double layer has to be charged, giving rise to a capacitive current, Ic. The capacitive current decays very rapidly to zero and therefore is only significant when very fast reactions are investigated. Furthermore, it decays via et rather than t and hence can be distinguished from the Faradaic current component.
E RC e Ic ( t ) = R
t
1 2

(2.32)

In this equation R is the solution resistance [] and C is the double layer capacitance [F]. According to the Cottrell equation, a plot of Faradaic current versus square root of time results in a linear dependence. Such a plot is shown in Figure 2.12 right. Contributions from the charging current can be avoided by subtracting the results obtained from a background experiment.
If, faradaic current Ic, capacitive current Itotal, total current
If, faradaic charge Ic, capacitive charge Itotal, total charge

current

Eforward

Erest time

charge

potential

b)
0 0

0 0

time

time

1/2

Figure 2.12. Single step chronocoulometric experiment. a) Current response after step due to Faradaic component and double layer charging (insert: applied potential). b) plot of charge versus square root of time.

The Cottrell plot obtained from a potential step experiment is a very accurate tool for determining the diffusion coefficient for a species. Alternatively, the electroactive area of a planar electrode can be measured by taking a species of known concentration if n and the value for the diffusion coefficient are available. The slope of the plot is linearly proportional to the area and proportional to the square root of the diffusion coefficient. Chronocoulometry is a good tool for determination of adsorbed reactants. Since the material is adsorbed at the electrode surface, its reaction contributes to the current only immediately after the potential step and is proportional to the amount adsorbed so that the total charge at time t is given by:
Dt 2 + nFA0 + Q c
1

Q = 2nFAc 0

(2.33)

where 0 is the surface concentration of the adsorbed species. The amount adsorbed can easily be determined graphically by a plot of charge versus square root of time via an Anson plot [119]. Adsorption will lead to an offset of the linear plot, similar to the black line in Figure 2.12. (right).

Chapter 2: Experimental and Theoretical Details Potential Step and Pulse Techniques

page: 52

Since the total offset (FA0 + Qc) includes the capacitive charge, it is essential to perform experiments on the background. In double step chronocoulometric experiments, the potential is stepped back to the initial value after a certain time. The return current response contains information about the oxidized/reduced form of the investigated species. For an electrochemically and chemically reversible process it is a requirement that a straight line in a charge versus time plot is obtained for both the forward and the return step [120]. 2.6.2. Differential Pulse Voltammetry In differential pulse voltammetry the potential is swept in one direction similarly to linear scan voltammetry, but during the linear sweep, pulses are applied on top of the waveform as shown in Figure 2.13.
tp td measurement i2 i1

Emax E

1/2

current (i2-i1)

potential

W 1/2 scan direction

E pulse height

time

potential

Figure 2.13. Differential pulse voltammogram. a) Applied waveform and b) voltammogram.

The potential is stepped every td seconds. At the end of this step, the current is measured (i1) and then a short pulse (E) is applied on top of the base potential, typically in the range of 10100mV. This pulse (tp) is about 20100 times shorter than the base step. At the end of this step, a second current measurement (i2) is taken and the final current output is the differential current (i2-i1). Since tdrop is relatively long (scan rates used are within 425mVs1), the solution environment at the electrode surface is given by the applied potential at the base step. The deviation of the applied pulse is chosen to be as short as possible, ideally just short enough so that when the second current measurement is made at the end of this pulse, the majority of the charging current has decayed away to a negligible value. For an infinitely small pulse height, the half peak width is given by
W12 = 3.52 RT nF

(2.34)

The location of the peak potential, Emax, depends on the scan direction and on the pulse amplitude used
E r12 = E max + E 2

(2.35)

Chapter 2: Experimental and Theoretical Details Potential Step and Pulse Techniques

page: 53

If the process is irreversible, the peak height relative to the reversible one will diminish, the peak will broaden and Emax will shift. Due to the measurement of differential currents and effective elimination of contribution from the charging current, this technique is a powerful analytical technique in trace element analysis. Solution phase detection limits are about 108M, although by utilizing reactant adsorption the detection limit can be lowered by several orders of magnitude [106]. The present record in trace element analysis for the limit of detection is 51014M [121] for the detection of mercury by differential pulse adsorptive stripping voltammetry. However, the requirement for a good signal response in a solution phase reaction is a highly reversible redox process. 2.6.3. Controlled Potential Bulk Electrolysis Voltammetric techniques are almost non-destructive because the amount of substance reduced or oxidized during an experiment (potential scan or step) is negligible compared to that present in bulk solution, so that the solution phase composition practically remains unaltered during an experiment. In contrast, in a bulk electrolysis experiment the aim is to have all of the material participating in the electrochemical reaction. Quantitative electrolysis is achieved by applying the desired potential until the charge passing per unit time becomes negligible. Faraday's law correlates the amount of charge passed with the amount of electrons transferred
Q = nF m0 Mw

(2.36)

where m0 is the amount [g] of substance participating in the reaction, Mw is molecular weight [gmol1] of the substance and n the number of electrons transferred. The product can then be characterized in situ by spectroscopic and voltammetric techniques or isolated and identified. The time domain of this experiment depends on the experimental setup, but usually requires about 20 minutes in a conventional cell. Conclusions can be drawn from a bulk electrolysis experiment regarding the long-term chemical reversibility and stability of a redox system and whether synthetic attempts will be successful to isolate the oxidized or reduced form of the compound. The ultimate proof for reversibility is an exhaustive oxidative (reductive) bulk electrolysis followed by a reverse exhaustive reductive (oxidative) bulk electrolysis, which should result in no noticeable change in the chemical composition of the solution. Furthermore, the number of electrons transferred in each single electrolysis experiment must be the same. Commonly bulk electrolysis is employed to determine the electrons transferred in a redox process and to electrogenerate substances.

Chapter 2: Experimental and Theoretical Details Spectroscopic Techniques

page: 54

2.7. Spectroscopic Techniques


2.7.1. UV/VIS Spectroscopy and Optically Transparent Thin Layer Electrolysis (OTTLE) Electronic spectra of molecules can give information about their electronic structures. The basic effect studied in electronic spectroscopy is the interaction (absorption) of light (photons) with a substance. As light passes through a substance, it is attenuated according to Lambert's law of extinction: I(x) = I0 exc0 (2.37)

where is called the molar absorption coefficient [M1cm1], c0 is the concentration of the substance [M], x is the thickness of the layer or cell [cm], and I0 is the light intensity at x=0. In a molecule, electrons reside in specific energy levels, which are characterized by their quantum numbers. The interaction of a photon and a molecule or atom is subject to the laws of conservation of energy, specifically in this case of linear and angular momentum. Thus, excitation of an electron within the energy levels of a molecule requires the supply of energy from an incoming photon. UV/VIS spectroscopy allows direct probing of the molecular orbital arrangements. By using reference compounds and molecular orbital calculations, bands can be often assigned to transitions arising between specific orbitals of a complex, which in turn may allow quantification of the energy levels. A more detailed discussion on molecular orbitals can be found in chapter 1.8.2. Optically transparent thin layer electrolysis is a combination of electronic spectroscopy and controlled potential electrolysis. The electrolysis is conducted in a thin (1mm) quartz cuvette, as commonly employed in UV/VIS spectroscopy, with a gauze type (mostly light transparent) working electrode, which is placed in the optical path. The reference and counter electrode are separated from the solution by a frit and are placed in the top compartment. The schematic setup of an OTTLE cell is illustrated in Figure 2.14.
re ce

we

lig

ht

Figure 2.14. OTTLE cell. x = path length, re = reference electrode, we = working electrode, ce = counter electrode.

Chapter 2: Experimental and Theoretical Details Spectroscopic Techniques

page: 55

The thin-layer arrangement results in a comparable short electrolysis time of only a few minutes. An OTTLE experiment offers the convenient opportunity to spectroscopically characterize one or more oxidation states of a compound. Furthermore, conclusions regarding the reversibility of the investigated redox couples can be drawn. If a simple reaction, such as R R+ + e, is investigated, where only two species are in an equilibrium, electronic spectra recorded during the course of the reaction must show the presence of isosbestic points. Isosbestic points occur where both R and R+ have the same molar absorbance value. That is, it is an energy value where the spectra of the pure species, as well as mixtures, intersect, as long as the total concentration remains unchanged. The absence of isosbestic points indicates the occurrence of a side reaction, where three or more species are present. However, the presence of isosbestic points is not a proof that only two species are present in equilibrium. For example, the series of [Ru(bpy)3]2+/+/0/ [122] as well as [Ir(bpy)3]3+/2+/+/0 [123] exhibit an isosbestic point at 3104cm1, although each system consists of four species [124]. 2.7.2. Electron Spin Resonance Spectroscopy The technique of electron spin resonance (ESR) spectroscopy can only be applied to species having one or more unpaired electrons. Free radicals, diradicals or compounds in triplet or other states satisfy this rule and can therefore be investigated. Due to its high sensitivity, the ESR technique is particularly valuable to study highly reactive radicals which may be prepared in situ by electrochemical oxidation or reduction. Furthermore, ESR spectroscopy is frequently used to study transition metal compounds and complexes with unpaired free electrons, such as high spin complexes. From the paramagnetic behavior of these complexes its alternative name is derived, electron paramagnetic resonance spectroscopy. Each electron in a system is characterized by a series of quantum numbers, one of them is the electron spin quantum number, which can have a value of + or . It also characterizes the orientation of the electron in an applied external magnetic field. In the presence of a magnetic field (B0), a molecule having one unpaired electron has two electron spin levels in regard to the magnetic field and the energy levels are given by E = g BB0MS (2.38)

Here B is the Bohr magneton (9.274081024JT1), MS is the electron spin quantum number (+ or ) and g is the proportionality factor, equal to 2.00232 for the free electron, and is the value of interest. The simple picture of two energy levels must be modified if the electron can interact with a neighboring nuclear spin. The energy levels then become E = g BB0MS + aMSmI (2.39)

where mI is the nuclear spin quantum number for the neighboring nucleus and the constant a is called the hyperfine coupling constant. Thus, a single nucleus with spin (e.g. 1H) will split the

Chapter 2: Experimental and Theoretical Details Spectroscopic Techniques

page: 56

present set of energy levels into two and information regarding the electronic neighborhood of the electron might be retrieved. In the above discussion, ESR spectra obtained for isotropic systems have been considered, as can be obtained for solutions or particles with spherical symmetry. In practice, solids or frozen solutions are studied. Substances, which have axial or lower symmetry, will have a preferred orientation to the magnetic field and the momentum of the electron needs to be treated spatially. A molecule with axial symmetry can have two non-equal orientations in regard to the magnetic field, either parallel or perpendicular. Hence, two g-values (g and g ) can be measured for either orientation. Similarly, a molecule with lower symmetry will exhibit three g-values, one for each of the three principal axes (gx, gy, gz) and yield a rhombic signal. 2.7.3. Electrospray Mass Spectrometry In a mass spectrometric (MS) experiment, ions of a substance are produced and then separated by their masstocharge ratios (m/z). Generally, during the highly energetic ionization process, fragmentation of the molecules occurs which in turn allows structural information to be retrieved from examination of mass spectra. However, the electrospray (ES) ionization technique is a very gentle ionization technique, which on its own does not produce charged species from uncharged species or even alter the charge of any species. It requires the presence of charged species (ions) in solution. The electrospray process removes the solvent (by evaporation) and extracts the charged particles by applying a potential (cone voltage) between the sample inlet and the mass spectrometer inlet. The ions are successively accelerated by a series of electronic lenses and drawn from the outer (ambient pressure) into the inner (high vacuum) of the mass spectrometer. Ideally, the only signal found is the molecular peak. Although ES-MS needs the presence of charged species, some neutral compounds (M) may be investigated by adding a proton source to the solvent, thereby generating protonated forms of the neutral compound (M-H+, M-H22+, etc.). As mentioned above, the electrospray technique is very gentle and generally does not result in fragmentation. Conversely, since it is such a gentle technique leaving even very weakly bound complexes intact, very often the formation of adduct complexes (e.g. M2-H+, M-Solvent-H+, etc.) can be seen. These are formed during the electrospray process rather than being present in solution. By increasing the cone voltage, the higher energy of the extraction process can break up such complexes and as a consequence their formation can be avoided. However, very high cone voltages can result in fragmentation. 2.7.4. Nuclear Magnetic Resonance Spectroscopy Nuclear magnetic resonance (NMR) spectroscopy is the most powerful spectroscopic technique available for elucidating structural information [125, 126]. In ESR spectroscopy, unpaired electron spins provide the basis for the measurement. In NMR spectroscopy, unpaired nuclear spins are investigated. For every isotope there is a ground state nuclear spin quantum number, I, which has a

Chapter 2: Experimental and Theoretical Details Spectroscopic Techniques

page: 57

value of zero or spin of


1 2

n 2

, where n is an integer. Most commonly investigated are isotopes with a nuclear

, such as 1H and 13C, because isotopes with a higher spin generally give more complicated

spectra. Isotopes with no nuclear spin, I=0, cannot be investigated. In a magnetic field, two energetic levels exist for a nuclear spin, parallel and anti-parallel to the applied field direction. Resonance phenomena occurs when the sample is irradiated with a frequency corresponding to the energy difference between those two levels. Thus, NMR spectroscopy measures the resonance frequency of nuclei. Its value strongly depends on the chemical environment of the nucleus. Spinspin coupling to adjacent nuclei will split the set of energy levels further, so that multiplet signals are observed. Summarizing, the resonance frequency (shift) and multiplicity of the signal allows elucidation of the chemical environment of the investigated nucleus. The total number of signals measured reflects the number of nuclei with different chemical environments and therefore allows conclusions about the symmetry of the compound to be drawn. In this work only 1H and 13C-NMR spectroscopy were employed to confirm the structure and the purity of compounds.

2.8. Electrochemical Quartz Crystal Microbalance


When a thin quartz disk is subjected to pressure, charges are dislocated in the material and a small electric signal may be monitored if electrodes are placed onto its surface. That is, such a crystal can convert mechanical displacement into electricity. This phenomenon was discovered by Jacques and Pierre Curie in 1880 [127] and termed piezoelectricity, which means "pressure electricity". Conversely, if an electric field is applied on the crystal it will induce a mechanical displacement. This is called the converse piezoelectric effect and one of the most fascinating applications of this phenomenon can be found in Atomic Force Microscopy where the position of an atomic tip is controlled by a converse piezoelectric crystal to a precision better than 1 (0.1nm). The resonance frequency of the quartz crystal is determined by its physical properties such as density, thickness and shear modulus. However, if any of these physical parameters change, the resonance frequency of the crystal will also change. For example, if a mass is attached to the surface of the crystal disk, e.g. by adsorption or deposition, the resonance frequency will decrease. This relationship was studied by Sauerbrey [128] who proposed an equation which relates the resonance frequency to the mass of the crystal.
F = f 02 m N q

(2.40)

F is the frequency change, m is the mass change per unit area, f0 is the fundamental resonance frequency of the crystal, N is the frequency constant of the crystal and q is the density of the

Chapter 2: Experimental and Theoretical Details Spectroscopic Techniques

page: 58

crystal. This relationship enables mass changes to be measured that occur on the surface of the crystal. A device constructed from a crystal and a frequency meter is called Quartz Crystal Microbalance (QCM) and is extremely suitable for measuring very small mass changes. One of the first applications of the QCM was to measure the thickness of materials deposited on surfaces by vacuum evaporation techniques. In the early 1980's, Konash and Bastiaans [129] and Nomura [130] demonstrated that the QCM can be applied in liquids, provided that the two sides of the crystal were electrically isolated from each other. This is most easily accomplished by allowing only one side of the crystal to be in contact with the solution. The validity of the Sauerbrey relationship in liquids requires that mass changes must not cause a frequency shift greater than approximately 2% of the fundamental resonance frequency of the crystal. For higher degrees of mass loading, the relationship is not linear and alternative models have to be used. If the crystal surface is coated with a conducting material, e.g. gold, and used as the working electrode, the Electrochemical Quartz Crystal Microbalance (EQCM) arrangement is formed. This allows mass changes in the nanogram range to be recorded as a function of an electrochemical experiment.

Chapter 3

Electrochemical and spectroscopic studies on the oxidation and reduction of the cis-(Et2-dcbpy)2RuX2 series of photovoltaic sensitizer precursor complexes (Et2-dcbpy =2,2'-bipyridine-4,4'-diethoxydicarboxylic acid, X = Cl, I, NCS, CN).

part of this work has been published in

'Electrochemical and spectroscopic studies on the oxidation of the cis-(Et2dcbpy)2RuX2 series of photovoltaic sensitizer precursor complexes (Et2-dcbpy =2,2'-bipyridine-4,4'-diethoxydicarboxylic acid, X = Cl, I, NCS, CN).' G. Wolfbauer, A.M. Bond, D.R. MacFarlane, Inorganic Chemistry, 1999, in press. 'Electrochemical and spectroscopic studies on the reduction of the cis-(Et2dcbpy)2RuX2 series of photovoltaic sensitizer precursor complexes (Et2-dcbpy =2,2'-bipyridine-4,4'-diethoxydicarboxylic acid, X = Cl, I, NCS, CN).' G. Wolfbauer, A.M. Bond, D.R. MacFarlane, J. Chem. Soc., Dalton Trans., 1999, submitted.

Chapter 3: Studies on Model Esters Compounds Introduction

page: 60

3.1. Introduction
Ruthenium(II) polypyridyl complexes have been known of for many years. They are characterized by strong metal to ligand charge transfer (MLCT) bands in the visible region which make these compounds highly colored. Several other advantageous photophysical properties, such as long lived triplet states and other luminescence phenomena [62, 131-133] explain why ruthenium polypyridyl complexes have been widely used as photosensitizers [25, 31, 134-136]. An important application of these photosensitizers is the conversion of sun light to electrical or other forms of energy. Currently, most photovoltaic sensitizers are based on ruthenium polypyridyl complexes. For example [Ru(bpy)3]2+ (bpy = 2,2'-bipyridine) is a well known sensitizer in so called water splitting systems [56, 58, 62]. Light induced excitation of one metal based electron into one of the ligands is followed by a chemical charge separation and finally reduction of the solvent, water, which will produce the valuable energy source hydrogen [137]. More recently, complexes of the formula (H2-dcbpy)2RuX2 have become very popular sensitizers for photoelectrochemical cells (PEC) [29, 67, 70, 74, 134]. Introducing carbon acid units onto the bipyridine groups enables more effective attachment to semiconductor surfaces. The replacement of one bipyridine by two halides results in a preferable red shift of the absorbance bands. The combination of a semiconductor with a ruthenium sensitizer results in a very efficient charge separation and the possibility of inexpensive PECs with energy conversion efficiencies of up to 10% [29, 36, 72, 138, 139]. As would have been expected, photophysical studies [68, 86, 93, 94] on these kinds of sensitizers have been extensive. However, perhaps surprisingly, general studies characterizing the nature of the oxidized states of those complexes, which is intrinsically important in understanding the physical properties of photovoltaic cells, are yet to appear. Such information is of particular importance in understanding the stability of the oxidized state in photosensitizer applications, high stability being a requirement for a long lived device, despite the relatively short period of time (<1s [50]) over which the oxidized state exists in the device before it is reduced to its original state by a charge mediator (commonly iodide).

COOC2H5 H5C2OOC N X X = Cl INCS CN COOC2H5 N Ru N X COOC2H5

Figure 3.1. Structure of cis-L2RuX2.

Chapter 3: Studies on Model Esters Compounds Introduction

page: 61

The most common photovoltaic sensitizer currently used in PECs has the structure cis-(H2dcbpy)2RuX2, where X = CN or NCS and H2-dcbpy = 2,2'-bipyridine-4,4'-dicarboxylic acid. The dye attaches onto the TiO2 surface via formation of ester linkages [67-69], by two acid groups on one bipyridine ligand with the other two acid groups remaining unchanged. On exposure to an incident light photon, a metal based electron from the ruthenium center is promoted into these esterified ligands, resulting in generation of a ((-Ti-OOC)2-bpy)Ru radical, which then initiates the energy producing step by injecting the electron into the conduction band of the TiO2. Thus, studying the esterified analogues of the cis-(H2-dcbpy)RuX2 sensitizers is particularly relevant to the photosensitizer applications of (H2-dcbpy)2RuX2 compounds. Literature reports on the electrochemical behavior of transition metal complexes with the esterified ligand, [M(Et 2dcbpy)3]2+ (M = Ru, Os, Fe) [133, 140-147], show that these complexes exhibit feature rich and highly reversible electrochemistry, whereas few reports on the electrochemical behavior of complexes with protonated acids may be found [70, 148], with no reports of reduction processes. Thus, the cis-(Et2-dcbpy)2RuX2 complexes, where the Et2-dcbpy ligand is the ethyl esterified analogue of the H2-dcbpy ligand, appear to be ideal model compounds for the protonated acid sensitizers used in photovoltaic cells. Furthermore, since the electronic influence of the esterified carboxylic acid groups (-COOC2H5) on the bpy ligand should be small compared to the protonated acid groups (-COOH), it can be proposed that the ester analogues can be conveniently used as superior model compounds to bpy compounds for understanding the redox and spectroscopic properties of the considerably more surface active and reactive protonated acid sensitizer compounds (see chapters 4, 5 and 6). Commonly, [Ru(bpy)3]2+ is taken as the model compound [62] used for comparison of electrochemical and spectroscopic data of newly synthesized ruthenium polypyridyl complexes. However, the electron withdrawing carboxylate group present in the H2-dcbpy and Et2-dcbpy ligands generate distinctively different physical and chemical properties to those of the parent bpy ligand. For example, [Ru(bpy)3]2+ [122, 145, 149-153] exhibits up to three reversible reductions and one reversible oxidation when voltammetric measurements are made under standard laboratory conditions. However, under ultra-inert conditions and at low temperature, six reversible reduction waves can be observed [145]. All reduction processes in ruthenium bipyridine complexes are considered [62, 154] to be bipyridine based. Replacing bpy by Et2-dcbpy to give [Ru(Et2-dcbpy)3]2+, offers the possibility of adding up to ten electrons under voltammetric conditions to the complex and ten reversible reduction waves have been observed [145] under ultra-inert and low temperature conditions. However, Morris et al. [155] reported that rigorous dry box conditions are necessary to observe full chemical reversibility for this class of compounds. Theoretical molecular orbital calculations [144] have shown that the rich cathodic electrochemistry observed with [Ru(Et2-dcbpy)3]2+ arises because of the presence of very low lying 1* and 2* orbitals. The availability of energetically low 2* orbitals has been attributed to the -conjugation effect of the carboxylate groups. A common and interesting property associated with the reduction

Chapter 3: Studies on Model Esters Compounds Introduction

page: 62

of these polypyridine type complexes is the specific localization of added electrons on one ligand [124], which leads to ligand based mixed valency states and induces "electron hopping" [151, 156] between ligands.

3.2. Experimental
Electrochemical measurements in organic solvents were carried out using a standard threeelectrode arrangement with a platinum wire as the counter electrode and a Ag/Ag+ (CH3CN, 10mM AgNO3) double junction reference electrode. The working electrodes were in-house constructed platinum and glassy carbon macro and microdisk electrodes of stated diameters. Before each experiment, the electrodes were polished with an aqueous aluminium oxide (0.3m) slurry and rinsed with acetone. A Metrohm 628-10 assembly was used for rotating-disk electrode measurements. CYSY-1R (Cypress Systems) or BAS100 (Bioanalytical Systems) electrochemical systems were used to obtain the voltammograms. For very fast scan rate cyclic voltammetric experiments, a two electrode cell arrangement (working and reference electrode) was used with a PAR 175 waveform generator and an AMEL 551 (Milano, Italy) potentiostat. The fast scan data were captured with a GOULD 4035 (Essex, England) digital oscilloscope. The DigiSim 2.1. software package (Bioanalytical Systems) was used for digital simulation of voltammetric results. The potential of the reference electrode was calibrated against that of the ferrocene/ferricenium (Fc/Fc+) redox couple via measurement of the ferrocene oxidation process under identical conditions (technique, temperature, solvent and electrolyte) being used for the compounds of interest. For measurements in aqueous solution an aqueous Ag/AgCl (3M KCl) reference electrode was used. In the latter case the potential is quoted directly versus the potential of the Ag/AgCl reference electrode. Oxygen was removed from solutions by purging with high purity nitrogen or argon. The electrolyte used for all voltammetric experiments was tetrabutylammonium hexafluorophosphate (Bu4NPF6), prepared according to a literature method [112], and dried under vacuum at 90C for 10 hours. For solid state voltammetric measurements, the compound was mechanically attached to a 5mm diameter edge plane pyrolytic graphite disk electrode by gently rubbing the clean electrode surface against a filter paper onto which powdered compound had been placed. The electrode, containing mechanically attached microcrystals of the solid, was then placed into the solvent (with electrolyte) of interest. The bulk electrolysis platinum electrode cell was in-house built and had the appearance of a beaker (25mL total volume, 5cm diameter) into which a glass cylinder with frit at bottom was placed (12mL total volume). The working and counter electrodes were both cylindrical large surface area platinum gauze basket designs. The working electrode was placed inside the glass cylinder containing the frit, along with the solvent, electrolyte and analyte of interest. The Ag/Ag+ reference electrode was also placed inside the inner compartment and the counter electrode was

Chapter 3: Studies on Model Esters Compounds Experimental

page: 63

placed in the outer compartment which was filled only with the solventelectrolyte mixture. The lid to the cell contained five inlets/outlets to allow additional electrodes to be placed into the inner solution so that monitoring of the course of bulk electrolysis experiments was feasible and to permit electrical contact with the working, reference and counter electrodes to be made. Bulk electrolysis at platinum was usually conducted with sample volumes of 511mL (inner compartment). During the electrolysis, the solutions in both compartments were vigorously stirred by a stream of nitrogen. Exhaustive electrolysis was said to have occurred when the charge passed per minute was less than 1% of the charge passed in the first minute. This definition typically resulted in electrolysis times of 1520 minutes for a one-electron process. X-band electron spin resonance (ESR) spectra were recorded with Varian E-12 or Bruker ESP380 spectrometers. All spectra were measured at 77K. Gains in the range 1.25102 to 2.5103 and modulation amplitudes around 0.01mT and 0.4mT were typically used. The g values were determined through comparison with the external standard solid diphenylpicrylhydrazyl (DPPH, g=2.0037). Samples were generated ex situ by controlled potential electrolysis experiments at a platinum gauze electrode under an argon atmosphere and transferred, with the exclusion of air, into a nitrogen degassed quartz ESR tube. Temperature control for ESR experiments was achieved using an ITC4 thermostat (Oxford Instruments Ltd., UK). Optically transparent thin-layer electrolysis (OTTLE) experiments were carried out at a platinum gauze electrode, a 1mm quartz cuvette, a Varian Cary 5 NIR/VIS/UV spectrometer and an EG&G PAR 273 potentiostat. The detailed experimental setup is described in reference [157] and chapter 2.7.1. Mass spectrometric experiments were carried out with a Micromass Platform II (8327E) mass spectrometer, which was equipped with an electrospray co-axial probe as the ion source. Cone voltages were typically in the range of 1050V, the cone temperature was 5560C, nitrogen was used as the nebulizing gas and methanol as the mobile phase. For experiments involving mass spectrometric examination of bulk electrolyzed solutions, only a ten fold instead of the usual hundredfold concentration excess of electrolyte (Bu4NPF6) was used, in order to minimize signal suppression arising from the electrolyte. All electrolyzed samples were also diluted 1:100 in methanol before commencement of mass spectrometric measurement. For NMR experiments, Bruker AM300 or AC200 instruments were used. All chemical shifts are reported versus an internal TMS standard.
13

C spectra reported are proton broad band decoupled. For J-

modulated spin echo experiments a value of =7.1ms was chosen. Infra Red measurements were recorded in nujol mulls with a Perkin Elmer 1600 FT-IR. Elemental analyses were conducted by the Department of Chemistry, University of Otago, Dunedin, New Zealand. Solvents for electrochemical and spectroscopic measurements were of HPLC grade (Mallinckrodt, UltimAR or ChromAR). Dimethylformamide (DMF), as received, contained 0.005% water and was treated with activated 3 molecular sieves prior to experiments to reduce the water content. Reagent grade chemicals were used for synthesis of compounds. RuCl3`3H2O was

Chapter 3: Studies on Model Esters Compounds Experimental

page: 64

purchased from Pressure Chemicals Inc. and 2,2'-bipyridine-4,4'-dicarboxylic acid (H2-dcbpy) was kindly donated by Dr. Leone Spiccia and Prof. Glen B. Deacon of the Department of Chemistry, Monash University.

3.3. Synthesis
An overview about the synthesis of the (Et2-dcbpy)2RuX2 series of sensitizers is shown in Figure 3.2. Et2-dcbpy can be obtained in high yields by esterifying the acid, H2-dcbpy, in absolute ethanol with a catalytic amount of sulfuric acid. Reaction of Et2-dcbpy with RuCl3 also gives the desired chloride complex, cis-(Et2-dcbpy)2RuCl2, in good yield, although close attention has to be paid to the reaction conditions (see below). The remaining three complexes are synthesized in onepot reactions without isolation of (Et2-dcbpy)2RuCl2 by addition of appropriate salts (KCN, NH4SCN and KI) dissolved in DMF or DMF/H2O mixtures. In case of cyanide, de-esterification occurs due to the significant increase of the solution pH during the reaction and allows isolation of the protonated acid, which can be converted into the ester by refluxing in an absolute ethanol/sulfuric acid mixture.
EtOOC HOOC COOH H2SO4 cat. EtOOC COOEt COOEt

+
N N

RuIIICl3x3H2O 2h 125C, DMF EtOOC N

N Ru Cl

N
(II)

3d RF, C2H5OH abs

COOEt

H2-dcbpy

83% Et2-dcbpy HOOC COOH

Cl

72% (Et2-dcbpy)2RuCl2 1h 80C DMF/H2O + 20 KCN N COOH EtOOC COOEt 1h 80C DMF + 15 NH4SCN or 2h 100C DMF/H2O + 100 KI

N HOOC COOEt H2SO4 cat. N EtOOC N NC Ru N


(II)

N Ru
(II)

EtOOC

NC CN 74% (Et2-dcbpy)2Ru(CN)2 2d RF, C2H5OH abs EtOOC

N N X Ru

N
(II)

COOEt

COOEt

CN

65% (Et2-dcbpy)2Ru(CN)2

71% (Et2-dcbpy)2Ru(NCS)2 40% (Et2-dcbpy)2RuI2

Figure 3.2. Reaction scheme for the syntheses of the series of (Et2-dcbpy)2RuX2 sensitizers.

Et2-dcbpy: 500mg H2-dcbpy and 0.5mL concentrated H2SO4 were refluxed in 30mL absolute ethanol for 23 days. When the reaction mixture became clear the reaction was stopped and the solution was cooled, allowing the product to precipitate. The solution was filtered and crystalline white needles of Et2-dcbpy were collected and then washed with cold ethanol and diethylether. Drying in vacuum gave 510mg (83%) pure Et2-dcbpy. cis-L2RuCl2: 500mg (1.67mmol) Et2-dcbpy and 218mg (0.83mmol) RuCl3`3H2O were dissolved in ca. 35mL DMF. The solution was bubbled with nitrogen and then heated to 125C whilst being vigorously stirred. After 2 hours the reaction was stopped and the solvent removed by rotary evaporation. 100mL methanol was added to the solid residue and the mixture stirred over

Chapter 3: Studies on Model Esters Compounds Experimental

page: 65

night. Dark green-brown crystals were collected by filtration. 460mg (72%) NMR-pure product was obtained after drying in vacuum. Microanalysis (calculated): C: 49.8% (49.8%), H: 4.2% (4.2%), N: 7.0% (7.2%), Cl: 9.0% (9.2%) cis-L2Ru(NCS)2: The reaction was commenced as described for L2RuCl2. After 2 hours the reaction temperature was lowered to ca. 80C. 1.0g NH4NCS (13mmol) was added and the solution stirred for another hour. DMF was removed by rotary evaporation after which 100mL MeOH was added and the solution was stirred over night. Dark brown crystals were collected by filtration. 480mg (71%) NMR-pure product was obtained after drying in vacuum. Infra Red: CN=2093cm1. Microanalysis (calculated): C: 48.8% (49.9%), H: 3.9% (3.9%), N: 10.0% (10.3%) cis-L2RuI2: The reaction commenced as described for L2RuCl2. After 2 hours the reaction temperature was lowered to ca. 100C, 15g KI (90mmol) dissolved in 20mL H2O was added to the solution and stirring undertaken for two hours. DMF was removed by rotary evaporation and the solid dissolved in dichloromethane. The organic phase was washed with aqueous 0.01M NaOH, water, and dried over anhydrous Na2SO4. Evaporation of the solvent gave 320mg (40%) dark green-brown crystals. Microanalysis (calculated): C: 39.1% (40.2%), H: 3.4% (3.4%), N: 5.7% (5.8%), I: 25.7% (26.5%) cis-L2Ru(CN)2: The reaction commenced as described for L2RuCl2. After 2 hours the reaction temperature was lowered to ca. 80C, 1.0g KCN (15mmol) dissolved in 15mL H2O was added to the solution and stirring undertaken for another hour. The heating of the aqueous cyanide solution lead to an unavoidable increase in pH which eventually hydrolyzed the ester moieties. The solvent was then removed by rotary evaporation. The hydrolyzed residue was dissolved in 40mL H 2O and the strongly basic (pH~12) solution was titrated with 10% CF3COOH to pH 2, which precipitated (H2-dcbpy)2Ru(CN)2. The acid was separated from the solution by centrifugation and washed twice with acidified water (pH 2.2). Freeze drying yielded 410mg (74%) pure bright orange (H2-dcbpy)2Ru(CN)2`2H2O. Conversion to the ester was achieved by refluxing in 50mL absolute ethanol and 1.0mL concentrated H2SO4 for two days. An excess of water was added to the cooled solution, which was titrated to pH 1011 by addition of 1.0M NaOH. The aqueous phase was extracted with dichloromethane. The organic phase was washed with water and dried over anhydrous Na2SO4. After removal of the solvent, 400mg (65%) pure L2Ru(CN)2 was obtained as a brown powder. Infra Red: CN=2085, 1996cm1. Microanalysis (calculated): C: 54.4% (54.2%), H: 4.2% (4.3%), N: 11.1% (11.1%) 3.3.1. Synthetic Remarks: The reaction temperature and time were crucial for the synthesis of the cis-L2RuCl2 product. Below 120C no noticeable reaction occurred, above 130C the reaction proceeded much faster, but also led to significant decomposition, as did use of an extended reaction time. Thus, the reaction needed to be stopped at a critical time to optimize the yields and obtain a pure product.

Chapter 3: Studies on Model Esters Compounds Experimental

page: 66

Monitoring the UV/VIS spectra gave a very good indication of the progress of the reaction. Optimal relative intensities (absorbances) for the bands at 321nm (ligand) : 435nm (MLCT) : 591nm (MLCT) in DMF were found to be: 2.86:0.94:1 just prior to the onset of decomposition. It is also noteworthy that use of methods based on the use of the protonated ligand, H2-dcbpy, instead of Et2-dcbpy required a 1015C higher temperature which resulted in lower yields and less pure material. Although the literature [29, 70] reports state that reaction times of up to 8 hours are needed for this reaction at reflux temperatures of ~165C, it was found that close attention had to be paid to the actual reaction temperature and time to avoid unwanted side reactions, such as the formation of the tris compound ([Ru(Et2-dcbpy)3]2+). In the synthesis of the cyanide derivative, the reaction utilizing the esterified ligand proceeded smoothly to give higher yields and purity of the (H2-dcbpy)2Ru(CN)2 intermediate than when commencing the reaction sequence with the hydrolyzed acid. Both, UV/VIS and TLC monitoring showed that chloride substitution by strong -back bonding cyanide and thiocyanate ligands proceeded within a few minutes and that only a slight excess was needed. In contrast, the substitution of chloride by iodide required a very large excess (100 fold) of iodide along with higher reaction temperatures and longer reaction times. It is well known that formation of ruthenium bipyridine compounds containing thiocyanate yields an isomeric mixture of N- and S-bonded isomers [80], where the N-bonded thiocyanate is the thermodynamically favored form. Formation of the unwanted fully S-bonded isomer can be avoided in the synthesis of the photovoltaic sensitizer (H2-dcbpy)2Ru(NCS)2, but no synthetic route has been found yet to completely remove the mixed N- and S-bonded isomer. In the case of the esters, the methanol addition step quantitatively removes all S-bonded linkage isomers. The crude product obtained prior to purification with methanol shows a 1H-NMR resonance at 9.85ppm, assigned to the 6-H proton of the mixed N- and S-bonded bonded isomer. This resonance is absent after purification by methanol. All L2RuX2 products were chromatographed on silica gel 60 before being used in electrochemical experiments. The iodide complex was found to occasionally contain traces of the mixed Cl/I complex. L2RuI2 and L2Ru(NCS)2 were chromatographed with dichloromethane and ethylacetate 5:1 as the mobile phase. For L2RuCl2 a 1:1 mixture of dichloromethane and ethylacetate and for L2Ru(CN)2 a 6:1 mixture of acetone and ethanol was used. The first band contained the desired product. Yields obtained from these chromatography experiments were always greater than 80%.

Chapter 3: Studies on Model Esters Compounds Oxidation

Table 3.1. 1H and 13C-NMR chemical shifts (ppm) referenced versus TMS and coupling constants (Hz) data. Primed atoms are in the trans position relative to the halides.
a) b)

CDCl3 d6-DMSO c) CD3CN d) peak not resolved

page: 67

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 68

3.3.2. Characterization by NMR Spectroscopy Assignment of NMR signals (Table 3.1) was made on the basis of literature information [38, 80, 126] and J-modulated
13

C-NMR spectra. NMR data confirmed that all L2RuX2 complexes were

present in the cis configuration, but for convenience the cis prefix will be omitted in further discussion. The proton and carbon in the 6 position are in close proximity to the halides and therefore provide a direct probe of their electronic influence. Iodide which has the largest electronegativity as well as the largest van der Waal's radius, causes the largest chemical shift (10.82ppm) and splitting (=3.13ppm) of the H-6 and H-6' protons, whereas thiocyanate causes only a shift of 9.46ppm with a splitting of =1.65ppm. The C-6 and C-6' carbons both gave very broad
13

peaks, indicating a long relaxation time, with the C-6' carbon resonances sometimes not being resolved.

3.4. Oxidation
3.4.1. L2RuCl2 Voltammetric data obtained for the oxidation of a range L2RuX2 complexes and their derivatives are summarized in Table 3.3 on page 71. In general one one-electron metal centered oxidation processes were observed which are detailed below. (a) Voltammetry and Bulk Electrolysis Experiments in DMF The voltammetric oxidation of L2RuCl2 represents an almost ideal example of a chemically and electrochemically reversible process. L2RuCl2 [L2RuCl2]+ + e (3.1)

Figure 3.3a contains a series of cyclic voltammograms in DMF as a function of scan rate () at a 1mm diameter platinum macrodisk electrode. The rotating disk electrode voltammograms obtained at the same electrode material are shown as function of rotation rate () in Figure 3.3b. Data obtained under a range of conditions are summarized in Table 3.2. No additional processes are observed up to the solvent limit. Since the voltammetric oxidation of the other compounds is far more complex, data obtained for this system can be conveniently used as a reference. A plot of ipox versus 1/2 from cyclic voltammograms (ipox = oxidation peak current, = scan rate) is linear over the scan rate range of 10mV to 5000mVs1 implying that the process is diffusion controlled. Mass transport control was also demonstrated via the linear dependency of iL on 1/2 (iL = limiting current, = 2f = angular velocity) in rotating disk electrode experiments. The Er1/2values were calculated as described in Chapters 2.4 and 2.5 and are given in Table 3.2.

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 69

Cyclic Voltammetrya) / mVs1 10 26 50 100 200 500 1000 2000 5000 10000 Epox / mV 163 161 161 163 167 169 177 187 209 237 Epred / mV 103 101 107 103 101 97 91 83 59 33 Ep / mV 60 60 54 60 66 72 86 104 150 204 Er1/2 / mV 133 131 134 133 134 133 134 135 134 135 |ipox/ipred| 1.02 1.01 1.00 1.00 1.00 1.00 1.02 1.00 0.99 1.00

Rotating Disk Electrodeb) f / min


1

Microdisk Electrode / mV r / m 5.6 slope / mV 60 Er1/2 / mV +133

slope / mV 63 63 65 67 69 69

Er1/2

500 1000 1500 2000 2500 3000

134 138 141 142 144 146

Table 3.2. Voltammetric data obtained for the oxidation of 1.1mM L2RuCl2 in DMF (0.1 M Bu4NPF6) at a) 1mm diameter and b) 3mm diameter platinum macrodisk electrodes as a function of scan rate or rotation rate. All peak potentials are reported versus Fc/Fc+, uncertainty 2mV. For CV data Er1/2-values calculated as (Epox + Epred)/2, for steady state techniques slopes and Er1/2 calculated from "log-plots". ( = scan rate, f = rotation frequency, r = electrode radius, Ep = peak potential, Ep = Epox Epred , ip = peak current)
a)
2A

b)

20A

1A

c)

-600 -400 -200

200

400

600
+

800

potential [mV] vs. Fc/Fc

Figure 3.3. Oxidation of 1.1mM L2RuCl2 in DMF (0.1M Bu4NPF6) under conditions of a) cyclic voltammetry, =10, 26, 50, 100, 200 and 500 mVs1 b) rotating disk electrode, f=500, 1000, 1500, 2000, 2500 and 3000min1 and c) simulation of a cyclic voltammogram at a scan rate of 100mVs1 with parameters given in text. Full line: experiment, circles: simulation.

As expected for a reversible process, the Er1/2-value is independent of the voltammetric technique, except when a small contribution from uncompensated resistance is present, as is the case when fast rotation rates are used in rotating disk electrode experiments. Thus the reversible halfwave potential has a value of 1333mV vs. Fc/Fc+. Further confirmation of a mass transport

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 70

controlled reversible one-electron process was obtained from noting that the ratio of ipox/ipred (cyclic voltammetry) was close to unity (Table 3.2) and that of plots of E versus log((iL-i)/i) ("log plot") had values close to 58mV at 22C (rotating disk at slow rotation speeds and microdisk electrode voltammetry, Table 3.2). The almost completely ideal reversible behavior under the near steadystate conditions present at a microdisk electrode, compared to small departures from ideality with the other techniques implies that a small contribution from uncompensated resistance is present with techniques based on the use of macrodisk electrodes. Bulk oxidative electrolysis of L2RuCl2 at a platinum electrode using a controlled potential of 0.5V vs. Fc/Fc+ required between 15 to 20 minutes for completion (number of electrons transferred nox=1.000.03) and was accompanied by a color change of the solution from dark green to pale yellow. [L2RuCl2]+ as obtained from bulk electrolysis experiment in DMF is very stable although after a few hours a very slow color change of the solution from yellow back to green indicates that slow reduction of the oxidized complex occurs to regenerate the starting compound. After exhaustive electrolysis, rotating disk and microelectrode voltammograms were identical in slope and Er1/2-value as prior to commencing the experiment. However, as expected for a fully reversible system, the voltammetric response was now fully reductive instead of oxidative. Thus, on both short time voltammetric and long time synthetic conditions the reduction process was consistent with the reaction: [L2RuCl2]+ + e L2RuCl2 (3.2)

On reversing the direction of the bulk electrolysis experiment by applying a potential of 0.0V, the L2RuCl2 complex could be almost quantitatively recovered (nred=0.980.03). The measured iL-value obtained at a microdisk electrode and use of Equation 2.28 (page 48) enabled the diffusion coefficient to be calculated as D=3.1(0.2)106cm2s1. This value and the measured Er1/2-value (see above) were used to simulate cyclic voltammetric experiments. The result shown in Figure 3.3c demonstrates that excellent agreement with experimental data is obtained for this very well defined system. (b) Voltammetric Studies in other Solvents Studies in acetonitrile and dichloromethane also gave very well defined voltammetric responses and the reversible Er1/2-values can be found in Table 3.3. In acetonitrile a second oxidation process is detected just prior to the solvent limit. Under conditions of cyclic voltammetry, the peak potential for this process is located at +1.81V at a scan rate of 100mVs1 and the peak current magnitude is similar to the preceding initial oxidation process, which indicates that this process may be associated with a one-electron charge transfer step. However, the process does not become fully reversible in the chemical sense under conditions of cyclic voltammetry even when a scan rate of 2000mVs1 is used. The potential separation of the initial RuII/III process and this second process is in the range that would be expected if the second process is a RuIII/IV reaction [158]. However,

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 71

ligand based oxidation in this very positive potential region cannot be ruled out. Since the focus of this work is on the RuII/III process, this second process at very positive potentials in acetonitrile is not further discussed.
compound: L2RuCl2 solvent DMF CH3CN CH2Cl2 DMF CH3CN CH2Cl2 DMF CH3CN CH2Cl2 DMF CH3CN CH2Cl2 DMF DMF CH3CN CH3CN CH3CN CH2Cl2 D [106cm2s1] 3.10.2 7.20.2 5.50.2 3.50.3 6.40.2 5.80.35 a) 4.40.4 6.80.2 3.50.2 7.50.5 5.40.4 Er1/2 [mV] c) +1333 +1825 +1863 +2295 +2584 +2265 +42020b) +4706 +4513 +6034 +6458 +67510 +4606 +62015 +64010 +124030 +81010 +54015

L2RuI2

L2Ru(NCS)2

L2Ru(CN)2

[L2RuI(DMF)]+ [L2Ru(DMF)2]2+ [L2RuI(CH3CN)]+ [L2Ru(CH3CN)2]2+ [L2Ru(CN)(CH3CN)]+ [L2RuI]+

Table 3.3. Summary of voltammetric data obtained in different solvents at 202C. a) multielectron process, see text. b) peak potential obtained from differential pulse voltammetry. c) potential vs. Fc/Fc+.

(c) Spectroelectrochemical Studies The result of an OTTLE experiment on the oxidation of L2RuIICl2 to [L2RuIIICl2]+ is shown in Figure 3.4.

40

molar extinction [10 M cm ]

-1 3 -1

30

20

10

0 5000 10000 15000 20000 25000


-1

30000

35000

40000

energy [cm ]
Figure 3.4. Spectroelectrochemical monitoring of the course of oxidation of 0.7mM L2RuCl2 in an OTTLE electrochemical cell in DMF (0.1M Bu4NPF6).

Four isosbestic points are observed. Oxidation of L2RuCl2 (4d6) leads to an electron being removed from a metal orbital resulting in a 4d5 configuration (RuIII) and results in a collapse of both the MLCT bands present in the spectrum of L2RuCl2. However, after oxidation a chloride lone

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 72

pair electron is now able to be promoted into the empty t2g metal orbital, resulting in the detection of a ligand-to-metal charge transfer (LMCT) at 24400cm1. This change in electronic spectra is not unusual for metal based redox processes when the metal changes from a d6 to a d5 electron configuration [159]. Reduction of the initially oxidized solution regenerates the electronic spectrum observed with the starting compound. (d) ESR Spectrum of [L2RuCl2]+ The ESR spectrum of [L2RuCl2]+ in frozen DMF is shown in Figure 3.5.

0.20

0.25

0.30

0.35

0.40

0.45

field [T]
Figure 3.5. ESR spectrum of [L2RuCl2]+ a) in DMF glass at 77K, b) simulated spectrum with g =2.6103 (H=8.8mT), g =2.3972 (H=4.3mT), g =1.6801 (H=14.0).
b a

As expected for compounds of this symmetry, parallel and perpendicular components are found. Both the large linewidth and splitting between both components are typical for a ruthenium centered spin. However, the perpendicular component is split into two components (g =2.610 and
a

g =2.397) which indicates a deviation from true axial symmetry and splitting of the t2 electron
b

level. A sample of [(bpy)2RuCl2]+, which was generated under identical conditions as [L2RuCl2]+, exhibited an almost indistinguishable ESR spectrum to that shown in Figure 3.5. For RuIII(pic)3 (Hpic = picolinic acid) a similar pseudo axial symmetry has been found [160]. DeSimone and Drago [161] reported the ESR signal for [RuIII(bpy)3]3+ and noted a minor deviation from perfect axial symmetry. Unusual quadrupole splitting in Mssbauer [162] studies and X-ray adsorption [163] data showed that bipyridine is not such a rigid ligand as phenanthroline. In bipyridine, both aromatic rings are only connected by a single bond, which allows rotation out of the plane. In contrast, in phenanthroline, both rings are additionally connected via a double bond which

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 73

increases aromaticity and hence forces both pyridine rings into a single plane. No hyperfine structure due to coupling to
14

N is observed, which is an indication for the significant metal

character of the redox orbital. Most isotopes of ruthenium have spin S=0. Only two ruthenium isotopes are ESR active, but because of their high spin S=5/2 and low abundance, their signal, as in the case of [L2RuCl2]+, is often not observed. (e) Electrospray-MS Studies on Oxidized Solutions of L2RuCl2 The formation of the singly charged oxidized complex after bulk oxidative electrolysis was also detected by electrospray mass spectrometry. The main isotopic peak was found at 772.1 (m/z+) and the isotopic pattern clearly identified the cationic compound giving rise to the signal as having the chemical composition [L2RuCl2]+. (f) Voltammetry of Solid L2RuCl2 Attached to an Electrode Surface The high stability of oxidized [L2RuCl2]+ and the general hydrophobic character of diethylester bipyridine ruthenium complexes prompted me to investigate the solid state voltammetry of L2RuCl2 in aqueous solution. The voltammetric response obtained from redox reactions of solid material is often dominated by the crystallographic properties of the material, specifically by solidsolid phase transformation, nucleation and crystal growth processes [164, 164, 165]. In case of an oxidation of a neutral compound, the nature of the electrolyte anion is also of importance since it needs to be incorporated into the oxidized material to maintain charge neutrality. Thus, very often the electrolyte participates these kind of solid state reactions.

b)
50A

a)
50A

-0.5

0.0

0.5

1.0

potential [V] vs. Ag/AgCl (3M KCl)


Figure 3.6. Cyclic voltammetric oxidation of solid L2RuCl2 attached to a 5mm diameter carbon electrode in H2O. First five cycles, first cycle bold, =100mVs1, a) 0.1M KNO3, b) 0.1M LiClO4.

The cyclic voltammetric response for the oxidation of solid L2RuCl2 was investigate in water in the presence of the following electrolytes: NaF, NaCl, LiClO4, Na2SO4 and KNO3 (all at a concentration of 0.1M). The results for KNO3 and LiClO4 are shown in Figure 3.6. The presence of more

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 74

than one wave for the oxidation and reduction process suggest that complex morphologic transformations accompany the redox process. The stablest responses could be recorded with LiClO 4 as the electrolyte (see Figure 3.6a). After the first cycle all consecutive cycles traced over each other which is as expected for a highly reversible and insoluble system. In the presence of the other electrolytes successive cycling diminished the voltammetric peak height (Figure 3.6b). These results lead to the conclusion that in the latter case oxidized [L2RuCl2]+ dissolves into the solution and consequently L2RuCl2 is removed from the electrode. 3.4.2. L2RuI2 (a) Voltammetry in DMF Cyclic voltammograms as a function of scan rate (Figure 3.7a) show, that in contrast to the case with L2RuCl2, the oxidation process is chemically irreversible at slow scan rates (<100mVs1), although chemical reversibility is found at scan rates 200mVs1, where ipox/ipred values are close to unity (Table 3.4).
1.5 1.0 current [A] 0.5 0.0 -0.5 -1.0 2.0 current [A] 1.0 0.0 -1.0 800 current [pA] 600 400 200 0 -0.6 -0.3 0.0 0.3 0.6
+

a)

A B

b)

F C D E

c)
0.9 1.2

potential [V] vs. Fc/Fc

Figure 3.7. Oxidation of 0.75mM L2RuI2 in DMF (0.1M Bu4NPF6) at a platinum electrode a) as a function of scan rate: =10mVs1, 26mVs1, 50mVs1 and 100mVs1. b) consecutive scans at =50mVs1. c) microdisk electrode r=5.6m.

Under conditions where the process is chemically irreversible, a second oxidation process is observed at a potential of about 230mV more positive than the initial process. A third response at +620mV is also observed at slow scan rates lower than 50mVs1. An analogous second (or third) response is not observed in the shorter time domain rotating disk or microdisk electrode (Figure 3.7c) experiments. From rotating disk electrode experiments mass transport control of the initial process at short time domains was confirmed by a linear plot of iL versus 1/2, which passed

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 75

through the origin. Additionally the slope of a plot of log((iL-i)/i) versus potential was close (see Table 3.4) to the theoretical value of 58mV expected for an electrochemically reversible oneelectron transfer process at 20C. This latter plot also allows the Er1/2-value to be determined in a similar fashion to that for L2RuCl2 and the data obtained are summarized in Table 3.4. L2RuI2
Cyclic Voltammetry Rotating Disk Electrode slope / mV Er1/2 / mV
r / mVs1 E 1/2 / mV |ipox/ipred| f / min1

10 26 50 100 200 500 1000 2000 5000

219 227 230 230 231 231 232 232 234

1.95 1.36 1.17 1.09 1.07 1.02 1.04 1.04 1.16

67 229 68 232 67 234 66 235 67 237 67 237 Microdisk Electrode r / m slope / mV Er1/2 / mV 5.6 65 +232

500 1000 1500 2000 2500 3000

L2Ru(CN)2
Cyclic Voltammetry / mVs 10 26 50 100 200 500 1000 2000 5000
1

Rotating Disk Electrode f / min


1

Er1/2

/ mV

ipox/ipred 1.79 1.63 1.49 1.46 1.20 1.17 1.12

slope / mV

Er1/2 / mV

598 600 601 604 603 604 606

72 601 73 603 73 604 75 605 76 607 75 607 Microdisk Electrode r / m slope / mV Er1/2 / mV 5.6 70 604

500 1000 1500 2000 2500 3000

L2Ru(NCS)2a)
Cyclic Voltammetry / mVs1 Epox / mV 10 100 500 530 556 600

Table 3.4. Voltammetric Data obtained for the oxidation of L2RuX2 in DMF (0.1M Bu4NPF6) at platinum electrodes. a) irreversible oxidation process.

The detection of additional processes under slow scan rate cyclic voltammetric conditions (Figure 3.7b), indicates that new products are formed by a chemical reaction following the charge transfer step which generated [L2RuIIII2]+. For other complexes containing easily oxidizable ligands [166, 167], electrochemically generated RuIII complexes have been shown to undergo an internal reaction to produce RuII and the oxidized ligand [168]. In the particular case of L2RuI2 the oxidized form of the ligand is I2 which itself is electroactive in the potential range of interest [169]. The redox chemistry of the iodide/iodine system has been classically described as

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 76

6I

2I3 +

4e 2e

(3.3) (3.4)

2I3 3I2 +

In Figure 3.8. cyclic voltammograms of I and I2 in DMF solution are shown. Iodide is already oxidized in DMF at a potential of 0.01V, that is 0.24V more negative than the oxidation of L2RuI2. However, because of surface based reactions [170], the voltammograms are considerably more complex than expected by inspection of Equations 3.3 and 3.4 It should be noted that according to steady-state measurements about 10% of I2 was present in a reduced form in the DMF solution.

a)
1A

b)
0.5A

-0.5

0.0

0.5

potential [V] vs. Fc/Fc

1.0

Figure 3.8. Cyclic voltammograms obtained for the iodine/iodide system in DMF (0.1M Bu4NPF6) at a 1mm diameter platinum electrode. a) 1.1mM I2, =100mVs1, b) 1.0mM Bu4NI, =100mVs1.

The following kind of reaction scheme therefore may be postulated in order to rationalize the electrochemical oxidation of L2RuI2. E: C': C'': E': C''': C'': E'': [L2RuIII(DMF)]+ L2RuIII2 [L2RuIIII2]+ + e [L2RuIII(DMF)]+ + I (3.5) (3.6) (3.7) + e + I (3.8) (3.9) (3.10) + e (3.11)

DMF + [L2RuIIII2]+  2I

 I2

[L2RuIIII(DMF)]2+ 

DMF + [L2RuIIII(DMF)]2+ 2I

L2RuII(DMF)2]2+

 I2

[L2RuII(DMF)2]2+ [L2RuIII(DMF)2]3+

where E0 and k-values are formal reversible potential and rate constant for the relevant reaction. The cyclic voltammetric behavior for this reaction scheme can be readily simulated if the

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 77

contribution from the highly complex iodine system is omitted (Eqns. 3.7 and 3.10). With this approximation, good fits between experimental and simulated voltammograms were obtained over the scan rate range of 25 to 1000mVs1. Figure 3.9 shows a comparison of the first five cycles of a cyclic voltammogram obtained by experiment and digital simulation at a scan rate of 500mVs1. The best fit to the experimental data over the scan rate range of 25 to 1000mVs1 was obtained with ka=101.30.3s1 (Keql = 1000), kb=102.30.5s1 (Keql = 1000), E10=+0.229V, E20=+0.46V and E30=+0.62V. The third electrochemical step, which is attributed to the oxidation of the bis-DMF complex (Eqn. 3.11) is only visible at very slow scan rates 50mVs1 (processes B and C in Figure 3.7b) in both experimental and simulated digital cyclic voltammograms. The voltammetry of iodine is detected in experimental results and marked as processes F and E in Figure 3.7b (scan rate 50mVs1) but is also almost absent at higher scan rates (compare Figure 3.9).
8 6 4 2 0 -2 -4 -6 8 6 4 2 0 -2 -4 -6 -0.2 0.0 0.2 0.4 0.6 0.8

current [A]

a)
D

current [A]

b)

potential [V] vs. Fc/Fc

1.0

Figure 3.9. Comparison of experimental and simulated cyclic voltammograms obtained from the oxidation of 2mM L2RuI2 in DMF (0.1M Bu4PF6) at a scan rate of 500mVs1 with parameters given in text. a) experiment, b) simulation.

(b) Bulk Electrolysis in DMF Exhaustive controlled potential oxidative bulk electrolysis of a 0.9mM solution of L2RuI2 in DMF at +200mV (Er1/2=229mV) consumes 1.00.1 electrons per molecule. The potential used for the electrolysis was chosen to avoid electrochemical oxidation of products formed by reaction of [L2RuI2]+ and DMF (Eqn 3.6). Figures 3.10a and b show rotating disk voltammograms obtained before and after electrolysis. As expected from data obtained from cyclic voltammograms of L2RuI2,

on the bulk electrolysis time scale, the mono-DMF substituted complex, [L2RuI(DMF)]+,

was formed according to Equations 3.5-3.7, and it is this complex that gives rise to the reversible oxidative wave at about +460mV. The preceding wave at ca. +180mV is due to formation of I2, the oxidized ligand. However, the iodine is present as a mixture of I2 and I3, since the potential for bulk electrolysis is close to that expected for the I2/I3 process.

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 78
2+/+

[L2RuI(DMF)]

[L2Ru(DMF)2]

3+/2+

[L2RuI2] I2/I3
-

+/0

c)
20A

b)

a)
-0.4 -0.2 0.0 0.2 0.4 0.6 0.8
+

1.0

potential [V] vs. Fc/Fc

Figure 3.10. Rotating disk voltammetry (3000rpm) before and after oxidative bulk electrolysis of 0.9mM L2RuI2 in DMF (0.1M Bu4NPF6). a) before electrolysis, b) after exhaustive electrolysis at +200mV, c) after exhaustive electrolysis at +440mV

Analysis of plots of log((iL-i)/i) versus potential for the oxidation of [L2RuI(DMF)]+ in DMF obtained from rotating disk voltammograms (5003000rpm) after bulk electrolysis yields a slope of 653mV and an E1/2 of +0.460.01V. However, the limiting current, iL, was only moderately reproducible and 8510% of that of the starting compound, implying that some L2RuX2 material has been lost during the course of bulk electrolysis experiments. Exhaustive oxidative bulk electrolysis at the more positive potential of +440mV consumed a further 1.10.2 electrons per molecule. From Equations 3.103.11 the formation of the bis-DMF complex, [L2Ru(DMF)2]2+, is expected on this long time scale experiment. A rotating disk voltammogram of the more extensively electrolyzed solution is shown Figure 3.10c. The oxidative wave now observed at +0.63V is associated with the formation of the new [L2Ru(DMF)2]2+ product. The iodine wave is now fully reductive and double in height relative to that observed after electrolysis at +200mV as expected when a second equivalent of iodine has been formed as predicted by Equation 9. "Log-plot" analysis of data obtained from rotating disk electrode voltammograms for the oxidation of [L2Ru(DMF)2]2+ gives a slope of 655mV. The limiting current for this process is 8010% of that of the [L2RuI(DMF)]2+ complex which implies loss of L2RuX2 material via other unknown reaction pathways also has occurred on the time scale of bulk electrolysis. (c) ES-Mass Spectrometric Studies on Bulk Electrolyzed Solutions in DMF Further evidence for the formation of the solvent complexes was obtained using electrospray mass spectrometry. A positive ion mode electrospray mass spectrum of the exhaustive bulk electrolyzed DMF solution of L2RuI2 at +200mV is shown in Figure 3.11a. The major peak occurs at 902.1m/z+ and is assigned to a singly charged species which exhibits the typical ruthenium

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 79

isotopic pattern. Simulation (Figure 3.11a insert) shows the excellent match with the expected spectrum of the mono-DMF substituted complex [L2RuI(DMF)]+. The signal at 861.1m/z+ has been identified as arising from [L2RuI(CH3OH)]+ where the DMF solvent has been replaced by the methanol carrier solvent introduced when undertaking the mass spectrometric experiments. The signal at 1289.4m/z+ is due to adduct formation with the Bu4NPF6 electrolyte to give [L2RuI(DMF)(Bu4NPF6)]+ and the signal at 1198.3m/z+ is assigned to adduct formation of unreacted starting material with the electrolyte cation to give [L2RuI2(Bu4N)]+..
902.1

424.2

a)
1016.9 1198.3

b)
861.1

401.7

493.1

890

900

910

420

425

430

800

1000

m/z m/z Figure 3.11. Electrospray mass spectra obtained from bulk oxidatively electrolyzed solutions of L 2RuI2 in DMF (0.01M Bu4NPF6). After exhaustive electrolysis at a) +200mV. b) +440mV. Inserts show measured (bottom) and simulated (top) spectra of the solvent complexes [L2RuI(DMF)]+ at 902.1m/z+ and [L2Ru(DMF)2]2+ at 424.1m/z+. See text for additional experimental details.

1200

1400

300

400

500

600

A positive ion mode electrospray mass spectrum of the solution exhaustively electrolyzed at 440mV is shown in Figure 3.11b. The main signal is now found at 424.2m/z+. Again simulation (Figure 3.11b insert) provides the excellent agreement with the signal expected for the doubly charged bis-DMF substituted complex [L2Ru(DMF)2]2+. A weak signal attributable to a doubly charged species with a ruthenium isotope pattern is seen at ca. 402m/z+ and is consistent with the spectrum from [L2Ru(DMF)(CH3OH)]2+. The mass spectrometric results imply that the DMF complexes have considerably stability even in solutions containing a hundredfold excess of CH3OH. (d) Studies in Other Solvents Cyclic voltammograms of L2RuI2 in CH2Cl2 and CH3CN gave a well defined initial metal centered oxidation in both solvents which is reversible at fast scan rates. However, in both solvents this initial process becomes irreversible at slow scan rates in which case it is followed by a series of overlapping processes. When the rotating disk or microdisk electrode techniques are used, "logplot" analysis of the first oxidation process in both solvents under these short time domain conditions again was consistent with a reversible one-electron charge transfer process (Table 3.3). Exhaustive bulk electrolysis of L2RuI2 at +250mV in CH3CN produced the mono-substituted complex [L2RuI(CH3CN)]+ as determined by ES-MS with the major signal at 870.1m/z+, although the yield was only about 70%. An additional signal derived from the presence of a singly charged ruthenium compound of unknown identity was observed at 929.3m/z+. The signal expected from formation the CH3OH substituted complex, [L2RuI(CH3OH)]+, at 861m/z+ was not observed in the

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 80

presence of methanol added prior to undertaking mass spectrometric measurements, indicating that CH3CN forms even more stable complexes than DMF with this type of ruthenium compound. From rotating disk electrode measurements undertaken on the electrolyzed solution the reversible halfwave potential for the oxidation of [L2RuI(CH3CN)]+ was calculated as Er1/2=0.64V. Bulk electrolysis carried out at a potential of +610mV formed a low yield of the [L2Ru(CH3CN)2]2+ complex as verified from electrospray mass spectra via observation of a signal of a doubly charged ruthenium compound at 392.1m/z+. The oxidation wave of this complex was partly overlapped with other unidentified processes, but was determined to have a value of E1/2=1.24V from rotating disk electrode experiments. Oxidative bulk electrolysis of L2RuI2 in CH2Cl2 at +240mV proceeded in a similar manner to bulk electrolysis in DMF and CH3CN and consumed 0.90.2 electrons per molecule. After the exhaustive electrolysis, a new oxidation wave at 0.540.01V was observed using rotating disk voltammetry. Electrospray mass spectrometry did not show a signal at 913.0m/z+ as expected for the mono substituted solvent complex [L2RuI(CH2Cl2)]+; rather the main signal was located at 829.0m/z+, which is attributable to the formation of the five coordinate unsubstituted [L2RuI]+ complex. A signal at 861.1m/z+ is assigned to formation of [L2RuI(CH3OH)]+. Although there is no direct mass spectrometric evidence for the existence of the mono substituted CH2Cl2 complex [L2RuI(CH2Cl2)]+, it might still be present in solution phase, since a weakly coordinating solvent molecule of this kind could have been removed in the electrospray process. 3.4.3. L2Ru(CN)2 (a) Voltammetry and Bulk Electrolysis Experiments in DMF As was the case with L2RuI2, cyclic voltammograms of the cyanide analogue recorded at slow scan rates were not fully reversible until scan rates >200mVs1 were used. Cyclic voltammograms at a platinum macrodisk electrode as a function of scan rate are shown in Figure 3.12 and data are summarized in Table 3.4. Data obtained from microdisk electrode and rotating disk electrode experiments (Table 3.4) showed almost ideal behavior for a reversible one-electron oxidation process. The Er1/2-value for the [L2Ru(CN)2]0/+ process was determined to be 0.60V, which is very much more positive than those found for the chloride and iodide derivatives. The peak potential for oxidation of cyanide present in a 1mM solution of KCN in DMF (0.1M Bu4NPF6) occurs at +0.63V (=200mVs1, platinum electrode). The process is irreversible, very broad and commences at ca. 0.0V. A mechanism related to that for the oxidation of L2RuI2 under conditions where the process is chemically irreversible, would lead to the formation of [L2Ru(CN)(DMF)]+ and [L2Ru(DMF)2]2+. However, the bis-DMF complex has a reversible halfwave potential of +0.62V and under conditions of controlled potential electrolysis of L2Ru(CN)2 (Er1/2=+0.60V) would be immediately oxidized to highly reactive [L2Ru(DMF)2]3+.

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 81

1.5

current [A]

1.0 0.5 0.0 -0.5 -1.0 0.8

a)

current [nA]

0.6 0.4 0.2 0.0 -0.4 0.0 0.4 0.8


+

b)
potential [V] vs. Fc/Fc

Figure 3.12. Oxidation of L2Ru(CN)2 in DMF (0.1M Bu4NPF6). a) As a function of scan rate at a platinum 1mm electrode, =10, 26, 50, 100 and 200mVs1. b) Under near steady state conditions at a platinum microdisk electrode r=5.6m.

Bulk electrolysis of a 0.9mM L2Ru(CN)2 solution in DMF (0.1M Bu4NPF6) at a platinum electrode was carried out with Eappl=+500mV which corresponds to the foot of the initial oxidation process and where subsequent oxidation of products formed by EC type processes are expected to be minimized. The exhaustive oxidation of L2Ru(CN)2 at this potential requires the transfer of two electrons (n=2.050.1) and leads to a color change of the solution from dark purple to pale yellow, which may be indicative of the formation of RuIII species.
[L2Ru(CN)2] 0.5A
0/+

a)

[L2Ru(NCS)2] 0.2A

0/+

b)

-0.4

-0.2

0.0

0.2

0.4

0.6
+

0.8

1.0

potential [V] vs. Fc /Fc


Figure 3.13. Oxidative bulk electrolysis of L2Ru(NCS)2 in DMF monitored by differential pulse voltammetry which demonstrates the conversion into L2Ru(CN)2 and the formation of a new product, a) 0.9mM L2Ru(CN)2, Eappl=+0.50V, b) 1.0mM L2Ru(NCS)2, Eappl=+0.27V.

The electrolysis was monitored by differential pulse voltammetry (Figure 3.13a) which reveals the formation of a new product which can be reduced at +0.30V. Steady-state microdisk electrode voltammetric techniques showed that this new product was formed in small yields of less than

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 82

20%. Upon reduction of the bulk electrolyzed solution at 0V, the color changed again to dark purple, indicating the formation of a RuII species. (b) ES-Mass Spectrometric Studies in DMF Electrospray mass spectrometric studies on bulk oxidized L2Ru(CN)2 solutions in DMF initially showed only one major product peak with a typical ruthenium isotopic pattern at 762.60m/z+. The peak pattern indicated that this resulted from a doubly charged tri-nuclear ruthenium species. The fact that the intensity was insensitive to the cone voltage, indicates that a compound present in the solution phase produces the spectrum and that it is not from a compound formed in the electrospray source. Furthermore, this response remained after reduction at 0.0V, so it is not associated with the product that gives the voltammetric process at +0.30V. After the bulk electrolyzed solution was left to stand for several minutes, the appearance of several new peaks assigned to bi- and trinuclear ruthenium compounds were seen between 600m/z+ and 1600m/z+, whilst the peak at 762.60m/z+ disappeared. No ruthenium patterns were found in the negative ion mass spectrum mode. The uncharacterized products formed by electrochemical oxidation of L2Ru(CN)2 are believed to be complexes formed by reaction with cyanogen [171], the oxidized form of the ligand (CN)2. (c) Studies in other Solvents In CH2Cl2 higher scan rates (>1000mVs1) are required than in DMF to achieve complete chemical reversibility for the RuII/III oxidation process. No additional processes are observed for oxidation of [L2Ru(CN)2]+ up to the solvent limit at +1.6V. The voltammetry in CH3CN is similar with no additional processes for oxidation of [L2Ru(CN)2]+ being observed up to +2.0V. Oxidative bulk electrolysis of L2'Ru(CN)2 at +605mV in CH3CN occurred with the transfer of n=0.90.1 electrons per molecule. Rotating disk electrode voltammetric measurements made during the course of electrolysis show the disappearance of the L2Ru(CN)2 oxidation process and the appearance of a new oxidative wave at 0.89V. The limiting current at the rotating disk electrode for this wave was only about 30% of that for oxidation of L2Ru(CN)2. ES-mass spectra show one major peak at 770.3m/z+ and the mass distribution indicates that this response arises from a doubly charged binuclear ruthenium species and is consistent with the formula [L2Ru(CN)(CH3CN)]22+. Exhaustive oxidative electrolysis of this solution at +805mV consumed a further 0.60.2 electrons per molecule and voltammetric results were similar to those obtained in DMF in the sense that the disappearance of the voltammetric response at 0.89V was noted and the only new well defined voltammetric process was a reductive wave at 0.35V with the limiting current being less than 20% of that observed on oxidation of the initial solution of L2Ru(CN)2. Furthermore, electrospray mass spectra again showed only one major peak at 762m/z+ consistent with the presence of a tri-nuclear species. Thus, the products formed upon exhaustive oxidation of L2Ru(CN)2 appear to be solvent independent and the reaction pathways are different to those observed for L2RuI2, since no evidence for the formation of the solvent complexes has been found.

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 83

3.4.4. L2Ru(NCS)2 (a) Voltammetry and Bulk Electrolysis Experiments in DMF Under conditions of cyclic voltammetry, the oxidation of L2Ru(NCS)2 in DMF was chemically irreversible at all scan rates used up to 10000mVs1. Minor processes following the initial metal based oxidation are believed to be caused by oxidation of surface based products, although these processes were less pronounced in DMF than after oxidation of (H2-dcbpy)2Ru(NCS)2.12 Since the oxidation process is not reversible the Er1/2-value in DMF cannot be calculated and only peak potentials (Table 3.4) are available. As observed for (H2-dcbpy)2Ru(NCS)2,12 the magnitude of the limiting current per unit concentration obtained via steady state techniques indicate that the oxidation process corresponds to an n>1 electron process. Oxidative bulk electrolysis (Eapp=+270mV) of 1.0mM L2Ru(NCS)2 in DMF yields the corresponding cyanide complex L2Ru(CN)2 (Equation 3.12). Kohle and coworkers [81] have reported a slow conversion of (H2-dcbpy)2Ru(NCS)2 to the cyanide analogue upon illumination. The analogy with electrochemical oxidation is appropriate since a photo excitation involves a metal-to-ligand charge transfer (MLCT) in which the metal changes from a d6 to a d5 electron configuration, which is equivalent to a metal based oxidation. L2Ru(NCS)2 L2Ru(CN)2 + oxidized "S" + ne (3.12)

The number of electrons transferred during the course of the bulk electrolysis was 62 electrons per molecule and the large value is a consequence of oxidation of sulfur, which exists in the oxidation state 2 in thiocyanate and can be oxidized up to +3, depending on the reaction pathway [172]. The transformation described by Equation 3.12 was monitored by differential pulse voltammetry (Figure 3.13b) during the course of the bulk electrolysis (Eapp=+270mV). Prolonged and further oxidation at a potential of Eappl=+500mV yielded the same product of unknown identity at 0.30V as obtained by direct oxidation of L2Ru(CN)2 (Figure 3.13a). It is noteworthy, that under no conditions was the formation of the mixed thiocyanate-cyanide complex, L2Ru(NCS)(CN), observed. Assuming ligand additivity [157, 173], the potential for this complex would be expected to lie between those for the thiocyanate and cyanide complex at +0.51V. However, even when the applied oxidation potential for the thiocyanate complex was set at +0.27V, only the formation of the bis cyanide complex was observed. This result implies that the sulfur elimination reaction may occur via formation of a cyclic intermediate and some form of rearrangement, as the thiocyanate ligand flips from N- bond to C- bond after the loss of sulfur. Both thiocyanate groups are in close vicinity, crystallographic data [38] report an angle of 86 degrees between both -NCS groups and the near linearity of the -NCS groups. If a metal centered oxidation occurs initially and the vacant electron from the t2g orbital is filled by an electron, the NCS group would lose its linear structure and the terminal sulfur could bend and interact with the
12

For more details on surface based processes accompanying the oxidation of (H2-dcbpy)2Ru(NCS)2 see Chapter 4.

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 84

other -NCS group. The thiocyanate ligands are thermodynamically capable of reducing the Ru(III) center, since free thiocyanate can be oxidized at a lower potential than L2Ru(NCS)2 (see Chapter 4.3.4 on page 124). Transition products are probably associated with the surface based phenomena, not observed with the other complexes. The large number of post waves, especially at low temperatures (5 superimposed post waves following the initial oxidation process can be observed in acetone at 60C), imply that a very complex reaction pathway is associated with the conversion of L2Ru(NCS)2 to L2Ru(CN)2. (b) Studies in other Solvents The oxidation of L2Ru(NCS)2 in CH2Cl2 at a platinum macrodisk electrode under conditions of cyclic voltammetry is shown in Figure 3.14.
3
ox

current [A]

1
0 0 10 20 30 40
1/2

ip [A]

50

60
1/2 -1/2

70

80

0 -1 -2 -3

(scan rate)

[mV s

-0.5

0.0

0.5
+

1.0

potential [V] vs. Fc/Fc

Figure 3.14. Cyclic voltammograms (=100mVs1) for the oxidation of 1.0mM L2Ru(NCS)2 in CH2Cl2 (0.2M Bu4NPF6) at a 1mm platinum electrode. Full line: measured. Circles: simulated using parameters described in the text. Insert: Anodic peak current (ipox) versus 1/2.

As was the case with L2RuCl2 in CH3CN, a second process (not shown) with a comparable peak current to the initial oxidation process is observed at +1.5V at a scan rate of 200mVs1. This process is not fully reversible and might be a metal centered RuIII/IV or a ligand based thiocyanate oxidation process. The cyclic voltammetric responses obtained in this non-coordinating solvent suggest the presence of a stable product since even at scan rates as low as 10mVs1 a reverse peak is found. However, careful examination of data obtained over the scan rate range of 105000mVs1 reveals that diffusion controlled mass transport is not achieved because plots of ipox and ipred versus 1/2 are not linear (see insert Figure 3.14) and the peak-to-peak potential separation at slow scan rates is 48mV which is less than that observed for the ideal model compound L2RuCl2 and as expected theoretically for a one-electron process. In contrast, when the microdisk or rotating disk electrode techniques are used, the L2Ru(NCS)2 oxidation process corresponds to a chemically reversible mass transport controlled one-electron process. That is, slopes calculated from "log plots" obtained from these techniques were 675mV and similar to those obtained for the oneelectron oxidation of L2RuCl2 under the same conditions. Furthermore, data obtained from rotating disk electrode voltammograms show that a plot of iL versus 1/2 was linear and passed through the

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 85

origin. From data obtained under the latter conditions values of D and Er1/2 may be calculated (Table 3.3). Cyclic voltammograms simulated using these values are compared to a recorded voltammogram in Figure 3.14. The substantial difference between the mass transport controlled (simulation) and the measured voltammogram can be clearly seen. Not only is the calculated halfwave potential different by 5mV, the experimentally measured peak current is almost double than predicted by the simulation. Apparently reactant and product adsorption provide a major contribution to the cyclic voltammetric response, as is the case with oxidation of (H2-dcbpy)2Ru(NCS)2 (see next chapter).
5V/s

a)

5A

100V/s

b)

0.2A

500V/s

c)

0.5A

1000V/s

d)

1A

-0.5

0.0

0.5

1.0

potential [V] vs. Pt-wire


Figure 3.15. Fast scan cyclic voltammetry for the oxidation of L2Ru(NCS)2 in CH3CN (0.1M Bu4NPF6) at platinum electrodes. a) =5Vs1, 500m electrode diameter, b) =100Vs1, 70m electrode diameter, c) =500Vs1, 70m electrode diameter, d) =1000Vs1, 70m electrode diameter

In cyclic voltammetric studies in CH3CN, whilst the peak current ratio ipox/ipred is almost unity, the peak-to-peak separation Ep=405mV for scan rates between 10 and 100mVs1, which is significant smaller than theoretically expected for a one-electron charge transfer process and as observed for the ideal L2RuCl2 model system. Further, ipred does not scale with 1/2 and at fast scan rates >50Vs1 ipred is almost independent of scan rate which is illustrated in Figure 3.15. Data obtained from fast scan cyclic voltammetric experiments in CH3CN shown in Figure 3.15 also show that the oxidation peak current, ipox, scales approximately linear with the amount of background current, which is dominated by the double layer charging current under these conditions.

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 86

However, the charging current scales linearly with scan rate, hence the ipox scales linearly with the scan rate as well, as would be expected when the process is controlled by product adsorption and not by diffusion (compare Equation 2.18). Additionally, at 40C the reduction peak potential shifts to a very much more negative potential, resulting in a Ep-value of 24020mV. This is, surface based processes perturb the oxidation process as is the case in CH2Cl2 and will be shown for (H2-dcbpy)2Ru(NCS)2 in Chapter 4. However, again when the microdisk electrode or the rotating disk electrode are used, the oxidation of L2Ru(NCS)2 in CH3CN is chemically reversible and mass transport controlled as evidenced from slopes of "log-plots" of 633mV and plots of iL versus 1/2 which pass through the origin in the case of rotating disk electrode experiments. From these log-plots the reversible half-wave potential in CH3CN can be calculated and the value is listed in Table 3.3. Exhaustive bulk electrolysis of L2Ru(NCS)2 in CH3CN proceeds as in DMF, with the formation of the cyanide analogue, L2Ru(CN)2, confirmed by ES-MS. 3.4.5. Spectroscopic Considerations In a recent XPS study [94] on (H2-dcbpy)2Ru(NCS)2 it was suggested that the oxidation of (H2dcbpy)2Ru(NCS)2 might be thiocyanate rather than metal based as assumed for other derivatives. Comparison of data obtained by work described in chapter 4 shows that the reversible potential for the oxidation of (H2-dcbpy)2Ru(NCS)2 is almost unaffected by esterification of the carboxylate groups so that potential data obtained from esters may be used as a basis for comparison. A linear relationship is found (see Figure 3.16) between the energy of both MLCT bands (values taken from Table 3.7b on page 98) and the reversible potential for the [L2RuX2]0/+ process in DMF using Er1/2values obtained from differential pulse voltammograms (lower energy band: slope 4.3103cm1V1, intercept 2.3104cm1, r=99.7%; higher energy band: slope 4.0103cm1V1, intercept 1.6104cm1, r=99.8%). The linear relationship is also obtained when CH3CN is chosen as the solvent. No unusual feature in the value for the reversible potential for the [L2Ru(NCS)2]0/+ process emerges from this comparison.
2.5 2.4

energy [10 cm ]

-1 4

2.3

1.9 1.8 1.7 0.1 0.2 0.3 0.4 0.5 0.6

potential [V]
Figure 3.16. Plots of the oxidation potential of the [L2RuX2]0/+ process in DMF versus the energy of each of the two MLCT bands. Data from Table 3.3. and Table 3.7b. Points are (left to right) for X=Cl, I, NCS and CN.

Chapter 3: Studies on Model Esters Compounds Oxidation

page: 87

The MLCT transition reflects the relative energy levels of the metal center and the participating ligand [157, 174]. In the case of the bipyridine ligands an electron from the metal -orbital becomes promoted into an empty *-bpy orbital during the transition. Since the bipyridine part stays unchanged in the L2RuX2 series, the change in the MLCT transfer bands should directly reflect the change in -orbital energy and hence the reversible oxidation potential. If the oxidation process is metal centered, the change in the MLCT transfer transitions should scale linearly with the change in oxidation potential. Thus, this linear relationship implies that it is highly likely that the oxidation of L2Ru(NCS)2 is indeed a RuII/III oxidation rather than a thiocyanate ligand oxidation process.

3.5. Reduction
Voltammetric data obtained for the reduction of a range L2RuX2 complexes and their derivatives are summarized in Table 3.5 on page 90. Up to four one-electron ligand based reduction processes were observed which are detailed below. 3.5.1. Et2-dcbpy The free ligand Et2-dcbpy exhibits one reversible one-electron reduction process under conditions of cyclic voltammetry in DMF (0.1M Bu4NPF6). A cyclic voltammogram is displayed in Figure 3.17a and from a series of voltammograms (data not shown) the reversible half-wave potential can be calculated to be 2.090.01V vs. Fc/Fc+. This value is in reasonable agreement with the literature value of 2.05V [142] (potential converted from SCE by assuming 0.45V for Fc/Fc+ vs. SCE in DMF (0.1M Bu4NPF6) [175]). Another study [176] has shown, that under ultrainert conditions in tetrahydrofuran two reversible one-electron reduction processes are accessible at room temperature and three at 74C. The reduction of Et2-dcbpy occurs more than 0.5V more positive than unsubstituted bipyridine under similar conditions (Er1/2=2.55V vs. Fc/Fc+ [142]), which is as expected when the electronic influence of the carboxylate groups is considered. Figure 3.17b shows a cyclic voltammogram of the reduction of Et2-dcbpy recorded under identical conditions as in Figure 3.17a, but in the presence of atmospheric oxygen. The preceding reductive process with a peak potential at 1.56V can be attributed to the reduction of oxygen and formation of peroxide according to Equation 3.13. O2 + e O2 (3.13)

The reduction process of Et2-dcbpy is now irreversible in the presence of oxygen. Apparently Et2-dcbpy reacts with oxygen or peroxide in an EC type reaction. The reaction of electrogenerated peroxide with voltammetrically reduced aromatic pyridine esters has been studied in depth by Webster and co-workers [177, 178]. They found that de-esterification occurred according to the following reaction scheme by reaction of the electrogenerated peroxide anion with the ester radical:

Chapter 3: Studies on Model Esters Compounds Reduction


O C R OR' O

page: 88

O2

C O

OR'

R'O

(3.14)

O C R O O

O C

+ O2

+ O2

(3.15)

Due to the adverse effect of oxygen present in solution on the voltammetry of Et2-dcbpy based reductions, extreme care was taken to effectively remove oxygen from (Et 2-dcbpy)2RuX2 solutions for studies on their reduction processes, since the reduction in these complexes are also Et2-dcbpy based.
1.0 0.5

current [A]

0.0 -0.5 -1.0 -1.5 -2.0 1 0

a)

current [A]

-1 -2 -3 -4 -2.5 -2.0 -1.5 -1.0


+

b)
O2 + e O2

-0.5

potential [V] vs. Fc/Fc

Figure 3.17. Cyclic voltammogram for the reduction of 1.0mM Et2-dcbpy in DMF (0.1M Bu4NPF6). a) 2mm glassy carbon electrode, =100mVs1; b) same as a) but with residual oxygen.

3.5.2. L2Ru(CN)2 (a) Voltammetry in DMF Four very well defined processes are accessible in the voltammetric reduction of L2Ru(CN)2 at glassy carbon electrodes in DMF. Assuming the reductions are Et2-dcbpy ligand based, then they may be formulated as follows: (Et2-dcbpy)2Ru(CN)2 + e [(Et2-dcbpy)(Et2-dcbpy)Ru(CN)2] + e [(Et2-dcbpy)2Ru(CN)2]2 (3.16) (3.17) (3.18) (3.19)

[(Et2-dcbpy)(Et2-dcbpy)Ru(CN)2] [(Et2-dcbpy)2Ru(CN)2]2 + e

[(Et2-dcbpy)(Et2-dcbpy2)Ru(CN)2]3 + e [(Et2-dcbpy2)2Ru(CN)2]4

[(Et2-dcbpy)(Et2-dcbpy2)Ru(CN)2]3

Figure 3.18a and b show cyclic voltammograms in DMF (T=22C) at a 1mm diameter glassy carbon disk electrode at scan rates of =100mVs1 and =500mVs1 respectively. The first three processes (Eqns. 3.16-3.19) represent almost ideal models of chemically and electrochemically

Chapter 3: Studies on Model Esters Compounds Reduction

page: 89

reversible processes, whilst the reversibility of the fourth process (Eqn. 3.19) depends strongly on the scan rate, purity of the sample as well as on the water content of the solvent and temperature. Under optimum conditions, the fourth process was fully reversible at scan rates >500mVs1 at 22C. A voltammogram obtained under near steady-state conditions at the same electrode material but with a microdisk electrode configuration is shown in Figure 1c and exhibits four reversible oneelectron reduction steps. Cyclic voltammetric data obtained for the four processes under a range of conditions are summarized in Table 3.5.
2A

a)

4A

b)

1nA

c)

-3.0

-2.5

-2.0

-1.5

potential [V] vs. Fc/Fc

-1.0

-0.5

Figure 3.18. Voltammetric reduction of L2Ru(CN)2 in DMF (0.1M Bu4NPF6) at 22C a) cyclic voltammetry, scan rate 100mVs1, glassy carbon electrode d=1mm, c=1.1mM b) cyclic voltammetry, scan rate 500mVs1, glassy carbon electrode d=1mm, c=1.1mM c) near steady-state conditions at a glassy carbon micro disk electrode d=11m, scan rate 10mVs1, c=1.7mM.

Plots of the reduction peak height, ipred, versus 1/2 from cyclic voltammograms for the first three processes are linear over the scan rate of 10mV to 2000mVs1 as expected if the processes are diffusion controlled. Mass transport control was also demonstrated via the linear dependency of the limiting current, iL, on the square root of the angular velocity, 1/2 in rotating disk electrode experiments for the first three processes and for low rotation rates (1500rpm) for the fourth process. The Er1/2-value as well as other voltammetric data are given in Table 3.5. As required for reversible processes, the values for the initial three processes are almost independent of the voltammetric technique. Minor deviations relative to the theory for a reversible process are observed for the fourth process. The reversible half-wave potentials obtained for all four processes are summarized in Table 3.8. Confirmation that the processes arise from reversible one-electron processes was obtained from "log plots" obtained from rotating disk and microdisk electrode voltammetry the slopes of which had values close to 58mV at 22C (Table 3.5). The almost exact agreement with theory for these

Chapter 3: Studies on Model Esters Compounds Reduction

page: 90

reversible processes under the near steady-state conditions of a microdisk electrode, compared to small departures from ideality with other techniques, implies that a small contribution from uncompensated resistance is present when the other techniques are used. [L2Ru(CN)2]0/
Cyclic Voltammetry / mVs1 10 26 50 100 200 500 1000 2000 E / mV 62 60 71 74 88 108 132 154 Er1/2 / mV 1570 1571 1574 1572 1574 1571 1571 1574 |ipred/ipox| 1.06 1.03 1.02 1.04 0.98 1.01 1.06 1.00 Rotating Disk Electrode f / min1 500 1000 1500 2000 2500 3000 slope / mV Er1/2 / mV

74 1573 77 1573 80 1573 84 1574 83 1579 86 1579 Microdisk Electrode r / m slope / mV Er1/2 / mV 5.6 64 1569 62 1814 68 1815 66 1820 65 1827 66 1833 60 1836 Microdisk Electrode r / m slope / mV Er1/2 / mV 5.6 58 1797 71 2353 64 2353 65 2363 64 2377 58 2382 57 2394 Microdisk Electrode r / m slope / mV Er1/2 / mV 5.6 58 2326 67 2582 67 2608 66 2633 78 2645 Microdisk Electrode r / m slope / mV Er1/2 / mV 5.6 69 2555 500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000 500 1000 1500 2000 2500 3000

[L2Ru(CN)2]/2
10 26 50 100 200 500 1000 2000 58 59 71 69 82 106 128 170 1792 1793 1794 1794 1794 1792 1795 1799 1.07 1.00 0.96 0.93 0.97 0.91 0.93 1.01

[L2Ru(CN)2]2/3
10 26 50 100 200 500 1000 2000 70 84 90 90 105 106 128 178 2320 2321 2324 2322 2321 2320 2321 2330 1.08 1.07 1.04 1.07 1.04 1.00 0.98 1.03

[L2Ru(CN)2]3/4
10 26 50 100 200 500 1000 2000 90 74 70 82 106 120 144 196 2520 2528 2530 2532 2538 2547 2549 2549

Table 3.5.Voltammetric data (T=22C) for the four L2Ru(CN)2 reduction processes in DMF (0.1M Bu4NPF6) at a 2mm diameter glassy carbon electrode. E1/2-values have an uncertainty of 3mV, ipox/ipred-values of 0.04 and slopes of 3mV.

Data obtained from rotating disk electrode measurements and use of the Levich equation (Eqn. 2.31 on page 51) enabled the diffusion coefficient to be calculated with a value of D=3.0(0.4)106cm2s1. From data obtained from microelectrode voltammograms and use of Equation 2.28 (page 48) an average diffusion coefficient of D=2.9(0.2)106cm2s1 can be

Chapter 3: Studies on Model Esters Compounds Reduction

page: 91

calculated. These diffusion coefficients are in excellent agreement with the value of D=3.1(0.2)106cm2s1 measured by using the oxidation process (see section 3.4.1a). (b) Temperature Dependence An interesting thermodynamic effect was observed when voltammetric data obtained at 22C were compared to data obtained at 58C (close to the freezing point of DMF which is 61C). At 22C the reversible Er1/2-values for the first pair of processes are separated by
0 / , / 2 ) 2 / 3 , 3 / 4 ) E (12 = 226mV and the second pair of processes by E (12 = 224mV. However,

the

separations

were

significantly

smaller

at

58C

0 / , / 2 ) ( E (12 = 191mV

and

2 / 3 , 3 / 4 ) E (12 = 210mV). Peak potentials obtained from differential pulse voltammetry were

used to determine these Er1/2 values over the temperature range from 60C to +80C and a linear
0 / , / 2 ) 2 / 3 , 3 / 4 ) and E (12 (Figure 3.19. and dependence on temperature was found for both E (12

Table 3.6). However, interestingly, since the slopes for both pairs of reduction processes are different, a cross over occurs at T=1515C. The data measured for the second set of reduction processes contains an error, since the fourth process is not fully reversible in the higher temperature range. However, this error must be small since under conditions where this process is fully reversible (i.e. T<0C) the same linear relationship is found as is also the case when rotating disk electrode voltammetry and cyclic voltammetry at moderately fast scan rates (=1000mVs1) are used to calculate Er1/2-values.
-260 -250

E 1/2

(0/- , -/2-)

potential difference [mV]

-240 -230 -220 -210 -200 -190 -180 200

E 1/2

(2-/3- , 3-/4-)

220

240

260

280

300

320

340

360

temperature [K]

Figure 3.19. Temperature dependence of the E1/2 separation for the first and second pairs of reduction processes for L2Ru(CN)2. First pair: ....
0 / , / 2 ) ( 2 / 3 , 3 / 4 ) E (12 , second pair: .... E 1 . 2

The observation that the E1/2-value for the first set of reduction processes is smaller than the second set of processes, as observed in this study on L2Ru(CN)2 at temperatures <15C, has been commonly explained in terms of overall charge of the complex; the higher the overall charge of the complex, the higher the energy needed to add an additional charge. Although no discussion concerning the opposite potential separation found at higher temperatures in this work appears to be available in the literature, there are related systems which may be examined in this context. For

Chapter 3: Studies on Model Esters Compounds Reduction

page: 92

[Ru(bpy)3]2+ and [Ru(Et2-dcbpy)3]2+, voltammetric data for the reduction processes in DMF at +20C [142] and 54C [145] have been reported. Unfortunately, the accuracy of the reported potential values (5mV) only allows qualitative comparison to be made for both compounds. In the
2+ / + , 0 / ) case of [Ru(bpy)3]2+, the first set of reduction processes yield E (12 = 0.43V at 20C and 2+ / + , 0 / ) 0.38V at 54C. For [Ru(Et2-dcbpy)3]2+ E (12 = 0.30V at 20C and 0.28V at 54C

whereas E (12 / 2 , 3 / 4 ) = 0.52V at 20C and 0.43V at 54C. Apparently the latter compound exhibits a stronger temperature dependence for the second set of reduction processes than L2Ru(CN)2, which in contrast shows a stronger temperature dependence for the first set of reductions. complex L2Ru(CN)2 process
0 / , / 2 ) E(12 2 / 3 , 3 / 4 ) E(12

intercept [mV] 92 169 98

slope [mVK1] 0.46 0.19 0.42

r [%] 99.3 95.5 99.3

L2Ru(NCS)2

E 12

( 0 / , / 2 )

Table 3.6. E1/2 data obtained by least squares linear regression of plots of potential separations of indicated reduction processes versus temperature. r = correlation coefficient.

The linear relationship between temperature and potential (energy) for E1/2-values may be explained in terms of entropy. By assuming that E1/2=Er1/2=E0', the relationship between the measured E1/2 and the Gibb's free energy, and hence entropy, is given by

E1 / 2 =

G H TS = nF nF
0

(3.20)

where S=S-S0 (S0 and S being the entropies of [L2Ru(CN)2] and [L2Ru(CN)2] , respectively)
0 / , / 2 ) can be formulated as and n=1 for the [L2Ru(CN)2]0/ process. Thus, E (12

( 0 / , / 2 ) E1 = /2

H T( S 2 + S 0 2S ) F

(3.21)

and H is the difference between the enthalpy changes in both processes. Assuming a localized electron model, in the one-electron reduced species, [L2RuX2], the electron can "hop" between the ligands. This hopping can be considered to provide an additional degree of freedom which can be expressed as an additional entropy term. The two electron reduced form, [L2RuX2]2, which has both Et2-dcbpy ligands filled up with an electron, and L2RuX2 lack this ability to permit electron hopping, and hence lack the additional entropy term. The entropy of [L2RuX2] may be written as
S = S system + S electron hopping

(3.22)

whereas the entropy for L2RuX2 and [L2RuX2]2 is only

Chapter 3: Studies on Model Esters Compounds Reduction

page: 93

S 0 = S 2 = S system

(3.23)

Thus, Equation 3.21 reduces to


H T 2 S electron hopping F

( 0 / , / 2 ) E1 = /2

(3.24)

0 / , / 2 ) so that E (12 is governed by the magnitude of S . Assuming H is independent of electron hopping

temperature for all the processes yields a linear relationship of E1/2 with T with a slope of
2 / 3 , 3 / 4 ) /F. Analogous arguments apply for the second E (12 -value, which is 2 S electron hopping
governed by the magnitude of S 3 . The different dependencies of E1/2 on T imply that electron hopping and S 3 have different values. This localized electron model leads to the prediction S electron hopping electron hopping

that, since [L2RuX2] and [L2RuX2]3 have an additional electron hopping term, they are in a preferred lower energetic state. That is, the reduction of L2RuX2 to the one-electron reduced form and the reduction of [L2RuX2]2 to [L2RuX2]3 therefore occur more readily and hence at a more positive potential than would be the case without the additional entropy term. In case of [L2RuX2] the additional stabilization accounts to 136mV at room temperature and S = 22JK1mol1. electron hopping (c) Bulk Electrolysis at 22C and Examination of Reduction Products by ES-MS Although the first two reduction processes were found to be chemically and electrochemically fully reversible at 22C under voltammetric time scales, reduced temperatures were necessary to quantitatively generate [L2Ru(CN)2] and [L2Ru(CN)2]2 (Equations 3.16 and 3.17) under conditions of controlled potential bulk electrolysis (vide infra). The technique of electrospray mass spectrometry (ES-MS) was employed to identify decomposition pathways that occur on longer time scales. In one bulk electrolysis experiment, a 0.6mM solution of L2Ru(CN)2 in DMF (0.01M Bu4NPF6) was exhaustively electrolyzed at 22C. The controlled potential was set at a value slightly more negative than the second reduction process (2.05V vs. Fc/Fc+). The electrospray mass spectrum of the nominally two-electron bulk electrolyzed solution of L2Ru(CN)2 (Figure 3.20) in the negative ion mass spectrum detection mode over the range of 1501200m/z revealed only two major ruthenium products. Simulation (Figure 3.20b) enabled the products to be identified as the fully deesterified and deprotonated acid [((COO)2-bpy)2Ru(CN)2]4, detected as the singly protonated form [((COO)2-bpy)2Ru(CN)2]H3, which has an overall 3 charge at 213m/z. The second signal at 204m/z, also having a 3 charge, can be identified as the fully de-esterified and deprotonated acid [((COO)2-bpy)2Ru(CN)]3 which also has lost one cyanide ligand. If the bulk electrolysis potential is set after the first process at Eappl=1.6V, only mixtures of partially de-esterified starting compound are observed in mass spectra. That is, evidence for reductively eliminated cyanide could

Chapter 3: Studies on Model Esters Compounds Reduction

page: 94

only be found in solutions exhaustively electrolyzed at potentials more negative than the second process. No signals due to ruthenium compounds were found in the positive ion detection mode.
204.3 213.0

b)

204.0

213.1

a)

190

195

200

205

m/z -

210

215

220

Figure 3.20. Negative ion mode electro spray mass spectrum of a) exhaustive bulk electrolyzed solution of 1.0mM L2Ru(CN)2 at Eappl=2.0V in DMF (0.01mM Bu4NPF6), b) simulation of mass spectra for [((OOC)2bpy)2Ru(CN)]3 m/z=204.3 and [((OOC)2bpy)2Ru(CN)2]H3 m/z=213.0.

The ES-MS experiments suggest that after the reduction one-electron bulk electrolysis of (Et2-dcbpy)2Ru(CN)2 to [(Et2-dcbpy)2Ru(CN)2] at 22C in DMF, loss of the ester groups occur, whilst after further exhaustive reduction of the solution reductive elimination of one cyanide ligand occurs. Both classes of reaction are documented in literature. For example, aromatic esters of pyridine derivatives have been shown [177-180] to produce the deprotonated acid analogues when electrochemically reduced in organic solvents and with related (bpy)2RuX2 systems [181] loss of X has been observed when the complex was reduced. Further proof for these reactions pathway will be presented in this chapter. (d) ESR Measurements on L2Ru(CN)2 Reduced at 55C ESR data on the reduced species of compounds of the type [M(bpy)3]2+ (M = Fe, Ru, Os) [153, 182, 183] and related compounds [152, 155, 156] provide evidence for electrons being localized in a single bpy ligand after a reduction process. Thus, ESR signals on the one-electron reduced compounds usually have g-values around the value of the free electron and do not show any fine structure. Additionally, multiple electron reduced species exhibit S= behavior and temperature dependent line broadening which has been interpreted as resulting from an interligand electronhopping. These observations are most easily rationalized by assuming the presence of single chelate ligand (spatially isolated) redox orbitals, so that multi-electron reduced species are multi radicals with little interaction occurring between the electrons. The one and two-electron reduced complexes were prepared ex situ by bulk electrolysis at 55C in DMF. The reduction potential used to obtain the one-electron reduced form had to be carefully chosen to obtain a pure product, because the separation between the first two reduction

Chapter 3: Studies on Model Esters Compounds Reduction

page: 95

processes is only 190mV. Thus, the potential was set only ca. 80mV more negative than the reversible potential for the first reduction process, so that <0.5% of the undesired two electron product is formed, as calculated via use of the Nernst equation. Additionally, considerable care had to be taken to avoid reaction of [(Et2-dcbpy)2Ru(CN)2] with oxygen, especially during the transfer of the electrolyzed solution into the ESR tubes. After transfer of solutions, the ESR tubes were sealed and the solution was immediately frozen in liquid nitrogen. The one-electron [(Et2-dcbpy)2Ru(CN)2] reduced species was shown to have some stability at room temperature in DMF, as demonstrated by repeatedly warming to room temperature and then cooling to the glassy state and noting that only a small decay of the ESR signal was observed. In contrast, the twoelectron [(Et2-dcbpy)2Ru(CN)2]2 reduced species was very reactive and was never obtained in pure form when generated at 55C. The number of electrons consumed in the experiments was always larger than two per molecule. Steady-state voltammetric measurements undertaken during the course of the reduction at potentials more negative than the second process indicated that conversion of the desired product back to [(Et2-dcbpy)2Ru(CN)2] occurred during bulk electrolysis experiments, as would be expected if the two electron reduced compound can catalytically reduce residual water or even the electrolyte or solvent to yield a less reduced form [55, 65].

0.315

0.320

0.325

0.330

0.335

0.340

field [T]
Figure 3.21. ESR spectrum of [L2Ru(CN)2] a) in DMF glass at 77K, b) spectrum simulated with parameter given in Table 3.7.

The ESR spectrum of the pure singly reduced [L2Ru(CN)2] in DMF glass at 77K is shown in Figure 3.21. and may be seen to exhibit a single signal at g =2.000 with a broad wing at the higher

field side which is due to the superimposed parallel component g =1.980. The ESR spectrum is similar to that reported for [Ru(bpy)3]+ [153] which was interpreted by Morris and co-workers [155] to reflect a lowering of symmetry resulting from ligand localization. Thus, if the electron resides on a single ligand, the symmetry of [Ru(bpy)3]+ degrades from D3 to C2v. In the case of the

Chapter 3: Studies on Model Esters Compounds Reduction

page: 96

cis-[L2Ru(CN)2]0/ system, different symmetry considerations apply. Cis-L2Ru(CN)2 has C2v point group symmetry. Reduction and localization of an electron into one of the ligands should reduce the C2v symmetry to C1 and therefore yield a rhombic ESR signal, which clearly is not observed. Therefore in interpreting the ESR spectrum, it can be hypothesized that the ESR spectrum reflects only the ligand symmetry (C2v) and interpret the result as implying the electron resides within the *-orbitals of one single bipyridine moiety and that no significant metal-ligand orbital mixing takes place. As noted above, only solutions of partly two-electron reduced L2Ru(CN)2 (3050% [L2Ru(CN)2]2 content) could be prepared by bulk electrolysis. The ESR spectrum obtained from an impure solution was indistinguishable from that of [L2Ru(CN)2] except for the fact that the intensity of the signal was enhanced. Variable temperature ESR experiments on [L2Ru(CN)2] prepared in DMF and acetone showed that the intensity of the signal diminished with increasing temperature and became indistinguishable from the background above ~180K. The temperature dependent line broadening, as expected if inter-ligand electron hopping occurs, was not observed within the temperature range of 3180K. For [M(terpy)2]+ (M=Ru and Fe, terpy = terpyridine) no ESR signal was reported [155] at temperatures above 70C, whilst for [Os(bpy)3]0 the threshold temperature for observation of an ESR signal was at about 40C [182], although line broadening was observed between 60 and 40C. Reducing the temperature to 3K did not lead to any fine structure being resolved for [L2Ru(CN)2]. At g=2.15 an additional broad ESR signal with varying intensity was observed under some conditions for [L2Ru(CN)2]. This signal was enhanced at lower temperatures, in acetone and in samples that were stored in liquid nitrogen for several days. Analogous behavior has been reported for [Fe(terpy)2]n (n=+1, 0, 1). Crystalline solid state ESR spectra of [Ru(bpy)3]0 and [Fe(Me2phen)3]0 have been reported to be very different from frozen solution spectra [182]. Whereas in frozen solution the signals occurred at ca. g=2.00, as expected for a ligand based reduction, the solid state ESR signals were very broad and observed at g=2.152.3. These data have lead to the conclusion that upon crystallization and loss of solvation, the electronic structure of the compounds drastically changes and that the * and * orbitals swap in relative energy. Thus, the metal based *-orbital becomes the LUMO and the ligand based *orbital increases in energy and becomes the second lowest unoccupied molecular orbital. This has been interpreted to result in a d8 metal-localized configuration in the case of [Ru(bpy)3]0. (e) Spectroelectrochemical Studies on Reduced L2Ru(CN)2 Electronic spectra of the one and two electron reduced forms of L2Ru(CN)2 can be obtained at 58C in DMF (Figure 3.22) via in situ OTTLE spectroelectrochemical experiments. These low temperature, small volume (0.51.5ml) and relatively short time domain OTTLE experiments enabled the two electron reduced form of L2Ru(CN)2 to be generated almost quantitatively. Reduction to [L2Ru(CN)2] and re-oxidation back to the starting material occurred without any

Chapter 3: Studies on Model Esters Compounds Reduction

page: 97

detectable loss of material in the sense that the initial and final electronic spectra were identical. Analogous experiments associated with generation of [L2Ru(CN)2]2 and re-oxidation back to the starting material showed losses of material of less than 5%. Due to the considerable richness of the electronic spectra of both reduced and oxidized forms of the compounds at least six isosbestic points emerge during the course of reduction of L2Ru(CN)2 to [L2Ru(CN)2] and seven for the additional reduction to [L2Ru(CN)2]2.
0.20

5 4

0.15 0.10 0.05 0.00

a)
4000 4500 5000 5500

molar absorbance [10 M cm ]

3 2 1 0
0.20

-1

-1

4 3 2 1 0

0.15

b)

0.10

0.05 4000 4500 5000

10000

20000
-1

30000

40000

energy [cm ]
Figure 3.22. Electronic spectra obtained in an OTTLE cell during the course of reduction of 0.6mM L2Ru(CN)2 in DMF (0.2M Bu4NPF6) at 58C. a) reduction of L2Ru(CN)2 to [L2Ru(CN)2], b) reduction to [L2Ru(CN)2]2. Inserts show evidence for a weak NIR band for the one-electron reduced form.

Electronic spectral data are summarized in Table 3.7. The spectra of the reduced compounds reflect the occurrence of Et2-dcbpy ligand based reductions. In the case of [Ir(bpy)3]3+ and [Ru(bpy)3]2+, it has been elegantly shown [122-124], that the electronic spectra of the triply reduced forms are almost identical with the spectrum of the one-electron reduced free ligand. In this sense, the reduction from [L2Ru(CN)2] to [L2Ru(CN)2]2 leads to the complete disappearance of the characteristic Et2-dcbpy ligand * band at 31500cm1. Additionally, both MLCT bands present at 25000cm1 and 18800cm1 disappear, whilst new intraligand bands grow in in the same region. The spectrum of [L2Ru(CN)2] is complicated in the visible region by superimposition of the RuII(Et2-dcbpy)0 MLCT and the characteristic Et2-dcbpy1 intraligand transitions. The spectrum of [L2Ru(CN)2]2 resembles that of reduced bipyridine [124, 184], the major difference being the position of an intense NIR band at 5700cm1 occurring at significantly lower energies than in bpy and [Ru(bpy)3] (~12500cm1). Elliott [143] reported the disappearance of this NIR band in the

Chapter 3: Studies on Model Esters Compounds Reduction

page: 98

[Ru(Et2-dcbpy)3]2+ system when more than one electron was added into one ligand and assigned this transition to an interligand intervalence charge transfer (IVCT). However, Heath and coworkers [151] proved that the IVCT process could be detected at lower energies and with significantly lower absorbance and hence assigned the NIR band to an intra-ligand transfer. The large shift of this NIR band relative to bpy, reflects the substantially different energies of the molecular orbitals of the ligands and especially the differences in relative orbital energies as determined by MNDO calculations [144]. The calculated energy difference between the 1* and 2* orbital is approximately halved when comparing Et2-dcbpy and bpy as is also the case with the energy of this experimentally observed band. Consequently, this transition is tentatively assigned as arising from an Et2-dcbpy based 1*2* transition. This is consistent with Elliott's observation [143] that this band disappears when a second electron is placed onto the ligand, which will significantly lower the 1* energy and hence shift the band out of that region. On the basis of the above argument, the shoulder at ca. 7400cm1 will then be due to a 1*3* transition, as the 2* and 3* orbitals are almost identical in energy.
a) complex [(L)LRu(CN)2] [(L)2Ru(CN)2]2 b) absorbance energy [103cm1] (molar extinction [103mol1lcm1]) 31.5 (40.0) 27.7 (20.0) 22.3 (17.7) 18.2 (15.5) 17.0 (10.7) 6.0 (5.3) 3.5 (0.8) 27.5 (42.2) 20.5 (21.5) 5.7 (11.4) L L* MLCT MdL2* MLCT MdL1*

energy [103cm1] (molar extinction [103mol1lcm1]) L2RuCl2 L2RuI2 L2Ru(NCS)2 L2Ru(CN)2 c) [(L)LRu(CN)2] [(L)LRu(NCS)2] 31.1 (37.4) 31.2 (39.0) 31.3 (45.7) 31.5 (58.4) g (H [mT])

23.1 (13.1) 23.5 (13.6) 24.5 (14.3) 25.0 (18.0) g (H [mT]) 1.9800 (3.6)

16.9 (14.2) 17.2 (12.9) 18.1 (14.5) 18.8 (22.1)

1.9996 (1.3)

Table 3.7. Spectroscopic data for L2RuX2 and reduced forms of L2RuX2 in DMF. a)Electronic spectra data obtained for reduced forms of L2Ru(CN)2 at 58C. b) Electronic spectra data obtained L2RuX2 complexes at 22C. c) Parameters used for simulation of the ESR spectrum shown in Figure 3.21 on page 95.

The band for [L2Ru(CN)2] in the near infrared region at ~3500cm1, having a very low absorbance (see Figure 3.22 inserts), is also attributable to an inter-ligand IVCT mentioned above. Its presence provides further evidence of inter-ligand electron hopping. Unfortunately, solvent overtones and strong absorbance from residual water mask this process. However, data points obtained at lower energies enabled the position and peak intensity to be estimated by interpolation. This transition is caused by electron hopping [151, 156] from one Et2-dcbpy ligand to the other as described in Equation 3.25 As expected, this band disappears when [L2Ru(CN)2] is reduced to [L2Ru(CN)2]2 and this feature leads to an additional isosbestic point, as shown in the relevant insert in Figure 3.22. The position of this band is similar to that of reduced [Ru(bpy)3]2+ (4500cm1) [151], although the absorbance is about three times higher (see Table 3.7).

Chapter 3: Studies on Model Esters Compounds Reduction

page: 99

[(L)(L)RuX2]

[(L)(L)RuX2]

(3.25)

In recent papers [86, 93] the authors of pump and probe laser experiments reported an estimate of the electron injection rate of excited (H2-dcbpy)2Ru(NCS)2 into TiO2. The probing was undertaken at wavelengths of 750nm and 1100nm which were thought to be typical of the oxidized dye [(H2-dcbpy)2Ru(NCS)2]+, and hence indicative of the rate of arrival of the electron into the semiconductor. The calculated rates implied that electron injection is faster than the detector response time. Interestingly, spectra for the reduced forms of L2Ru(CN)2 (Figure 3.22), which are expected to be very similar to that of reduced (H2-dcbpy)2Ru(NCS)2, show very strong absorbances at the same wavelength regions where the probing was undertaken in these experiments. These bands, assigned in this work to intra-ligand charge transfers in the reduced form of the ligand, therefore are also expected to be present in the excited state, which formally has the configuration [(L)(L)Ru+X2]. Thus, it is likely that the formation of the excited state rather than the electron injection time has been measured in the work of reference [86] and [93]. Authors of a recent paper [77] recognized this problem and have used mid-IR frequencies (47m) to directly probe conduction band electrons. 3.5.3. L2Ru(NCS)2 (a) Voltammetry in DMF A cyclic voltammogram for reduction of L2Ru(NCS)2 is shown in Figure 3.23a and can be seen to exhibit many of the characteristics found for the cyanide analogue. The main difference is the decreased stability of the reduced forms. For this compound, only the first two reduction processes are chemically and electrochemically reversible at 22C. Under conditions of cyclic voltammetry, ipred for the first two processes scaled linearly with 1/2 and a Levich plot (rotating disk electrode) was linear and passed through the origin for both processes, establishing that both processes are mass transport controlled. Er1/2-values for both processes obtained from these two techniques were calculated to be 14803mV and 16974mV vs. Fc/Fc+ and coincided with values obtained from microdisk electrode experiments under near steady-state conditions. "Log-plots" from microdisk and rotating disk electrode voltammograms gave slopes of 633mV, which are close to the theoretical value expected for a reversible one-electron process. Slopes of rotating disk electrode Levich plots for the first and second process were identical within experimental error and lead to a diffusion coefficient for L2Ru(NCS)2 in DMF of D=3.20.4106cm2s1, which is very similar to the value obtained for L2Ru(CN)2. Cyclic voltammetric scan rates >200mVs1 were required to make the third reduction process of L2Ru(NCS)2 fully reversible. Under near steady-state conditions of a microdisk electrode, and also in rotating disk electrode experiments, the third process appears to be reversible in the sense that slopes calculated from "log-plots" were 615mV, close to the value of 58mV theoretically expected for a reversible one-electron process. Er1/2-values calculated from these "log-plots" for the third process were in agreement with those obtained from

Chapter 3: Studies on Model Esters Compounds Reduction

page: 100

cyclic voltammograms (>200mVs1). Thus the reversible half wave potential for the [L2Ru(NCS)2]2/3 process is established to be 223020mV.
a) b)

5A 1A

c)

d)

5A

1A

e)

f)

5A

1A

-2.5

-2.0

-1.5

-1.0

-0.5
+

-2.5

-2.0

-1.5

-1.0

-0.5
+

potential [V] vs. Fc/Fc

potential [V] vs. Fc/Fc

Figure 3.23. Cyclic voltammograms for reduction of L2RuX2 as a function of switching potential and temperature in DMF (0.1M Bu4NPF6) at a glassy carbon working electrode (d=1.5mm, =100mVs1). 1.0mM L2Ru(NCS)2, a) T=22C, b) T=58C; 1.0mM L2RuCl2, c) T=22C, d) T=58C; 1.1mM L2RuI2, e) T=22C, f) T=58C.

At temperatures below 20C, the third and fourth processes become fully reversible (Figure 3.23b) under all voltammetric conditions. The reversible half wave potentials at 58C for the third and fourth process were measured to be 21606mV and 24286mV vs. Fc+/Fc in DMF. As was the case with L2Ru(CN)2, the reversible half wave potentials of the reduction processes were investigated as a function of temperature. Due to the low reversibility of the third and fourth
0 / , / 2 ) processes at elevated temperatures, only the first two processes were investigated. E (12

was again found to vary in a linear manner with temperature over the temperature range of 60C to +80C, the slope being similar to that obtained with L2Ru(CN)2 (Table 3.6). It is postulated that the reduced back-bonding ability of the NCS ligand compared to the very good back-bonding ligand CN, contributes to the decreased stability of reduced forms of L2Ru(NCS)2. That is, the weaker back-bonding character of NCS results in an enhanced reductively induced elimination of the NCS ligand.

Chapter 3: Studies on Model Esters Compounds Reduction

page: 101

Reversible Potentialsa) Er1/2 / mV complexes L2Ru(CN)2 L2Ru(NCS)2 L2RuI2 L2RuCl2 [L2RuX2]0/ 15703mV 14803mV 155510mVb) 15884mV [L2RuX2]/2 17965mV 16974mV 179010mVb) [L2RuX2]2/3 23267mV 223020mV [L2RuX2]3/4 255020mV 24286mVb)

Table 3.8. Reversible half wave potentials for the reduction of L2RuX2 in DMF (0.1M Bu4NPF6). a) mV versus Fc/Fc+. b) T=58C, other data reported at T=22C.

(b) Bulk Electrolysis and ESR Spectra To generate the singly reduced [L2Ru(NCS)2] anion, controlled potential bulk electrolysis of a 1.0mM solution of L2Ru(NCS)2 in DMF (0.1M Bu4NPF6) was carried out at Eappl=1550mV and T=58C. The lower stability of reduced forms of L2Ru(NCS)2 compared to L2Ru(CN)2 did not allow the electrolysis to go to completion without product decomposition. Therefore samples for ESR measurements were collected after 0.5 electrons per molecule had been transferred to ensure that no appreciable decomposition of [L2Ru(NCS)2] had taken place. The ESR signal of [L2Ru(NCS)2] in DMF glass at 77K was indistinguishable from that of [L2Ru(CN)2]. 3.5.4. L2RuI2 and L2RuCl2 (a) Voltammetry and Reductively Induced Halide Ligand Elimination At 22C and under conditions of cyclic voltammetry at a scan rate of 100mVs1 (Figure 3.23c), L2RuI2 is irreversibly rather than reversibly reduced (peak potential: 1630mV). Under these conditions, the first two reduction processes expected on the basis of data obtained with the cyanide and thiocyanate derivatives have merged into a single irreversible process. Lowering the temperature to 58C increases the reversibility of the first two reduction processes and simplifies the voltammetry (Figure 3.23f). At this low temperature, the first reduction process is chemically and electrochemically fully reversible under the near steady-state conditions of the microdisk or rotating disk electrode and under conditions of cyclic voltammetry at scan rates 100mVs1. As expected, the measured Er1/2-value (see Table 3.8) was technique independent. Slopes of "log-plots" obtained from steady-state voltammograms were 424mV as theoretically expected (43mV at 58C) for a one-electron process. Only a peak potential of Epred=1720mV (=100mVs1, cyclic voltammetry) can be quoted for this second irreversible process under the low temperature voltammetric conditions used. Slow scan rate cyclic voltammograms of L2RuCl2 at 22C reveal the presence of one partly reversible reduction process (Figure 3.23c), followed by a series of irreversible processes. At scan rates 200mVs1 the first reduction process was close to reversible and the E1/2-value was constant (3mV) over the scan rate range of 25 to 2000mVs1, which implies that this value is close to the reversible Er1/2-value. Slopes calculated from "log-plots" obtained from near steady-state microdisk electrode and rotating disk electrode voltammograms were 623mV and Er1/2-values obtained from the latter techniques coincided with the value obtained from cyclic voltammograms (Table 3.8).

Chapter 3: Studies on Model Esters Compounds Reduction

page: 102

Again at reduced temperature, the chemical reversibility of this system increased. At 58C the first reduction process is fully reversible even at slow scan rates (Figure 3.23e) whilst the second reduction process becomes partly reversible. The low temperature reversible Er1/2-value for this second process is listed in Table 3.8. Cyclic voltammograms for L2RuI2 in DMF in the positive potential region before and after the potential is scanned over the first reduction wave are shown in Figure 3.24b. On the second cycle, two new oxidative processes appear if the reductive process is included in the initial scan. The first of these new processes precedes the metal centered oxidation of L2RuI2 (Epox=+0.05V) and is due to the oxidation of liberated I, whilst the second of the new processes, which follows the metal centered oxidation of L2RuI2, is attributable to the reversible one-electron oxidation of [L2RuI(DMF)]+ (see Section 3.4.2 on page 74). The major difference between oxidatively (Section 3.4.2) and reductively induced ligand elimination is the oxidation state of the eliminated ligand. Thus, chemically irreversible oxidation eliminates the oxidized ligand (I2), whilst irreversible reduction eliminates the reduced form of the ligand (I).
2A

b)

1A

a)
-2.0 -1.5 -1.0 -0.5 0.0 0.5
+

1.0

potential [V] vs. Fc/Fc

Figure 3.24. First (dashed) and second cycles (full line) of cyclic voltammograms for reduction of a) 1.0mM L2RuCl2 in DMF (0.1M Bu4NPF6) at a 1mm diameter platinum electrode, =100mVs1 b) 1.0mM L2RuI2 in DMF (0.1M Bu4NPF6) at a 1mm diameter glassy carbon working electrode, =500mVs1.

The potential of the [L2RuI(DMF)]+/2+ couple at E1/2=+0.45V vs. Fc/Fc+ is in very good agreement with the value E1/2=+0.46V vs. Fc/Fc+ obtained from electrochemically generated solutions of [L2RuI(DMF)]+ (compare Table 3.3 on page 71). Slower scan rates reveal the appearance of the bis-solvent [L2Ru(DMF)2]2+ complex in the positive potential range. Reversible oxidation for the [L2Ru(DMF)2]3+/2+ couple occurs at E1/2=+0.62V vs. Fc/Fc+ which agrees with E1/2=+0.62V vs. Fc/Fc+ for this process when using electrochemically generated solutions of [L2Ru(DMF)2]2+ (compare Table 3.3 on page 71).

Chapter 3: Studies on Model Esters Compounds Reduction

page: 103

The analogous behavior observed for L2RuCl2 is shown in Figure 3.24b. In this compound chloride is lost after the potential is scanned over the first two reduction waves and the formation of the mixed solvent complex is seen when the potential is scanned back to the potential region. Slower scan rates reveal the appearance of the bis-solvent [L2Ru(Solvent)2]2+ complex and the reaction details are summarized in Equations 3.26-3.28 L2RuIIX2 + e [(L)(L)RuIIX2] [(L)(L)RuIIX2] (L )(L)RuIIX(Solvent) + X [(L )(L)RuII(Solvent)2]+ + X (3.26) (3.27) (3.28)

(L )(L)RuIIX(Solvent) + Solvent

In contrast, reduced forms of L2Ru(CN)2 and L2Ru(NCS)2, which are much more stable, did not give rise to formation of these solvent complexes on the time scale of cyclic voltammetry. Even scanning up to the cathodic solvent limit did not generate any additional process(es) corresponding to oxidation of the pseudo halide/solvent mixed complex at positive potentials on the second cycle. This difference in stability of the reduced complexes can be explained by the capability of the isothiocyanate and cyanide ligand to establish strong -back-bonding, which may compensate for the loss in -bonding that occurs on reduction. That is the ability to maintain a strong -back-bond determines the stability of L2RuX2 (L = bipyridine, X = halide or pseudo halide) bipyridine based reduction products. The origin of the decreased stability of reduced forms of L2RuX2 where X=Cl, I compared to X=CN, NCS may be understood in terms of molecular orbital considerations. Cyanide and isothiocyanate are both good -donors as well as -acceptors. In contrast, chloride and iodide are not capable of maintaining a strong back-bond. After the Et2-dcbpy * orbital is occupied upon reduction, the bipyridine will lose its ability to adequately stabilize the d metal orbitals and thereby maintain a strong d* back-bond. On the other hand, the -donor strength increases, and indeed, in many cases, the reduced form of aromatic imine ligands have been found [185] to be better ligands than the unreduced ligands. This increased -donor strength lowers the energy of the d orbitals of the ruthenium metal which will then effect the -bond strength of other ligands which can lead to a loss of unreduced ligand, partly because of the trans effect [186]. An alternative explanation is based on the tendency of compounds to reach a state of minimal charge. Accordingly in this context, the complex frees itself from the additional charge imposed by the reduction process, by ejecting an anionic ligand. In (bpy)2RuX2 systems [181], the loss of X also occurred when the complex was reduced. (b) Bulk Reductive Electrolysis of L2RuI2 in DMF and Identification of Products by ES-MS Further evidence for the presence of reductively induced ligand elimination described in Equations 3.26-3.28 was obtained from electrospray mass spectrometric measurements. A 0.7mM solution of L2RuI2 in DMF (0.01M Bu4NPF6) was reductively bulk electrolyzed at a platinum

Chapter 3: Studies on Model Esters Compounds Reduction

page: 104

gauze electrode at a potential of 1.5V. Samples were taken during the course of the electrolysis and examined by electrospray mass spectrometry. A sample taken during an early stage (0.7 electrons per molecule transferred) is shown in Figure 3.25. A series of ruthenium products is detectable in the positive ion mode. The major signal at 424.1m/z is assigned to formation of the doubly charged bis-DMF complex, [(Et2-dcbpy)2Ru(DMF)2]2+. The signal for the mono-DMF complex, [(Et2-dcbpy)2RuI(DMF)]+, is also detectable at 902.2m/z but with less intensity. The signals at 401.7m/z and 861.1m/z can be assigned to complexes formed by replacement of DMF with the methanol solvent used for mass spectrometric experiments. No signals were found in the negative ion detection mode. This result is in agreement with conclusions made on the basis of cyclic voltammetric data. In another experiment, exhaustive electrolysis was carried out at 2.1V. In this case, the major ES-MS peaks were found in the negative ion mode and could be assigned to the fully de-esterified bis-solvent complexes.
424.1

902.2
424.1

902.2

a)
401.7
420

b)
424.1

861.1
902.2

425

430

895

900

905

910

300

350

400

m/z

450 +

500

550

750

800

850

m/z+

900

950

1000

Figure 3.25. Electrospray mass spectra obtained from a partly (0.7 electrons) electrolyzed (Eappl=1500mV) solution of L2RuI2 in DMF (Bu4NPF6). Inserts show experimental (bottom) and simulated (top) spectra of the solvated complexes (a) [L2RuI(DMF)]+ at 902.1m/z+ and (b) [L2Ru(DMF)2]2+ centered at 424.1m/z+.

3.5.5. Electronic Spectra Absorption data obtained from electronic spectra measured in DMF are contained in Table 3.7b. All complexes exhibit very intense bands in the visible and UV region of the electronic spectrum. The most intense transition is found at ca. 31200cm1, with molar absorbances of up to 58400M1cm1 (L2Ru(CN)2). This transition is assigned to a ligand (Et2-dcbpy) based * process, since the free ligand shows a transition at the same energy. Two MLCT bands, Ru(d)Et2-dcbpy(*), of almost identical intensity, shape and absorbance are found in the visible region. The observation of two bands is a result of the energetic closeness of the 1* and 2* Et2dcbpy orbitals. The band at lower energies is a d1* transition; the band at higher energies is the direct d2* transition into the next higher unoccupied orbital [187]. The differences between both bands, ~6300cm1=0.79eV, correlates well with the spacing of the 1* and 2* orbital of 0.6eV obtained from MNDO calculations [144] for the free ligand. Both bands have a tail on the lower energy side, which extends to approximately 12500cm1 (L2RuCl2). The compound therefore absorbs over the whole visible region of the spectrum. This tail has been interpreted [86] as a spin forbidden direct transition into the triplet state. No additional bands are observed down to 3300cm1.

Chapter 3: Studies on Model Esters Compounds Conclusion

page: 105

3.6. Conclusion
From this study it has emerged that L2RuCl2 is a model compound for voltammetric studies of the [L2RuX2]0/+ redox couple. Electrochemically generated [L2RuCl2]+ is stable on the time scale of hours and can therefore be characterized by a range of spectroscopic techniques. ESR spectra obtained from [L2RuIIICl2]+ suggest that the symmetry of the oxidized complex is less then C2v, which can be explained by assuming that distortion of the Et2-dcbpy ligands occurs after oxidation. A more complex cyclic voltammetric response was observed when solid L2RuCl2 was oxidized in water. [L2RuCl2]+ appeared to dissolve into the aqueous solution under most conditions, but this could be prevented when LiClO4 was used as the electrolyte. When more easily oxidizable halide or pseudo-halide ligands are present, the redox chemistry associated with the oxidation process becomes considerably more complicated. In the case of L2RuI2, electrochemical oxidation resulted in an overall internal reaction leading to loss of iodine. When coordinating solvents such as DMF and CH3CN were used, solvent substituted complexes were formed whereas in non-coordinating solvents the five coordinate complex [L2RuI]+ was formed. The formation of these charged complexes was proven by electrospray mass spectrometry and voltammetric techniques. The oxidation of L2Ru(CN)2 is complicated by the formation of mono and poly-nuclear ruthenium compounds. Products formed by reaction with the oxidized ligand, cyanogen (CN)2, or a derivative, are believed to be involved in the complex reaction pathways. Voltammetric oxidation of L2Ru(NCS)2 is accompanied by adsorption. It can be proposed that surface attachment occurs via the sulfur component of thiocyanate groups, since surface processes are absent when other halides or pseudo halides are present. Oxidation of L2Ru(NCS)2 leads to formation of the cyanide analogue L2Ru(CN)2 and appears to proceed without the formation of the mixed ligand complex L2Ru(NCS)(CN). Degradation in photovoltaic cells also may proceed via this route, since the main photoelectrochemical cycle involves the momentary oxidation of the sensitizer [81]. The reductive electrochemistry of L2RuX2 has been investigated in DMF. The presence of low lying 1*-orbitals of the Et2-dcbpy ligand enables up to four reversible Et2-dcbpy ligand based oneelectron reductions to be observed. The reversible potentials occur at considerably less negative potentials than reported for the bpy analogues. The degree of reversibility for the ligand based reduction processes are in almost reverse order to those observed for the metal centered oxidation processes. The stability of the reduced forms is determined by the ability of the halide or pseudohalide ligand to maintain a strong -back bond with the metal center after reduction. Thus I and Cl ligands, which are unable to form strong -back bonds are reductively eliminated and replaced by the solvent (DMF). A competing reaction associated with decomposition of the reduced complexes is de-esterification which gives rise to deprotonated carboxylate groups.

Chapter 3: Studies on Model Esters Compounds Conclusion

page: 106

Spectroscopic data obtained for reduced forms of L2Ru(CN)2, such as the detection of an interligand IVCT band in the NIR region of the electronic spectrum, provide evidence for the presence of spatially localized redox orbitals and electron hopping. The large temperature dependence of the reversible potentials of the reduction processes also appears to be attributable to the same phenomena.

3.7. Appendix: Temperature Dependence of Ligand Based Reductions


3.7.1. Introduction In the previous sections it has been demonstrated, that the electron hopping phenomena of ligand based electrons results in an unusual temperature dependence of the ligand based reduction potentials. The information available in the literature on this phenomenon of similar compounds is restricted [152, 153] to structures of the type [Me(bpy)3]2+ where bpy is a bipyridine derivative and Me = Ru, Os and Fe. That is, the electron hopping was only observed to occur between equivalent ligands. Heath and co-workers [151] were the first to observe the electron hopping induced ligandligand inter-valence charge-transfer of the [Ru(bpy)3]2+ system by near infra-red spectroscopy. They used Hush's theorem [188] to relate the frequency of the transition to the activation barrier for the electron hopping process, which was measured by Motten and co-workers [153] by variable temperature ESR spectroscopy. Hush [188] showed that optical transition occurs at four times the activation energy of the thermal transfer. Eop = 4E*th (3.29)

However, this situation only applies when the initial and the final states are identical, i.e. to a symmetrical homoorbital one-electron transfer process. This situation is illustrated in Figure 3.26a, both initial and final state orbitals are identical, so that the overall energy change, E00, is zero. It is interesting to note that no luminescence phenomena are expected to occur from this transition, since both states are thermoneutral. So far it has been seen as an essential requirement that both states are identical, which is not necessarily true. The case of a charge transfer between two nonequal orbitals is shown in Figure 3.26b. Clearly, the energy for the optical transition, Eop, will increase, on the other hand the oscillator strength (intensity of the transition), which is determined by the quality of the overlap between the wave functions of both orbitals, will be likely to decrease. Based on these considerations, preliminary experimental work has been conducted to show that electron hopping can occur between non-equivalent ligands. Certainly, the above stated thoughts greatly oversimplify this problem and proper mathematical treatment would be required for an accurate description of this phenomenon, which would be beyond the scope of this work. Complexes of the type [RuXYZ]2+ and [RuX2Y]2+ have been investigated, where X,Y and Z are similar, but different bipyridine derivatives.

Chapter 3: Studies on Model Esters Compounds Appendix

page: 107

a)

b)
Eop Eop E*th E*th

E00

Figure 3.26. Potential energyconfigurational coordinate diagrams for an inter-valence charge-transfer. Eop = energy of optical transition, E*th = thermal energy and E00 = overall energy change or zero point optical transition. a) transfer between identical orbitals (Eop=4E*th), b) transfer between non-identical orbitals. Adopted from reference [188].

3.7.2. Experimental: Compounds Samples of Ru(bpy)2Cl2, [Ru(bpy)(4,7-CH3-phen)(5,6-CH3-phen)](PF6)2, [Ru(bpy)2(phen)](PF6)2 and [Ru(bpy)(phen)(4,4'-CH3-bpy)](PF6)2, where phen = phenanthroline, were kindly donated by Dr. Sue Jenkins and Dr. Chris Kepert, Monash University. [Ru(bpy)2(Et2-dcbpy)]2+: 100mg (215mol) Ru(bpy)2Cl2 and 100mg (330mol) Et2-dcbpy were refluxed in ethanol for four hours. The solution was evaporated to dryness and the solid residue chromatographed on neutral Al2O3 with a solvent mixture of ethanol / CH2Cl2 1:5. Unreacted dark purple Ru(bpy)2Cl2 elutes first followed by the orange band of the desired [Ru(bpy)2(Et2-dcbpy)]Cl2, which is collected and evaporated to dryness. The solid is dissolved in ethanol and an excess of NH4PF6 is added to the solution. The solution is stored at 20C for five days in order to allow the PF6 salt to precipitate. 100mg (46%) orange crystals of [Ru(bpy)2(Et2dcbpy)](PF6)2 are removed from the solution by filtration. 3.7.3. Results and Discussion [Ru(bpy)2(Et2-dcbpy)]2+ exhibits one reversible one-electron metal centered oxidation process and four reversible one-electron ligand based reduction processes under conditions of cyclic voltammetry in DMF (0.1M Bu4NPF6). A cyclic voltammogram is shown in Figure 3.27. and the potentials may be found in Table 3.9. Since the reduction processes are ligand localized and Et2dcbpy is about 0.5V easier to reduce than bpy, each of the reduction processes can be unambiguously assigned to a specific ligand. Thus, the first reduction process can be assigned to the Et2dcbpy ligand, followed by two bpy reduction processes and the fourth reduction process adds a second electron into the * orbital of the Et2-dcbpy ligand yielding the formal configuration [Ru(bpy)2(Et2-dcbpy2)]2.

Chapter 3: Studies on Model Esters Compounds Appendix

page: 108

30 20 10

current [A]

0 -10 -20 -30 -40 -2.5 -2.0 -1.5 -1.0 -0.5 0.0
+

0.5

1.0

potential [V] vs. Fc/Fc

Figure 3.27. Cyclic voltammogram of 1.1mM [Ru(bpy)(Et2-dcbpy)2](PF6)2 in DMF (0.1M Bu4NPF6). 2mm glassy carbon working electrode, =200mVs1, T=20C.

redox process [V] vs. Fc/Fc+ compound [Ru(bpy)2(Et2-dcbpy)] [Ru(bpy)(4,7-CH3-phen)(5,6-CH3-phen)]2+ [Ru(bpy)2(phen)]2+ [Ru(bpy)(phen)(4,4'-CH3-bpy)]2+
2+

2+/3+ 0.92 0.75 0.81 0.745

2+/+ 1.40 1.78 1.74 1.775

+/0 1.83 1.965 1.915 1.965

0/ 2.04 2.24 2.18 2.03

/2 2.39

Table 3.9. Reversible half-wave potentials for [RuXYZ]2+ complexes in DMF (0.1M Bu4NPF6) at 20C. Calculated from a series of measurements by cyclic voltammetry, microdisk electrode steady-state voltammetry and differential pulse voltammetry. Working electrode: 3mm diameter glassy carbon disk. Potentials quoted have an error of less than 10mV.

The reduction potentials for the first set of reductions (2+/+, +/0 and 0/) are investigated in DMF as a function of temperature over the temperature range of 60 to +85C. Similar to the L2RuX2 compounds, linear dependencies are found between the potential differences of these three
2+ / + , + / 0 ) + / 0 ,0 / ) processes, E (12 and E (12 , and are shown in Figure 3.28a and the values are given

in Table 3.10.

E (122 + / + , + / 0 )
compound [Ru(bpy)2(Et2-dcbpy)]2+ [Ru(bpy)(4,7-CH3-phen)(5,6CH3-phen)]2+ [Ru(bpy)2(phen)]2+ [Ru(bpy)(phen)(4,4'-CH3-bpy)]2+

E (12+ / 0 , 0 / )

intercept [mV] slope [mVK1] intercept [mV] slope [mVK1] 365 99 106 105 0.23 0.32 0.24 0.29 91 164 242 254 0.39 0.31 0.06 0.10

Table 3.10. E1/2 data obtained by least squares linear regression of plots of potential separations of indicated reduction processes versus temperature.

In the singly reduced complex, the electron is located on the Et2-dcbpy ligand and electron hopping can only occur to one of the two unfavorable bpy ligands, which are energetically higher than Et2-dcbpy (see Figure 5.5 on page 147). In the doubly reduced complex, the second electron is located on one bpy ligand, so that electron hopping can occur between both bpy ligands where the
2+ / + , + / 0 ) orbital overlap can be assumed to be favorable. As expected, the value for E (12 is found to

Chapter 3: Studies on Model Esters Compounds Appendix

page: 109

+ / 0 ,0 / ) be 0.23mVK1 which is smaller than the value found for E (12 = 0.38mVK1, although the

one-electron reduced form has two empty bpy ligands available for the electron hopping process and the two-electron reduced form only has one available.
460 300

[Ru(bpy)2(Et2-dcbpy)](PF6)2
440 280

[Ru(bpy)2(phen)](PF6)2

potential difference [mV]

420

E 1/2

(2+/+ , +/0)

potential difference [mV]

260

E 1/2
240 200

(+/0 , 0/-)

400 220

a)
E1/2
220 240 260 280 300 320

b)
E1/2
(2+/+ , +/0)

200

(+/0 , 0/-)

180

180 340 360

160 220 240 260 280 300 320 340 360

temperature [K]
280

temperature [K]
300

[Ru(bpy)(4,7-Me2phen)(5,6-Me2phen)](PF6)2

[Ru(phen)(bpy)(4,4'-Me-bpy)] (PF6)2

2+

2-

potential difference [mV]

260

E 1/2

(+/0 , 0/-)

potential difference [mV]

280

240

260

E 1/2

(+/0 , 0/-)

220

c)
E 1/2
220 240 260 280 300 320

200

d)
E 1/2
220 240 260 280 300

200

(2+/+ , +/0)

180

(2+/+ , +/0)

180

160 320 340 360

340

360

temperature [K]

temperature [K]

Figure 3.28. Temperature dependence of potential separation for the [RuXYZ]2+ series. Potential separation measured by differential pulse voltammetry, least square linear regression was used for linear fit. For convenience all graphs are scaled exactly the same.

The results obtained from [Ru(bpy)2(phen)]2+ are in agreement with the above argument. In this complex the bpy ligand is easier reducible than the phen ligand, the E1/2 for the [Ru(bpy)3]2+/+ process is 1.34V vs. SCE and for [Ru(phen)3]2+/+ the potential is 1.41V vs. SCE [62]. Therefore a situation is created which is opposite to that observed for [Ru(bpy)2(Et2-dcbpy)]2+, the first reduction of [Ru(bpy)2(phen)]2+ generates a radical where the electron can hop between two equivalent bpy ligands, the second electron will be placed in the second bpy ligand, so that electron hopping can only occur between the bpy and phen ligand. This assumption is confirmed by the
2+ / + , + / 0 ) , is found to be 0.24mVK1 and the value for results, the value for the first reduction, E (12

the temperature dependence for the second reduction process is 0.06mVK1. It is noted, that the third reduction process for this compound lost its chemical reversibility above 30C under the differential pulse conditions used, which was observed as a minor broadening of the peak and decrease in peak height. Therefore only data obtained below 20C was used for the temperature relation where a linear relationship was found. The remaining two complexes, [Ru(bpy)(4,7-CH3-phen)(5,6-CH3-phen)]2+ and [Ru(bpy)(phen)(4,4'-CH3-bpy)]2+, exhibit systems where any electron hopping can only occur between non-

Chapter 3: Studies on Model Esters Compounds Appendix

page: 110

equivalent ligands. Shown in Figure 3.28 are the plots of temperature versus potential separation for these two complexes, which again result in a linear dependence. The slopes for both processes,
2+ / + , + / 0 ) + / 0 ,0 / ) E (12 and E (12 , of [Ru(bpy)(4,7-CH3-phen)(5,6-CH3-phen)]2+ are well defined, for

[Ru(bpy)(phen)(4,4'-CH3-bpy)]2+ the third reduction process is not fully reversible at higher temperatures under the differential pulse conditions used, resulting in a scattering of data points for
+ / 0 ,0 / ) value, although the expected trend can still be seen (Figure 3.28d). the E (12

3.7.4. Conclusion It has been shown that a linear dependence of potential separations with temperature is also found with a series of [RuXYZ]2+ complexes. It appears, that the amount of temperature dependence, which equals
2S electron hopping F

(see Equation 3.24), is a measure of the quality of orbital

overlap between the participating moieties and therefore is also a measure of oscillator strength and intensity of optical transitions. For instance, the inter-valence charge-transfer for the [(bpy)2Ru(py)2]2+ (py = pyridine) and [(bpy)3Ru]2+ system [151] occurs at significantly less energy and less absorbance than for the (Et2-dcbpy)2RuX2 system (see chapter 3.5.2(e) for details). This would imply that electron hopping occurs more readily between the Et2-dcbpy ligands than between the bpy ligands, thus, complexes where electron hopping can occur between Et2-dcbpy ligands should exhibit a larger temperature dependence than where the hopping occurs between bpy ligands. This assumption is confirmed by the results obtained in this study. The two systems investigated in this appendix, where electron hopping occurs between two bpy ligands, exhibit values of 0.24mVK1 and 0.38mVK1 (see Table 3.10) for the temperature dependence of the potential separation. The two systems investigated earlier in this chapter, where the hopping takes place between Et2-dcbpy ligands, have larger values of 0.46mVK1 and 0.42mVK1 (see Table 3.6). So far, electron hopping has only been measured by variable temperature ESR spectroscopy and near infra-red spectroscopy. This work suggests that a third method is available in which redox potentials are measured as a function of temperature. Furthermore, it has been shown that electron hopping also occurs between non-equal ligands. Certainly more experimental work is required to establish the assumptions raised in the text. It will be desirable to find ligands which provide complexes with a highly reversible voltammetric behavior in order to be able to investigate a large range of reversible redox reactions over a large temperature range. Et2-dcbpy is a good candidate, since its reduction potential is 0.5V more positive than bpy. An even better candidate for this purpose may be its analogue with the carbon ester groups in the 5-position, 5,5'-diethoxy-2,2'bipyridine-4,4'-dicarboxylic acid, which can be reduced at more positive potentials (0.32V more positive than 4,4'-Et2-dcbpy [142]).

Chapter 4

part of this work has been published in

'Voltammetric Determination of the Reversible Redox Potential for the Oxidation of the Highly Surface Active Polypyridyl Ruthenium Photovoltaic Sensitizer RuII(dcbpy)2(NCS)2 (dcbpy = 2,2-bipyridine-4,4-dicarboxylic acid)' A.M. Bond, G. Deacon, J. Howitt, D. MacFarlane, L. Spiccia, G. Wolfbauer, J. Electrochem. Soc., 1999, 146(2), 648656 'Determination of Reversible Redox Potentials of Highly Surface Active Polypyridine Ruthenium Complexes' G. Wolfbauer, A.M. Bond, J.A. Howitt, D.R. MacFarlane, Abstracts of RACI Inorganic Chemistry Division Conference, 1998, Wollongong, Australia, poster PT85

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Introduction

page: 112

4.1. Introduction
4.1.1. General The most common sensitizer used [29, 72] in ruthenium-titania photovoltaic systems is cis-(H2dcbpy)2Ru(NCS)2 (H2-dcbpy = 2,2'-bipyridine-4,4'-dicarboxylic acid) and the structure is contained in Figure 4.1.
COOH

HOOC N

N N Ru N N N C S C S

COOH

HOOC

Figure 4.1. Structure of cis-(H2-dcbpy)2Ru(NCS)2.

This sensitizer provides excellent absorption in the visible spectrum, a high electron injection rate, high turn over rates and high stability in photo electrochemical cells. However, whilst the photophysical properties of cis-(H2-dcbpy)2Ru(NCS)2 have been studied [67, 86, 93], relatively little information on the solution phase redox chemistry is available (but see [29]) even though the reversible potential of the Ru(III)/Ru(II) couple is an essential component in the thermodynamics of the photovoltaic cell (Reaction Scheme 4.1). excitation: e injection: dye regeneration: electrolyte regeneration: RuII(dcbpy2)(NCS)2 [RuIII(dcbpy)(dcbpy)(NCS)2]* [RuIII(dcbpy2)(NCS)2]+ + El [RuIII(dcbpy)(dcbpy)(NCS)2]* [RuIII(dcbpy2)(NCS)2]+ + e RuII(dcbpy2)(NCS)2 + El El

El + e

Reaction Scheme 4.1. Schematic representation of reactions that occur in a photoelectrochemical cell which demonstrates that both the ground and excited state redox potentials of the cis-[Ru(dcbpy2)(NCS)2]+/0 couple are important. El and El are the oxidized and reduced forms of the electrolyte (e.g. commonly I3 and I respectively).

On the basis of studies on simple ruthenium bipyridyl complexes available in literature [144, 145, 150] it would be expected that conventional cyclic voltammetric techniques at macrodisk electrodes could be used in a straight forward manner to measure the reversible potential of the metal based ground state [(H2-dcbpy)2Ru(NCS)2]+/0 redox couple. However, since cis-(H2dcbpy)2Ru(NCS)2 has been designed to be attached to electrode surfaces, the likelihood of surface activity of the reduced or oxidized forms of the compound onto the electrode surface is substantial. Importantly, if surface based reactions are coupled with the charge transfer process then the application of techniques such as cyclic voltammetry and diffusion controlled theory, as widely

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Introduction

page: 113

used in studies on [Ru(bpy)3]2+ (bpy = 2,2'-bipyridine), may not provide correct values of the reversible potential. For example Wopschall and Shain [189-191] have demonstrated the dramatic effect that reactant or product adsorption may have on cyclic voltammograms. In this chapter it is demonstrated that the combined use of a range of electrode materials, electrochemical techniques (cyclic, rotating disk and microelectrode voltammetry) and chemical modification of the electrode surface is required to ensure that the effects of surface and other reactions present with the photovoltaic sensitizer are minimized, so that the required reversible potential of the mass transport controlled solution phase process is correctly measured. The extent of the surface activity is also monitored by the electrochemical quartz crystal microbalance method. 4.1.2. Aim of the Work From data obtained in the previous chapter on the esterified analogues of cis-(H2dcbpy)2Ru(NCS)2, the voltammetric and chemical behavior of the oxidized state of cis-(H2dcbpy)2Ru(NCS)2 might be deduced. The esterified chloride analogue, cis-(Et2-dcbpy)2RuCl2, shows a highly reversible cis-[(Et2-dcbpy)2RuCl2]0/+ process, enabling cis-[(Et2-dcbpy)2RuCl2]+ to be characterized. However, the isothiocyanate analogue exhibits a significantly decreased stability for its singly oxidized state, specifically in dimethylformamide (DMF) where no conditions were found under which reversible voltammetric behavior was obtained. Furthermore, the oxidation process of cis-(Et2-dcbpy)2Ru(NCS)2 is accompanied by surface based processes, which result in significant deviations from the theoretically expected voltammetric waveform. Since the origin of these surface based processes is believed to be caused by the isothiocyanate groups, these effects are also expected to occur in case of cis-[(H2-dcbpy)2Ru(NCS)2]. As pointed out above, the potential for the oxidation process of cis-(H2-dcbpy)2Ru(NCS)2 is an essential thermodynamic value for describing and understanding the physical properties of a photoelectrochemical cell. The major aim of work described in this chapter is to establish accurate values for the potentials of the oxidation process of (H2-dcbpy)2Ru(NCS)2 in various solvents. This will require conditions under which reversible voltammetry for the [(H2-dcbpy)2Ru(NCS)2]0/+ process can be found and which minimize deviations from the theoretically required mass transport controlled conditions caused by surface based processes.

4.2. Experimental
4.2.1. Materials and Equipment The solvents, dimethylformamide (DMF), acetone, acetonitrile, and tetrahydrofuran (THF) were of HPLC (Mallinckrodt, UltimAR or ChromAR) grade and used without further purification. Tetrabutylammonium hexafluorophosphate (Bu4NPF6) was prepared according to a literature method [112] and dried in vacuum at 90C for 10 hours. Tetrahexylammonium perchlorate (Hex 4NClO4)

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Experimental

page: 114

(GFS Chemicals), tetrabutylammonium thiocyanate (Bu4NSCN) (Aldrich), [Ru(bpy)3](PF6)2 (Aldrich), Ferrocene (Aldrich) and 4,4'-bipyridyldisulfide (SS-bpy) (Aldrich, under the name Aldrithiol-4) were used as supplied by the manufacturer. The majority of the voltammetric measurements were carried out using a standard three-electrode arrangement and employing electrochemical equipment as described in Chapter 3.2 on page 62. Additionally, a MacLab/4e (ADInstruments) potentiostat was used. For solid state voltammetric measurements see Chapter 3.2 on page 62. In all experiments, the reference electrode was calibrated against the potential of the ferrocene/ferricenium (Fc/Fc+) redox couple via measurement of the ferrocene oxidation process under the same conditions (solvent and electrolyte) as used in the experiments on the ruthenium sensitizer. The temperature at which experiments were undertaken was 20(2)C. Electrochemical quartz crystal microbalance (EQCM) studies were carried out on an ELCHEMA-701 instrument (Postdam, New York) in conjunction with an ELCHEMA PS-205 potentiostat and Voltscan software (Intellect Software, Postdam, New York). 10 MHz AT-cut gold sputtered quartz crystals were purchased from Bright Star Crystals (Vermont, Vic., Australia). The diameter of the gold sputtered area was 5mm. The quartz crystal was mounted onto the side of a glass cell and was held in place by two O-rings. An Extech-1671 conductivity meter was used for conductivity measurements. Solutions of the sensitizer were measured at a concentration of 0.5mM in the respective solvents and corrected for the conductivity of the pure solvent. 4.2.2. Synthesis (a) cis-(H2-dcbpy)2RuCl2 1.50g (5.7mmol) RuCl33H2O and 2.79g (11.4mmol) H2-dcbpy are placed in a 500mL threeneck round bottom flask and about 200mL DMF solvent is added. The solution is bubbled with nitrogen for 10 minutes (N.B. no significant change in yields was observed when this step was omitted). Under vigorous stirring the solution was heated to 145C (160C oil bath temperature). The course of the reaction was monitored by UV/VIS spectroscopy, in a similar manner as described for synthesis of the ester analogue (see chapter 3.3.1 on page 65 for details). The reaction was stopped (2.5h) when the metal-to-ligand charge-transfer (MLCT) band at about 558nm had reached its relative maximum value compared to the higher energy MLCT band at 411nm and the ligand charge transfer band at 314nm. Typical values for the relative intensities that indicate completion of the reaction are 1 : 0.95 : 3.00 for bands at 558nm : 411nm : 315nm. When the reaction was not stopped in time the yield and purity of the sample decreased considerably.

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Experimental

page: 115
COOH

COOH HOOC RuCl3 3(H2O) +


N

COOH

HOOC HOOC N Ru N HOOC Cl Cl -OH -OH N N COOH + KSCN 150C DMF/H2O 3h HOOC N

N N Ru N N C S N C S

COOH

145C DMF ca. 2.5h

H2-dcbpy

cis-(H2-dcbpy)2RuCl2 85%

cis-(H2-dcbpy)2Ru(NCS)2 35-70%

Figure 4.2. Reaction scheme for the synthesis of cis-(H2-dcbpy)2Ru(NCS)2.

The hot solution was filtered through a G4 frit to remove any unwanted [Ru(H2-dcbpy)3]2+. The solvent was removed by rotary evaporation and the solid transferred with aqueous 2M HCl to a jar, centrifuged, washed twice with 2M HCl and the remaining gelatinous precipitate freeze dried over night. Yield: 3.63g (85%) cis-(H2-dcbpy)2RuCl2. For synthesis of cis-(H2-dcbpy)2Ru(NCS)2 no attempts are made to purify the intermediate chloride product, which usually contained mixtures of cis-(H2-dcbpy)2RuX2, where X = Cl or OH, and small amounts of DMF solvent. (b) cis-(H2-dcbpy)2Ru(NCS)2 1g (1.24mmol) cis-(H2-dcbpy)2RuCl2 was dissolved in 50mL DMF. 3.6g KSCN (37.2mmol, 30 fold excess) were dissolved in a minimal amount of water and added to the former solution. Alternatively, NH4NCS may be used, which does not require dissolution in water. The solution was gently heated to reflux and vigorously stirred. After three hours the solvent is removed by rotary evaporation and the remaining solid dissolved in a small amount of aqueous 0.05M NaOH solution. 20% aqueous CF3COOH solution is added dropwise to the solution until pH2.2 is reached, which causes the product to precipitate. The solution is centrifuged and washed twice with H2O (adjusted to pH 2.2 with CF3COOH). The solid is dissolved in 10% aqueous n-tetrabutylammonium hydroxide, Bu4N(OH), solution and very carefully titrated to pH 3.5 with 20% aqueous CF3COOH solution to precipitate the (Bu4N)2-salt. The precipitate is again centrifuged and washed with H2O. (Note: The latter step, precipitation of the (Bu4N)2-salt, might be omitted when the amount of impurities is low.) To the solid crude product, as obtained from centrifugation, a minimal amount (510ml) of aqueous NaOH solution is added until a pH of 810 is reached, which is sufficient by basic to completely dissolve the product. The solution is stirred at least for an hour and the pH is checked to ensure that all of the material has dissolved. Chromatography of this solution is then undertaken on Sephadex LH20 gel-exclusion resin with water as the eluent. Several bands separate on the column, but only one broad dark-purple band forms, which contains the desired product. All other fractions are rejected. Four to six fraction are collected from the product band. It is essential to collect such a large number of fractions, since thiocyanate linkage isomers of cis-(H2-dcbpy)2Ru(NCS)2 to some extent separate on the column enabling enrichment of the desired linkage isomer. The products are precipitated by titrating the solutions to pH 1.61.8 and separated by centrifugation followed by freeze drying. Yield: 0.64g (70%)0.32g (35%) of cis-(H2-dcbpy)2Ru(NCS)22H2O.

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Experimental

page: 116

S,S N,S

N,S

d)

S,S N,S

N,S

c)

N,S

N,S

b)

N,S

N,S

a)
10.0 9.5 9.0 8.5 8.0 7.5

ppm
Figure 4.3. 1H-NMR spectra of cis-(H2-dcbpy)2Ru(NCS)2 and thiocyanate linkage isomers (indicated by arrows). Spectra referenced versus (CH3)4Si (d6-DMSO) or C3D7COOD (D2O). a) less than 2% N-,S- isomer, solvent d6-DMSO; b) 9.6% N-,S- isomer, solvent 0.1M NaOD/D2O; c) 33% N-,S- isomer, 3.5% S-,S- isomer solvent 0.1M NaOD/D2O; d) 56% N-,S- isomer, 12% S-,S- isomer, solvent 0.1M NaOD/D2O.

NMR spectroscopy, carried out in 0.1M NaOD/D2O, shows, that the fractions collected last contain the largest amount of the desired doubly N-,N- bound thiocyanate ligand complex (see Figure 4.3a), cis-(H2-dcbpy)2Ru(NCS)2, whereas fractions taken at an earlier stage indicate mixtures of cis-(H2-dcbpy)2Ru(NCS)2 and the mixed -N,-S linkage isomer, cis-(H2dcbpy)2Ru(NCS)(SCN) (see Figure 4.3b). The assignment of the NMR signals for both isomers can be found in the literature [80]. Authors of this study [80] have determined the coordination mode of the thiocyanate ligand in this complex and found that thiocyanate primarily coordinates via the sulfur (kinetically preferred), whereas upon extended reaction time or higher temperature the thermodynamic favorable doubly N-,N- bond linkage isomer is formed. With this chromatography procedure cis-(H2-dcbpy)2Ru(NCS)2 can be obtained with 98% purity, only containing 2% of the mixed N-, S- bond thiocyanate linkage isomer (Figure 4.3a). When the reaction temperature and time for the thiocyanate substitution step is lowered to 100C and one hour, respectively, mixtures of N-,N- and N-,S- bond linkage isomers of up to 35% cis-(H2-dcbpy)2Ru(NCS)(SCN) content may be obtained. Mixtures with higher contents of the N-,S- isomer already exhibit significant amounts of the doubly S-,S- bond linkage isomer (see Figure 4.3c and d) as well as other impurities of unknown origin. NMR spectroscopy provides further proof that the ruthenium complexes formed have cis configuration and no trace of the trans isomer are found, which should be readily

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Experimental

page: 117

distinguishable from the cis isomer due to its higher symmetry. For convenience the cis prefix will be omitted in the remainder of this chapter.

4.3. Results and Discussion


4.3.1. Cyclic Voltammetry of (H2-dcbpy)2Ru(NCS)2 in Acetone It was expected, after comparison of the structures of (H2-dcbpy)2Ru(NCS)2 and (Et2dcbpy)2RuCl2, which was investigated in the previous chapter, that (H2-dcbpy)2Ru(NCS)2 also would exhibit one reversible metal based oxidation process, although irreversible processes due to oxidation of the thiocyanate ligand also could be expected [192]. However, results obtained from the model compound (Et2-dcbpy)2Ru(NCS)2 (see chapter 3.4.4 on page 83) indicate an unexpected significantly more complex oxidation process. A cyclic voltammogram for the oxidation of (H 2dcbpy)2Ru(NCS)2 in acetone is shown in Figure 4.4.
100 80 60 40 20 0 -20

current [A]

-0.5

0.0

0.5
+

1.0

potential [V] vs. Fc/Fc

Figure 4.4. Cyclic voltammogram for the oxidation of 1.1mM (H2-dcbpy)Ru(NCS)2 in acetone (0.1M Bu4NPF6). =100mVs1, 2mm diameter glassy carbon working electrode.

Simple inspection of the cyclic voltammogram for the oxidation of the dye (Figure 4.4) shows that several oxidation processes are present in acetone and that none of them is a simple diffusion controlled reversible process. The even more complex reduction processes are discussed in the next chapter. On the basis of analysis of cyclic voltammograms of the kind as shown in Figures 4.4 and 4.5, the mechanism for the oxidation of (H2-dcbpy)2Ru(NCS)2 is postulated to involve the expected solution phase one-electron charge transfer process coupled with surface based processes that occur prior to and after oxidation as well as oxidation of surface modified species at positive potentials. Equations 4.1 to 4.5 summarize the basic features of the proposed mechanism. Dsol Dads Dads Xsurf (4.1) (4.2)

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 118

Dsol

D+sol + D+sol Ysurf

(4.3) (4.4) (4.5)

X surf (or Ysurf)

X+surf (or Y+surf) + e product(s)

Where Dsol and Dads represent the dye molecule, (H2-dcbpy)2Ru(NCS)2 in the solution and adsorbed states respectively, and Xsurf or Ysurf represent surface attached forms of the dye, whose identities are unknown, but which give rise to the second and other oxidative processes (Equation 4.5). As theoretically predicted, when surface based phenomena are coupled to the electron transfer step, the shape and nature of voltammograms are strongly affected by the nature of the working electrode, the scan rate, the extent of potential cycling and the concentration. A series of voltammograms recorded under various conditions is shown in Figure 4.5.

e)

2A

d)

20A

c)

50nA

b)

40A

a)
-0.2 0.0

6A

0.2

0.4

0.6

potential [V] vs. Fc/Fc

0.8

1.0

Figure 4.5. Cyclic voltammograms of (H2-dcbpy)2Ru(NCS)2 in acetone (0.1M Bu4NPF6) obtained at different electrode materials, concentrations and scan rates. a) c=0.5mM, =50mVs1, electrode: glassy carbon d=2mm; b) c=0.5mM, =2000mVs1, electrode: glassy carbon d=2mm; c) c=0.2mM, =50mVs1, electrode: platinum d=1mm; d) c=0.5mM, =1000mVs1, electrode: glassy carbon d=2mm; e) c=0.5mM, =100mVs1, electrode: glassy carbon d=2mm.

The amount of material attached to the surface is related to the current magnitude of the processes at positive potentials, and was observed to be considerably less on platinum electrodes

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 119

relative to glassy carbon and gold electrodes (compare Figure 4.5c). The scan rate influences both the ipox/ipred ratio of the first oxidation process and the prominence of the second oxidation peak (see Figure 4.5b). However this interrelationship is expected, since the primary oxidation process (Equation 4.3) becomes increasingly diffusion controlled at fast scan rates when it would be predicted that the extent or formation of D+sol and hence Ysurf is minimized. The ratio of ipox/ipred was never unity as required for a completely diffusion controlled process, but at fast scan rates it approached 2:1 (see for example the cyclic voltammogram obtained at moderately fast scan rates of 1000mVs1 in Figure 4.5d). A ratio greater than unity is expected when reactant adsorption occurs [189]. The dependence on the scan rate mentioned above is clearly revealed when the voltammograms contained in Figures 4.5a and 4.5b are compared. Furthermore, it should be noted, that at a concentration of 0.5mM and with a glassy carbon macrodisk electrode the second peak is almost absent at a scan rate of 2000mVs1, whereas at a scan rate of 50mVs1 the second process is quite pronounced. Also noteworthy is the significant difference between the first and subsequent cycles (Figure 4.5e). Apparently, during the course of sweeping the potential, surface attached material is removed from the surface. Thus when the potential is returned to the initial value prior to commencing the second cycle, less time is available for reactant to become attracted to the electrode surface than is the case with the initial cycle where the electrode is held at the initial potential value for a significantly longer period of time. As expected, the reactant concentration also markedly influences the contribution of the surface based processes. For example, the ratio of the peak heights of the second and first oxidation process obtained from a 0.2mM solution of (H2dcbpy)2Ru(NCS)2 (Figure 4.5c) is much smaller than obtained when the concentration 0.5mM. For a solely diffusion controlled process, the reversible potential is expected to be approximated as the average value of Epox and Epred ((Epox + Epred)/2) as is the case for the ideal [(Et2dcbpy)2RuCl2]0/+ system considered in the previous chapter. The data in Table 4.1 actually indicate that the potential of the first oxidation process calculated in this manner is remarkably insensitive to the experimental conditions, unlike other features of the voltammetry. The (Epox + Epred)/2 potentials listed in Table 4.1 were all calculated from the third cycle of the potential, but no significant difference was found in values obtained from other cycles. The peak-to-peak separations are clearly larger than those obtained for oxidation of (Et2-dcbpy)2RuCl2 (compare Table 3.2 on page 69) or ferrocene at the same scan rates, and this feature may be attributed to the surface effects. However, the calculated (Epox + Epred)/2 value of 0.410.01V vs. Fc/Fc+ for the [(Et2dcbpy)2Ru(NCS)2]0/+ couple was found to be independent of concentration (0.05 to 0.5mM), scan rate (50 to 2000mVs1) and electrode material (platinum, glassy carbon and gold) at the 10mV uncertainty level implying that this is a good approximation of the reversible potential (Er1/2-value) for the solution phase diffusion controlled process. Thus, the influence of the surface based reactions is not significant in the thermodynamic sense, with respect to the first oxidation process.

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 120

Glassy Carbon Electrode conc. [mM] 0.5 [mVs1] 50 100 500 1000 2000 50 100 500 1000 2000 50 100 500 1000 2000 50 100 500 1000 2000 50 100 500 1000 2000 Ep [V]
ox

Platinum Electrode Er1/2 [V] 0.40 0.40 0.40 0.40 0.41 0.41 0.41 0.41 0.40 0.41 0.41 0.40 0.40 0.40 0.40 0.41 0.41 0.41 0.42 0.42 Ep [V]
ox

Ep [V]

red

Ep [mV] 140 160 220 120 130 150 230 110 120 150 210 120 120 160 200 110 100 100 110 170

Epred [V] 0.34 0.32 0.29 0.34 0.34 0.31 0.35 0.34 0.34 0.31 0.33 0.35 0.34 0.33 0.31 0.35 0.35 0.35 -

Ep [mV] 130 160 220 130 140 200 120 120 140 200 130 110 120 150 190 110 110 110 -

Er1/2 [V] 0.41 0.40 0.40 0.41 0.41 0.41 0.41 0.40 0.41 0.41 0.40 0.41 0.40 0.41 0.41 0.41 0.41 0.41 -

0.4

0.3

0.2

0.1

0.48 0.47 0.47 0.48 0.51 0.47 0.46 0.47 0.48 0.52 0.47 0.46 0.46 0.48 0.51 0.47 0.46 0.46 0.48 0.50 0.46 0.46 0.46 0.47 0.50

0.33 0.32 0.29 0.34 0.34 0.33 0.29 0.35 0.34 0.33 0.30 0.34 0.34 0.32 0.30 0.35 0.36 0.36 0.36 0.33

0.47 0.47 0.47 0.48 0.34 0.47 0.47 0.47 0.48 0.51 0.47 0.47 0.46 0.48 0.51 0.46 0.46 0.46 0.48 0.50 0.46 0.46 0.46 0.47 -

Table 4.1. Cyclic voltammetric data obtained for the oxidation of (H2-dcbpy)2Ru(NCS)2 in acetone (0.1 M Bu4NPF6) at 1mm diameter platinum and 2mm diameter glassy carbon macrodisk electrodes as a function of scan rate and concentration. All peak potentials are reported versus Fc/Fc+, uncertainty of 0.01V, Er1/2values calculated as (Epox + Epred)/2.

4.3.2. Voltammetric Oxidation of (H2-dcbpy)2Ru(NCS)2 in Acetone at a Microelectrode The steady-state voltammetry of (H2-dcbpy)2Ru(NCS)2 (0.05 to 0.5mM) at a platinum microdisk electrode in acetone shows no evidence of a second oxidative process (Figure 4.6a). In contrast, at a glassy carbon microdisk electrode there is a small second oxidative response which is enhanced at higher concentrations. It would appear that the greater flux of material toward (and away from) a microelectrode due to radial diffusion, which dominates under steady state conditions, has swept the majority of the oxidized material away from the electrode surface before significant interaction occurs with the surface. Furthermore, the current contribution from surface attached material is proportional to the electrode area, hence iads r2, whereas the current due to radial diffusion as encountered under microdisk electrode conditions is directly proportional to electrode radius, ilim r (Equation 2.28 on page 48). Thus, when decreasing the electrode size of microdisk electrodes the current contribution due to radial diffusion will decrease slower than the surface attached material component, which eventually will become negligible at sufficiently small electrode sizes. However, under cyclic voltammetric conditions, both current components are proportional to the square of the electrode radius (see Rendles Sevcik Equation 2.18 on page 43) and therefore no effect of electrode size on voltammograms is expected and was observed.

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 121

a)
10nA

20A

b)
-0.2 0.0 0.2 0.4 0.6
+

0.8

1.0

potential [V] vs. Fc/Fc

Figure 4.6. Steady state voltammograms of (H2-dcbpy)2Ru(NCS)2 in acetone (0.1M Bu4NPF6). a) Glassy carbon microdisk Electrode, r=6m, c=1.1mM, =10mVs1 b) 3mm diameter glassy carbon rotating disk electrode, =500, 1000, 1500, 2000, 2500 and 3000min1, =10mVs1.

The Er1/2-value obtained from microdisk electrode experiments was found to be 0.40(0.01)V vs. Fc/Fc+. Assuming that the limiting current is diffusion controlled under these near steady-state conditions and use of Equation 2.28, gives a calculated diffusion coefficient of 9.5(0.5)106cm2s1. A "log-plot" obtained under steady-state conditions at a microdisk electrode is shown Figure 4.7b. The slope is linear as expected for a diffusion controlled Nernstian process. The slightly larger slope of 72mV than the 59mV theoretically predicted for a one-electron transfer is probably due to the process being not fully chemically reversible at the 11m diameter microdisk electrode.
0.6 0.4 0.2
0.6

a)
log((iL-i)/i)

0.4 0.2 0.0 -0.2 -0.4

b)

log((iL-i)/i)

0.0 -0.2 -0.4 -0.6

380

400

420

440
+

460

-0.6

380

400

420

440
+

460

potential [mV] vs. Fc/Fc

potential [mV] vs. Fc/Fc

Figure 4.7. Log-plots obtained from rotating disk and microdisk electrode steady-state voltammograms shown in Figure 4.6.. a) Rotating disk electrode 500min1, slope: 69mV, b) microdisk electrode, slope: 72mV.

4.3.3. Rotating Disk Voltammetry of (H2-dcbpy)2Ru(NCS)2 in Acetone The rotating disk voltammetry of (H2-dcbpy)2Ru(NCS)2 in acetone is consistent with the suggestion that the second oxidation process is due to oxidation of surface attached material. At the lower rotation rates and at a concentration of 1.1mM, the second process is of almost the same size as the first, but as the rotation rate is increased the relative size of the second process decreases,

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 122

until at sufficiently high rotation rates it eventually disappears (Figure 4.6b). Again, the second process is less evident at platinum than glassy carbon or gold electrodes. The increased hydrodynamic flux of material at the electrode surface due to high rotation rates is able to sweep products away from the surface so they have less time to interact. This feature of the rotating disk electrode experiment results in the elimination of the second peak as is the case at a microdisk electrode. A plot of limiting current is dependent on the square root of the rotation frequency at a platinum electrode over the range of 500 to 3000min1 and passes through the origin. The |E3/4-E1/4| for the first oxidation is 66(3)mV at a rotation frequency of 3000min1 and 68(3)mV at a rotation frequency of 500min1. These values are in very good agreement with values calculated from logplots (see Figure 4.7a), where a value for the slope of 69mV is obtained for a rotation rate of 500rpm. Under these conditions the process is close to a mass transport controlled reversible oneelectron oxidation. The E1/2-value is independent of rotation rate and electrode material and is located at 0.409(0.002)V vs. Fc/Fc+ which is in agreement with values calculated from cyclic voltammetric and microdisk electrode voltammograms. Consequently, both the microdisk steadystate and rotating disk techniques give almost ideal voltammetric behavior for a reversible mass transport controlled one electron oxidation process. 4.3.4. EQCM Studies in Acetone To confirm the presence of the surface based processes, studies were carried out on gold electrodes using the electrochemical quartz crystal microbalance (EQCM) technique, where small mass changes on surfaces occurring during voltammetric experiments can be detected, assuming that the Sauerbrey equation (Eqn. 2.40 on page 57) is valid [193]. Figure 4.8 shows the EQCM response at the open circuit potential for an acetone solution (0.1mM Bu4NPF6) before and after the solution is spiked with 0.6mM (H2-dcbpy)2Ru(NCS)2. Apparently, a large amount of material becomes attached to the gold electrode under open circuit conditions at a stationary electrode. Under the conditions of Figure 4.8 a rapid mass increase of 410ng occurred onto the surface of the 5mm gold electrode, which, assuming a flat surface and that the material is adsorbed in an unreacted form, corresponds to a surface coverage of =3109mol cm2. The surface coverage was calculated assuming a flat electrode surface and the molecules were spheres having a diameter of 14. This diameter was estimated from X-ray structural data [38, 194]. If cubic packing for the molecule is assumed, a surface coverage corresponding to 30 mono-layers is calculated at the open circuit potential.

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 123

600

mass change [ng]

400
(H2-dcbpy)2Ru(NCS)2 added

200

0 -100 0 100 200 300 400 500 600

time [s]
Figure 4.8. Electrochemical Quartz Crystal Microbalance experiments in an acetone (0.1M Bu 4NPF6) solution which is spiked with (H2-dcbpy)2Ru(NCS)2 at the time indicated to give a final concentration of 0.6mM. Open circuit potential, 5mm diameter gold electrode.

EQCM experiments during a cyclic voltammetric experiment in acetone on a gold electrode give the results shown in Figure 4.9a . Results from this experiment show that holding the potential at the initial value of 450mV vs. Fc/Fc+ for 120s and then scanning until a potential of about 200mV is reached, leads to only a small decrease of material initially attached to the gold electrode. However, at about 200mV vs. Fc/Fc+ which corresponds to the onset of the oxidation process (Equation 4.5), a large amount of material is rapidly removed from the electrode surface. The amount lost of ca. 350ng is almost equal to the amount originally adsorbed onto the surface under open circuit conditions. When the potential is scanned to values slightly beyond the half wave potential, a very abrupt mass increase occurs, but as more positive potentials are reached, the mass of material attached to the electrode decreases to a value which is similar to that at the commencement of the experiment. Incorporation of PF6, the electrolyte anion, into solid attached to the electrode surface may occur after oxidation in order to achieve charge neutralization. Figure 4.9b shows that the mass fluctuations became much smaller on repetitive cycling of the potential until an almost constant mass situation is reached, with respect to the potential cycle number. The EQCM data are consistent with the mechanism proposed in Equations 4.1 to 4.5. The identities of the species attached to the electrode surface and associated with the second oxidation process at more positive potentials are unknown. Bron and Holze [195] investigated the adsorption of free thiocyanate onto gold electrodes. They found a strong affinity of thiocyanate toward gold electrodes over a very wide potential range. The thiocyanate ions adsorbed mainly through the sulfur atom, although an observed potential dependence was related to whether the adsorption occurred via the nitrogen or the sulfur. In case of (H2-dcbpy)2Ru(NCS)2, the thiocyanate ligands are coordinated via the nitrogen, which means that the sulfur atom is available for surface interactions. Another study [196] reports adsorption of the free H2-dcbpy ligand onto platinum electrodes, therefore electrode surface attachment via the carboxylate group could also occur as is

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 124

postulated at semi conductor titania electrodes [67-69]. On "soft" metal electrodes, sulfur-metal interactions could well be dominant.
50

mass change [ng]

400

a)

40

current [A]

200

30 20 10

-200 0 0 50 100 150 200 250


60

time [s]
0

mass change [ng]

-100 -200 -300 -400 -500 0 200 400 600

b)
40

current [A]

20

0 800

time [s]
Figure 4.9. Electrochemical Quartz Crystal Microbalance experiments on 0.5mM (H2-dcbpy)2Ru(NCS)2 in acetone (0.1M Bu4NPF6), 5mm diameter gold electrode, =20mVs1, dashed red line shows applied potential waveform (arbitrary units), Estart=0.45V, Eswitch=0.94V, vertical dashed line indicates start of potential scan. a) one scan, b) successive scans.

Free SCN added as Bu4NSCN was found to be irreversibly oxidized at a potential of 0.18V, 0.12V and 0.23V vs. Fc/Fc+ under conditions of cyclic voltammetry (scan rate 100mVs1) in acetone (0.1M Bu4NPF6) at platinum, glassy carbon and gold macrodisk electrodes respectively. Thus, the second oxidation process is not associable with oxidation of SCN released after oxidation of (H2-dcbpy)2Ru(NCS)2, although it could be attributable to oxidation of coordinated thiocyanate associated with surface attached form of the dye. However, the process is not associated with oxidation of the thiocyanate ligand coordinated to solution soluble [(H 2dcbpy)2Ru(NCS)2]+ generated via oxidation of (H2-dcbpy)2Ru(NCS)2, since the second process is absent when the [(H2-dcbpy)2Ru(NCS)2]0/+ process is reversible and diffusion controlled. 4.3.5. Cyclic Voltammetry in Acetone in the Presence of SS-bpy If the postulated mechanism is correct, then adding an electroinactive surfactant to the solution should suppress surface attachment of (H2-dcbpy)2Ru(NCS)2 and thereby eliminate or minimize the

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 125

magnitude of the second oxidation process. The gold electrode surface active substance, 4,4'bipyridyl disulfide (SS-bpy), has been widely used [197, 198] as an electrode modifier, and for example, this compound assists the detection of well defined voltammograms for cytochrome c. SS-bpy forms electroinactive clusters on the gold electrode surface at low modifier concentrations and a monolayer coverage at high concentrations.

12

current [A]

0 -0.2 0.0 0.2 0.4 0.6 0.8 1.0

potential [V] vs. Fc/Fc

Figure 4.10. Effect of addition of SS-bpy on the voltammetry of 1.5mM cis-Ru(dcbpy)2(NCS)2 in acetone (0.1M Bu4NPF6), 1mm diameter platinum macrodisk electrode, =100mV/s. - - - - - - - without electrode modifier, after addition of SS-bpy (c=2mM).

Figure 4.10 shows the effect on a voltammogram at a gold macrodisk electrode of adding SSbpy to a 1.5mM solution of (H2-dcbpy)2Ru(NCS)2 in acetone. At equal or higher concentrations of SS-bpy, the processes at positive potentials are eliminated, which implies that a rather high SS-bpy electrode modifier concentration is necessary to compete for the adsorption sites. SS-bpy also modifies the response in an analogous manner at platinum electrodes, but the oxidation processes at positive potentials remained when equivalent experiments were undertaken at the glassy carbon electrode. Reversible [(H2-dcbpy)2Ru(NCS)2]+/0 potential a) Er1/2 [V] d) Acetone THF Acetonitrile DMF +0.41 +0.32 +0.45 +0.39 Second oxidation process b) Epox [V] d) +0.74 +0.83 +0.74 Diffusion coefficient c) D [cm2s1] 9.5106 5106 10106 9106 e)

solvent

Table 4.2. Potentials and diffusion coefficient data for (H2-dcbpy)2Ru(NCS)2 in different solvents (0.1M Bu4NPF6). a) Calculated as (Epox + Epox)/2 from voltammograms at a platinum macrodisk electrode under conditions where chemical reversibility is obtained. b) Oxidation peak potential at glassy carbon electrode at a scan rate of 100mVs1. c) Calculated from platinum microdisk electrode experiments under near steady state conditions. d) Volt vs. Fc/Fc+. e) Maximum value. Diffusion controlled process with n=1 assumed, but evidence that n>1 was obtained (see text).

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 126

4.3.6. Voltammetry of (H2-dcbpy)2Ru(NCS)2 in THF, Acetonitrile and DMF Er1/2 and diffusion coefficient data obtained from voltammograms of (H2-dcbpy)2Ru(NCS)2 in a range of solvents are summarized in Table 4.2. Voltammetry in THF was similar to that in acetone. However, due to the limited positive potential range available in this solvent, detection of the second oxidation process was difficult. Under conditions of cyclic voltammetry in THF at macrodisk electrodes, chemical reversibility was evident at lower scan rates and higher (H2-dcbpy)2Ru(NCS)2 concentrations than in acetone, presumably due to reduced surface effects. Determining the half wave potential by steady state voltammetry at microdisk electrodes was straight forward, with a reversible one one-electron process being observed without evidence of adsorption. Half-wave potentials determined by the RDE method gave slightly more positive values, which is attributed to the presence of IR (ohmic) drop resulting from the high resistance of THF.

0.6

current [A]

0.4

0.2

0.0

-0.2

-0.5

0.0

0.5
+

1.0

potential [V] vs. Fc/Fc

Figure 4.11. Cyclic voltammogram of a 3.5105M solution of (H2-dcbpy)2Ru(NCS)2 in acetonitrile (0.1M Bu4NPF6), 2mm diameter platinum macrodisk electrode, =500mVs1, background subtracted.

Nazeeruddin et al. [29] have reported a value for the standard potential for oxidation of (H2dcbpy)2Ru(NCS)2 of +0.85V vs. SCE (corrected to +0.47V vs. Fc/Fc+ [175]) in acetonitrile. The value was obtained from cyclic voltammetric measurements, but few details on the mechanism of the process were provided. Furthermore, the oxidation process was described as highly reversible, indicating high stability of the oxidized sensitizer, at least on the cyclic voltammetric time scale. However, an attempt to duplicate the results obtained by Nazeeruddin et al. [29], indicated that pure samples of (H2-dcbpy)2Ru(NCS)2 were only sparingly soluble in acetonitrile, resulting in solutions in the 105M range. Even at such low concentrations evidence of adsorption is revealed under conditions of cyclic voltammetry at a macrodisk electrode (Figure 4.11) and no conditions were found where the oxidation process of (H2-dcbpy)2Ru(NCS)2 was reversible and diffusion controlled in this solvent, which is an essential requirement for deducing lifetimes for the oxidized dye. It should be noted, that the prominence of the reverse peak observed in cyclic voltammograms of (H2-dcbpy)2Ru(NCS)2 in acetonitrile somewhat varied from experiment to experiment, which

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 127

may lead to incorrect assumptions regarding the reversibility of the [(H2-dcbpy)2Ru(NCS)2]0/+ process, similar as has been observed for the ester analogue in CH2Cl2 and acetonitrile (see page 84). The symmetrical waveform of the cathodic process (Figure 4.11) also indicates the presence of surface based processes. Nazeeruddin et al. [29] further stated "From the reversible nature of the cyclic voltammogram at slow scan rates, it can be inferred that if such a process occurred (N.B. decomposition) it would require at least several seconds. ..... more than 107 cycles could be sustained in such a case without the significant destruction of the complex ...." and thus estimated lifetimes of [(H2-dcbpy)2Ru(NCS)2]+ from their cyclic voltammetric data. Based on these lifetimes of the oxidized state, minimal life expectations of solar cells were given. For further discussion see the conclusion section on page 131. Steady-state voltammograms (rotating disk and microdisk electrode) of the oxidation of [(H2dcbpy)2Ru(NCS)2]+ appeared to be reversible and in accordance with the assumed one-electron transfer. No signs of adsorption were found and plots of iL versus the square root of the rotation frequency scaled linearly (rotating disk electrode) indicative the presence of appropriate mass transport conditions. As a result an Er1/2-value of 0.450(0.005)V vs. Fc/Fc+ in acetonitrile can be quoted, which is in good agreement with the value reported by Nazeeruddin and co-workers [29].

1A

a)

2A

b)

-0.5

0.0

0.5
+

1.0

potential [V] vs. Fc/Fc

Figure 4.12. Fast scan cyclic voltammetry of (H2-dcbpy)2Ru(NCS)2 in DMF (0.1M Bu4NPF6) 70m diameter platinum microdisk electrode, background corrected, =1000Vs1. a) 2.5mM cis-Ru(dcbpy)2(NCS)2; b) 4.0mM (H2-dcbpy)2Ru(NCS)2 with ca. 30mM SS-bpy.

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 128

In DMF, transient cyclic voltammograms for oxidation of (H2-dcbpy)2Ru(NCS)2 require the use of very fast scan rates at platinum microdisk electrodes (ca. 200Vs1) before the onset of chemical reversibility becomes evident (Figure 4.12a). Increasing the scan rate beyond 200Vs1, improved the chemical reversibility, but as in acetone, a peak current ratio (ipox/ipred) of unity was never achieved in this solvent. Figure 4.13 shows the ipox/ipred ratio as a function of scan rate for the initial oxidation process in DMF. As in acetone and THF, a peak current ratio of about 3:2 was the maximum obtained, using scan rates in the range 100500Vs1. At high scan rates, the peak current ratio actually decreased and at scan rates above ca. 3000Vs1 no reverse peak could be detected. Under steady-state conditions at microdisk and rotating disk electrodes, the oxidation process in DMF was not chemically reversible and the limiting current value was consistent with a multielectron ECE type process.

current ratio ip /ip


ox red

1 0 10 20 30 40
1/2

50

60

70

(scan rate)

Figure 4.13. Cyclic voltammetric peak current ratio (ipox/ipred) for the oxidation of (H2-dcbpy)2Ru(NCS)2 in DMF (0.1M Bu4NPF6), without () and with (x) added SS-bpy, as a function of scan rate. 70m diameter platinum microdisk electrode, all scans were background corrected.

Addition of SS-bpy simplified the voltammograms obtained in DMF. In the presence of a sufficient concentration of this surface modifier, a close to chemically and electrochemically reversible response was observed at a scan rate at 100Vs1 (Figure 4.12b). Thus, provided ten fold excess of SS-bpy was present, the ipox/ipred ratio was close to unity over the scan rate range of 100Vs1 to 4000Vs1 (Figure 4.13). From these fast scan rate cyclic voltammograms obtained from 2.5mM (H2-dcbpy)2Ru(NCS)2 in the presence of a ten fold concentration excess of SS-bpy, the reversible half-wave potential for the oxidation process in DMF was found to be +0.39(0.01)V vs. Fc/Fc+. 4.3.7. Voltammetry in Acetonitrile of Solid (H2-dcbpy)2Ru(NCS)2 Attached to an Electrode Surface The relatively low solubility of (H2-dcbpy)2Ru(NCS)2 in acetonitrile enabled the electrochemistry of this compound to be examined while attached as a solid to an electrode. Figure 4.14 shows a voltammogram of (H2-dcbpy)2Ru(NCS)2 mechanically attached to a graphite electrode which is placed in acetonitrile (0.1M Hex4NPF6) containing a saturated solution of (H2-dcbpy)2Ru(NCS)2. The resemblance of the response obtained in acetonitrile under these conditions to the solution phase voltammetry in acetone is striking, with even the second oxidation process now being readily

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 129

observed. The first process has a well defined reduction peak, but repetitive cycling of the potential lowers the response as the oxidized product dissolves into the solution phase. The dissolution of oxidized material could be observed visually, as when the potential was scanned past the first oxidation peak, a purple colored solution was generated near the electrode surface. The average value of the oxidation and reduction components of the first oxidation process again leads to a calculated reversible redox potential in acetonitrile of +0.45V vs. Fc/Fc+. Similar experiments were also undertaken in CH2Cl2, where the solubility of (H2dcbpy)2Ru(NCS)2 is even less than in CH3CN, no detectable amount of (H2-dcbpy)2Ru(NCS)2 dissolves in this solvent. When the potential was scanned past the oxidation peak the dissolution of a purple compound from the electrode surface into the clear solution could be observed. In contrast to observation in CH3CN, no reverse peak could be seen and the reverse potential scan, that is the oxidation process was irreversible in this solvent. Furthermore, during consecutive scans, the intensity of the initial oxidation peak decreased significantly faster than in CH 3CN.
250 200

current [A]

150 100 50 0 -50 0.0 0.5 1.0


+

potential [V] vs. Fc/Fc

Figure 4.14. Cyclic voltammograms (first 5 cycles) of solid (H2-dcbpy)2Ru(NCS)2 mechanically attached to a graphite macrodisk electrode (d=5mm) which is then placed in acetonitrile (0.1M Hex4NClO4) presaturated with (H2-dcbpy)2Ru(NCS)2, =200mVs1.

4.3.8. The Influence of Isomers Synthesis of (H2-dcbpy)2Ru(NCS)2 yields an isomeric mixture predominantly containing the N-,N- bonded thiocyanate ligand and small amounts of the unsymmetrical isomer which contains one N- and one S- bonded thiocyanate. The coordination mode of the thiocyanate ligands has been recently confirmed by single crystal X-ray diffraction studies on the ethyl ester analogue [38]. It has been shown previously by NMR measurements [80] that the initial substitution product produced in the synthesis is the S-,S- bonded thiocyanate which is then converted into the N-,Nisomer. Upon refluxing for a substantial period of time in DMF, most of the product can be converted in the N-,N-bonded form [80] (see page 116 for details). To study the influence of the presence of the N-,S- isomer, voltammograms on solutions with N-,N-:N-,S- isomeric distributions ranging from 2% to 35% were examined. No additional

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Results and Discussion

page: 130

electrochemical responses were observed in any solvent for the 35% N-,S- bonded solutions and the measured Er1/2-values were also unaffected. Hence, it is likely that the reversible potential for the N-,S- isomer is similar (within 20mV) to that of the N-,N- isomer. However, in DMF, at increased (~>10% N-,S-) concentrations, a relative decrease in the magnitude of the second response was seen. That is, the oxidation peak at +0.74V is suppressed, as is the case when SS-bpy is present. In acetone, acetonitrile and THF the electrochemical behavior did not change significantly when the N-,S- isomer percentage was altered. 4.3.9. The Influence of Water The fully protonated form of (H2-dcbpy)2Ru(NCS)2 was found to be insoluble in slightly acidified water. At a neutral pH, a very small amount in the M concentration range dissolved. The dissolved material is assumed to be a deprotonated form of (H2-dcbpy)2Ru(NCS)2. The solubility of (H2-dcbpy)2Ru(NCS)2 in organic solvents may be significantly enhanced when small amounts of water are added. In DMF and DMSO, the solubility of (H2-dcbpy)2Ru(NCS)2 is substantial, but the rate of dissolution is slow in the absence of added water. In acetone the solubility is in the low millimolar region, while in THF and acetonitrile the solubility is in the sub millimolar range. Deliberate addition of water (10%Vol) to THF increased the solubility by a factor of about 50100. Since (H2-dcbpy)2Ru(NCS)2 as synthesized in this study contains two waters of solvent in its crystal lattice [29], water is always present in experiments reported in this chapter. Solvent THF Acetone DMF Methanol Conductivity a) / S mol1cm2 below LOQ b) 4.6 22.0 23.8

Table 4.3. Conductivity of (H2-dcbpy)2Ru(NCS)2 in different organic solvents. a) Determined at a concentration of 0.5mM b) Limit of Quantification.

To determine if any significant concentration of (H2-dcbpy)2Ru(NCS)2 was dissolved in a deprotonated or other ionized form, the conductivity of 0.5mM solutions were measured in a range of solvents (Table 4.3). In THF, the measured conductivity was indistinguishable from the solvent background value, while the molar conductivity of 4.6 Smol1cm2 in acetone is only slightly above the background value of the solvent. For comparison, the singly charged complex [Cr(en)2Cl2]+ [199] has a molar conductivity of 79 Smol1cm2, [Ru(bpy)2(NCS)(DMF)]+ a molar conductivity of 80 Smol1cm2 [200] and the electrolyte (Bu4NPF6) a molar conductivity of 103.4 Smol1cm2 in acetone [201]. Thus, the concentration of the deprotonated species in THF and acetone can be neglected. In contrast in DMF and methanol the conductivity was significantly higher (see Table 4.3), suggesting that up to 25% of (H2-dcbpy)2Ru(NCS)2 may have been dissolved in a deprotonated form in these higher dielectric constant solvents or alternatively, the negatively charged thiocyanate ligand may have been partially replaced by the solvent or small concentrations of water present.

Chapter 4: Studies on the Oxidation of cis-(H2-dcbpy)2Ru(NCS)2 Conclusions

page: 131

4.4. Conclusions
The reversible potential of the [(H2-dcbpy)2Ru(NCS)2]0/+ redox couple has been determined in a variety of solvents using different electrochemical techniques and electrode materials. Reactant and product interactions with the electrode surface cause departures from the mass transport controlled oxidation process for this compound under commonly used conditions of cyclic voltammetry at a macrodisk electrode. Surface effects were stronger on glassy carbon and gold than on platinum electrodes. Steady-state microdisk electrode and rotating disk electrode experiments provide close to chemically and electrochemically reversible mass transport controlled voltammograms in acetone and low concentrations of the dye and fast scan rates minimize the influence of the surface based effects under the transient conditions of cyclic voltammetry. In DMF, scan rates greater than 100Vs1 were needed to observe a significant level of chemical reversibility, whereas only moderate scan rates were required in other solvents. Addition of the electroinactive surfactant, SSbpy, was also found to minimize surface based effects. These surface based processes also restrict assumptions made from cyclic voltammetric data regarding the reversibility of the [(H2dcbpy)2Ru(NCS)2]0/+ process and stability of the oxidized form of the dye, which had previously led to confusion [29]. The chemical reversibility of the [(H2-dcbpy)2Ru(NCS)2]0/+ process was also found to depend strongly on the solvent used. As a worst case scenario, the reversibility of the oxidation process might be estimated in DMF from cyclic voltammetric data under conditions where the process is diffusion controlled (e.g. when SS-bpy is present) by assuming that an ipox/ipred-value of 0.8 is obtained at a scan rate of 100Vs1 and |EswitchEr1/2|=0.5V (see Figure 4.12). Nicholson and Shain [114] have set up a working curve, which allows estimation of kinetic parameters for a simple irreversible homogeneous chemical reaction following the charge transfer. This leads to a value for the half-lifetime of [(H2-dcbpy)2Ru(NCS)2]+ in DMF of 3ms. This value is several orders of magnitude less than reported in the literature [29] and assuming that the reduction back to the ground state occurs on the time scale of 1s [75] and not on the time scale of 108s [29], would possibly reduce the postulated life expectation of a dye sensitized solar cell device from "at least 12 years of continuous operation" [202] to a few weeks or even days. Of course, one has to bear in mind that the situation in the actual solar cell, where the dye is immobilized on the TiO2 surface, might be considerably different to results obtained from solution phase studies. In fact, long term stability tests [81] of dye sensitized solar cells have shown promising results. The substantial surface effects observed in this work where conventional electrode surfaces were employed give rise to another interesting question. The extremely high efficiency of photovoltaic sensitizers is partly attributed to the very effective attachment to the semi-conductor surface. Usually, it is postulated that the dyes are anchored by forming ester linkages with the titania electrode surface [67-69]. The results of this work suggest that a component of this attachment may be due to physical attachment and not solely the result of chemical attachment.

Chapter 5

part of this work has been published in

'Experimental and theoretical investigations of the effect of deprotonation on electronic spectra and reversible potentials of photovoltaic sensitizers: Deprotonation of cis-L2RuX2 (L = 2,2'-bipyridine-4,4'-dicarboxylic acid, X = CN, NCS) by electrochemical reduction at platinum electrodes.' Georg Wolfbauer, Alan M. Bond, Glen B. Deacon, Douglas R. MacFarlane, Leone Spiccia, J. Am. Chem. Soc., 1999, submitted 'Determination of Reversible Redox Potentials of Highly Surface Active Polypyridine Ruthenium Complexes' G. Wolfbauer, A.M. Bond, J.A. Howitt, D.R. MacFarlane, Abstracts of RACI Inorganic Chemistry Division Conference, 1998, Wollongong, Australia, poster PT85

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Introduction

page: 133

5.1. Introduction
The thermodynamically important oxidation process of the fully protonated photovoltaic sensitizer cis-(H2-dcbpy)2Ru(NCS)2 (H2-dcbpy = L = 2,2'-bipyridine-4,4'-dicarboxylic acid) has been investigated in depth in the previous chapter. From conductivity measurements (see previous chapter page 130) it has been concluded, that (H2-dcbpy)2Ru(NCS)2 dissolved in organic solvents is essentially present in its fully protonated form. However, the carboxylic acid groups of the dcbpy ligand, in addition to facilitating attachment to electrodes, also enable the L2RuX2 sensitizers to participate in acid-base reactions, which could significantly alter the electronic properties of the dcbpy ligand. Thus, whilst deprotonated forms of the sensitizer have been reported [203] to provide improved performance in photoelectrochemical cells by increasing the open circuit potential by 50100mV [203], surprisingly few studies [68] have specifically addressed the significance of the deprotonated forms of the sensitizer in either the thermodynamic or spectral sense. The absorbance of this class of sensitizers is dominated by metal-to-ligand charge-transfer bands (MLCT) which can formally be written as L2RuII(NCS)2 + h [(L)(L)RuIII(NCS)2]*. The values for the reduction processes of L2Ru(NCS)2 are also of importance, since the spectral response (absorbance) is determined by the difference of the oxidation and reduction potentials. Results obtained from the ester analogues (see Chapter 4) show that the carboxylate groups provide the bipyridine ligand with distinctly different physical and chemical properties to those of the unsubstituted bipyridine ligand. Therefore, ruthenium bipyridine complexes may not be taken as model compounds for complexes of Ru and H2-dcbpy. No studies of the reduction processes of [Ru(H2-dcbpy)3]2+ or derivatives are available in the literature. The results presented in Chapter 3 indicate that the ester analogues are ideal model compounds of the corresponding acids. From this data, it is anticipated that the (H2-dcbpy)2Ru(NCS)2 and (H2-dcbpy)2Ru(CN)2 sensitizers may exhibit reversible reductive electrochemistry. However, the H2-dcbpy ligand may participate in either acid-base reactions or deprotonation of the H2-dcbpy ligand, which is deliberately performed when the sensitizer is applied in real solar cell applications [203]. Both reactions will significantly alter the electronic structure of the ligand and complex and these effects cannot be deduced from the ester analogues. This study has explored the use of electrochemical reduction of the fully protonated cis-(H2dcbpy)2RuX2 (X = NCS, CN, for structure see Figure 5.1) complexes in dimethylformamide (and generation of hydrogen gas) as a means of systematically achieving deprotonation of the photovoltaic sensitizer. Spectroscopic (UV/VIS) and voltammetric studies on the reduction and oxidation of the electrochemically generated deprotonated species enable insights to be gained into changes of molecular orbital energy levels that occur with different levels of protonation. The experimentally derived results, combined with theoretical data provided by molecular orbital calculations, enhance

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Introduction

page: 134

the detailed understanding of the role of the protonation state in the performance of sensitizers used in photoelectrochemical cells.

A
e

S+/S*
fermi level

COOH HOOC N X X X = NCS CN COOH N Ru N COOH

e n e rg y

Eoc R/R

TiO2

electrolyte solution

Figure 5.1. Simplified diagram showing the electron transitions in a photoelectrochemical solar cell during operation. S .... sensitizer (for structure see right), R .... electrolyte (commonly I /I3), Eoc .... open circuit potential.

5.2. Experimental
5.2.1. Instrumentation and Electrodes for Voltammetry Voltammetric measurements were carried out using a standard three-electrode arrangement. Specific details can be found in Chapter 3.2 on page 62. 5.2.2. Instrumentation for other Techniques The setup for bulk electrolysis was identical to that described in Chapter 3.2 on page 62. Additionally, bulk electrolysis at glassy carbon electrodes utilized an identical setup as described above and a cylindrical single piece glassy carbon working electrode. For gas chromatographic determination of hydrogen gas generated during electrolysis, the lid was sealed by paraffin film and the inlets by rubber septa. Gases evolved during bulk electrolysis were analyzed by taking a sample of the supernatant gas phase (500L) and the composition was detected by gas chromatography. For gas chromatographic measurements, a Varian 3700 gas chromatograph equipped with a carbosphere 80/100 9'x''s.s. column and a TCD detector was used. Under conditions where Tcol=40C, Tvalve=100C, TTCD=200C and flow rate = 27mLmin1 (He), dihydrogen had a retention time of 532s. The MOPAC-97 software package (Fujitsu Ltd., Tokyo, Japan) was used for molecular orbital calculations. All calculations employed the semiempirical MNDO-PM3 Hamiltonian with the default parameters supplied. "Sparkles" were used to balance negative charges. The setup for electrochemical quartz crystal microbalance (EQCM) studies is described in Chapter 4.2.1 on page 113 and that for optically transparent thin-layer electrolysis (OTTLE) can be found in Chapter 3.2 on page 62.

counter electrode

S+/S

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Experimental

page: 135

5.2.3. Reagents and Synthesis of Compounds Solvents and electrolyte used for electrochemical and spectroscopic measurements were as described in Chapter 3.2 on page 62. An aqueous 40% tetrabutylammonium hydroxide, Bu4N(OH), solution and 48% aqueous HBF4 were purchased from Aldrich and used as received. cis-(H2dcbpy)2Ru(NCS)2 and cis-(H2-dcbpy)2Ru(CN)2 were synthesized as described in the previous two chapters and the cis prefix notation will be omitted in further discussion. (H2-dcbpy)2Ru(NCS)2 was present predominately as the diisothiocyanate linkage isomer with 4% of the mixed bond N,S- linkage isomer. Tetrabutylammonium "salts" of the sensitizers [68, 203] used for reference against deprotonated solutions prepared by electrochemical reduction of (H2-dcbpy)2RuX2 were synthesized via two different routes [68]. Firstly by precipitation of solid after acidification of aqueous Bu 4N(OH) solutions of (H2-dcbpy)2RuX2 by titration with a 5% CF3COOH aqueous solution over the pH range of 24.5. The "salts" formed via this procedure were separated from the solution by centrifugation and then freeze dried. This method produced salts of composition [(H2x/2dcbpyx/2)2RuX2](Bu4N)x, where 0.80x1.85 (x was determined by NMR spectroscopy). A second synthetic approach was utilized to prepare salts where x 2. In this method, solid (H2-dcbpy)2RuX2 was directly mixed with the required amount of aqueous Bu4N(OH) solution and then freeze dried under high vacuum to remove excess water. (H2-dcbpy)2RuX2 itself contains two to four waters of crystallization [29] and hence some water was always present in the parent compound as well as in the deprotonated salts.

5.3. Results and Discussion


5.3.1. Electrochemical Reduction of (H2-dcbpy)2Ru(NCS)2 in DMF (a) Voltammetry in DMF: The Initial Two One-electron Reduction Processes In order to establish a reference point against which to measure changes induced by deprotonation, detailed knowledge in the voltammetric behavior of (H2-dcbpy)2Ru(NCS)2 is required (for oxidation see Chapter 4, for reduction see this Chapter). The reduction of [Ru(bpy)3]2+ (bpy = 2,2'bipyridine) and related compounds usually occurs via an extensive series of reversible one-electron ligand based charge transfer processes [142, 150]. For example, the ester analogue of (H2dcbpy)2Ru(NCS)2 (see Chapter 3.5.3 on page 99) exhibits two chemically and electrochemically reversible reduction processes in DMF with half-wave potentials (Er1/2) of 1.48V and 1.70V vs. Fc/Fc+ at room temperature at glassy carbon and platinum electrodes which may be extended to four reversible processes at T=58C (Er1/2=1.49V, 1.67V, 2.16V and 2.43V). Since the electronic effect arising from esterification is expected to be negligible, it would be expected that fully protonated (H2-dcbpy)2Ru(NCS)2 should show analogous voltammetric reduction behavior.

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 136

1A

a)

b)
1A

c)
1nA

d)
2A

-2.0

-1.5

-1.0
+

-0.5

potential [V] vs. Fc/Fc

Figure 5.2. Voltammograms for the first and second (H2-dcbpy)2Ru(NCS)2 reduction process at T=25C in DMF (0.2M Bu4NPF6). a and b) cyclic voltammetry (=100mVs1) at a glassy carbon disk electrode (d=1mm) c) steady-state response (=20mVs1) at a glassy carbon microdisk electrode (d=11.2m) and d) cyclic voltammetry (=100mVs1) at a platinum disk electrode d=1mm.

The cyclic voltammetric response observed at a glassy carbon electrode (Figure 5.2a) in DMF over the potential range of 0.5V to 1.9V vs. Fc/Fc+ consists of two well defined processes, the first being chemically reversible as expected (Figure 5.2b) and the second process, whilst only partially reversible in the chemical sense, still occurs in the expected potential region. If these initial two reductions processes are H2-dcbpy ligand based, as is known to be the case with the ester analogue, the initial charge transfer processes may be formulated as in Equations 5.1 and 5.2 below: (H2-dcbpy)2Ru(NCS)2 + e [(H2-dcbpy)(H2-dcbpy)Ru(NCS)2] + e [(H2-dcbpy)2Ru(NCS)2]2 (5.1) (5.2)

[(H2-dcbpy)(H2-dcbpy)Ru(NCS)2]

where H2-dcbpy stands for the singly reduced fully protonated H2-dcbpy ligand. Further complex reduction processes at more negative potentials are present at glassy carbon electrodes (see later). Evidence that the first process (Eqn. 5.1) is reversible in both the chemical and electrochemical senses in DMF (0.2M Bu4NPF6) at a glassy carbon electrode is the fact that the peak-topeak separation (Ep) for the reduction (Epred) and oxidation (Epox) peak potentials was 655mV over the scan rate range of 20500mVs1 (concentration c0=1.1mM, temperature T=25C), which is close to the theoretically predicted value of 57mV for a reversible one-electron process. Values of Ep as a function of scan rate are contained in Table 5.1.

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 137

Cyclic Voltammetrya)

Rotating Disk Electrodeb)

[(H2-dcbpy)2Ru(NCS)2]0/
/ mVs1 10 26 50 100 200 500 1000 2000 5000 Epox / mV 1560 1557 1555 1555 1557 1559 1565 1577 1601 Epred / mV Ep / mV 1489 1497 1493 1493 1491 1489 1487 1479 1453 71 60 62 62 66 70 78 98 145 Er1/2 / mV |ipred/ipox| f / min1 1525 1527 1524 1524 1524 1524 1526 1528 1527 1.34 1.08 1.07 1.08 1.10 1.08 1.04 1.08 1.07 500 1000 1500 2000 2500 3000 slope / mV Er1/2 / mV

66 1522 67 1525 65 1527 64 1527 62 1530 64 1529 Microdisk Electrode d / m slope / mV Er1/2 / mV 11.2 59 1525

[(H2-dcbpy)2Ru(NCS)2]/2
/ mVs1 1000 1500 2000 3000 4000 5000 Epox / mV 1785 1791 1795 1801 1809 1817 Epred / mV Ep / mV 1665 1679 1683 1675 1667 126 116 118 134 150 Er1/2 / mV |ipred/ipox| f / min1 1728 1737 1742 1742 1742 1.45 1.26 1.19 1.12 1.09 500 1000 1500 2000 2500 3000 slope / mV Er1/2 / mV

66 1747 66 1751 66 1752 65 1753 64 1754 Microdisk Electrode d / m slope / mV Er1/2 / mV 11.2 57 1742

[(H2x/2-dcbpyx/2)2Ru(NCS)2]x/(x+1)
/ mVs1 20 50 100 200 500 1000 2000 Epox / mV 130 126 126 134 138 140 154 Epred / mV Ep / mV 40 56 56 52 52 48 34 90 70 70 82 86 92 120 Er1/2 / mV |ipred/ipox| f / min1 87 93 93 95 95 96 96 2.40 1.60 1.44 1.55 1.54 1.42 500 1000 1500 2000 2500 3000 3000 slope / mV 85 84 83 88 91 89 89 Er1/2 / mV 83 88 89 92 93 92 92

Table 5.1. Voltammetric data obtained for the reduction (T=25C) of 1.1mM (H2-dcbpy)2Ru(NCS)2 in DMF (0.2 M Bu4NPF6) and oxidation of deprotonated [(H2x/2-dcbpyx/2)2Ru(NCS)2]x at glassy carbon electrodes a) d=1mm and b) d=3mm as a function of scan rate and rotation rate. Peak potentials are reported versus Fc/Fc+ with an uncertainty of 2mV. Cyclic voltammetric Er1/2-values calculated as (Epox + Epred)/2, for steady state techniques slopes and Er1/2 calculated from "log-plots". ( = scan rate, f = rotation frequency, d = electrode diameter, Ep = peak potential, Ep = Epox Epred , ip = peak current).

At the glassy carbon electrode, the peak current ratio ipred/ipox (ipox = oxidation peak current, ipred = reduction peak current) was also close to the theoretically expected value of 1.0 over the scan rate range of 205000mVs1 (see Table 5.1). Furthermore, the reversible half-wave potential or Er1/2value (see Table 5.1) calculated from the values of (Epox + Epred)/2 (cyclic voltammetry) and the potential at iL/2 (rotating disk electrode where iL is the limiting current) is independent of scan rate and rotation rate, respectively, as required for a reversible process. The measured value of Er1/2=15253mV is similar to Er1/2=1480mV listed in Table 3.8 on page 101 for the ethyl ester. A plot of ipox versus 1/2 from cyclic voltammograms ( = scan rate) is linear over the scan rate of 10mV to 5000mVs1 confirming that the process is diffusion controlled. Mass transport control

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 138

(diffusion and convection) was also demonstrated via the linear dependence of iL on 1/2 ( = 2f = angular velocity) in rotating disk electrode experiments. A voltammogram for the initial process obtained under near steady-state conditions at a glassy carbon microdisk electrode (Figure 5.2c) was well defined and the calculated Er1/2-value coincided with that obtained from cyclic and rotated electrode techniques (see Table 5.1). The essentially ideal reversible one-electron process observed at a glassy carbon microdisk electrode, compared to minor deviations encountered in other techniques under some conditions (cyclic voltammetry at high scan rates and rotating disk electrode at high rotation rates) suggests that a small amount of uncompensated resistance is present when techniques based on the use of macrodisk electrodes are used. Confirmation of an initial one-electron charge transfer process was obtained by noting that "log plots" were linear and had values close to 59mV at 25C at rotating macrodisk and microdisk glassy carbon electrodes, see Table 5.1. Data obtained from rotating disk electrode measurements and use of the Levich equation (Equation 2.29 on page 49) enabled a value for the diffusion coefficient of D=2.8(0.2)106cm2s1 to be calculated for the protonated form of the (H2dcbpy)2Ru(NCS)2 complex in DMF. The observation of chemical reversibility of the second reduction process (Equation 5.2) at a glassy carbon electrode requires the use of cyclic voltammetric scan rates greater than 5000mVs1. Under these conditions, the ipred/ipox-ratio for this process was close to unity (see Table 5.1). Under near steady-state conditions with glassy carbon microdisk and rotating disk electrodes (rotation rates >1000rpm), the second process showed reversible behavior. Er1/2-values obtained from the latter two techniques coincided with (Epox + Epred)/2 values obtained from cyclic voltammograms at scan rates >1500mVs1. "Log-plot" analysis of the second process gave slopes close to the theoretical value expected for a reversible one-electron process. At 58C and at glassy carbon electrodes, the initial two (H2-dcbpy)2Ru(NCS)2 reduction processes are both chemically and electrochemically reversible even under longer time scale conditions (Figure 5.3a and b). Thus, under conditions of cyclic voltammetry, ipred for both processes scaled linearly with 1/2 and ipred/ipox-values are close to unity over the entire scan rate range (20 to 5000mVs1). Furthermore, "log-plots" from rotating disk electrode voltammograms recorded at this temperature gave slopes of 486mV for both processes, which are close to the theoretically expected value of 43mV at this temperature. The Levich plot (rotating disk electrode) was linear and passed through the origin for both processes. Er1/2-values and other data obtained from glassy carbon electrode voltammograms at 58C are reported in Table 5.2 on page 140 for both processes.

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 139

a)
0.5A

b)
2A

c)
-2.0 -1.5 -1.0

0.5A

-0.5
+

potential [V] vs. Fc/Fc

Figure 5.3. Voltammograms for the first two (H2-dcbpy)2Ru(NCS)2 reduction processes at T=58C in DMF (0.1M Bu4NPF6). a) cyclic voltammetry (=100mVs1) at a glassy carbon disk electrode (d=1mm), b) steadystate voltammogram at a glassy carbon rotating disk electrode (500rpm), c) cyclic voltammetry (=100mVs1) at a platinum disk electrode (d=1mm).

Even casual inspection of a cyclic voltammogram obtained for reduction of (H2dcbpy)2Ru(NCS)2 in DMF (0.1M Bu4NPF6) at a 1mm diameter platinum disk electrode (Figure 5.3d) shows that great differences are found relative to that at glassy carbon with the initial reduction processes for (H2-dcbpy)2Ru(NCS)2 are both very drawn out relative to the response expected for a reversible process, and even more significantly the peak potential of the first process (Epred = 1.20V vs. Fc/Fc+) is about 0.3V less negative than predicted on the basis of data obtained with the ester analogue. Even at 58C, cyclic voltammograms obtained at a platinum electrode for the reduction of (H2-dcbpy)2Ru(NCS)2 remain irreversible and highly complex (Figure 5.3c), although the reduction peak potentials are now in the same region as at glassy carbon electrodes. This platinum electrode dependence differs to the voltammetric behavior found for the ester analogue (see chapter 3.4) and [Ru(bpy)3]2+ [150] where results are independent of electrode material. Thus, in summary, the voltammetry of (H2-dcbpy)2Ru(NCS)2 at platinum electrodes contains unexpected, but as it will emerge, very useful characteristics that provide distinctively different voltammetry to that at glassy carbon where the first two processes exhibit essentially the behavior expected for this class of compound.

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 140

Electrochemical data a) complex RuII/III oxidation Er1/2 [mV] (H2-dcbpy)2Ru(NCS)2 [(H2x/2-dcbpy [(H2x/2-dcbpy
x/2

Spectroscopic data b) L L* MLCT MdL2* MLCT MdL1*

L/L reduction Er1/2 [mV] 15254 15493 d) 21545 162410 16345 d) 22675

L/L2 reduction Er1/2 [mV] 17425 17244 d) 250010 180810


d)

energy [103cm1] (molar extinction [103M1cm1]) 31.5 (51.6) 32.5 (45.7) 31.7 (46.8) 32.7 (42.4) 24.8 (15.1) 26.7 (14.7) 25.2 (14.3) 27.5 (13.1) 18.3 (15.3) 19.4 (14.1) 18.7 (17.1) 19.7 (14.9)

+39010 c)
x

)2Ru(NCS)2] )2Ru(CN)2]

+935 +5685

(H2-dcbpy)2Ru(CN)2
x/2 x

+2105

26178

Table 5.2. Summary of electronic spectral data and reversible half-wave potentials for (H2-dcbpy)2RuX2 and x/2 the deprotonated form ([(H2x/2-dcbpy )2RuX2]x). a) Electrochemical data obtained from voltammetry at glassy carbon electrodes and potentials are referenced against Fc/Fc+, b) data obtained by in-situ electrolysis in an OTTLE (platinum gauze) experiment, c) value taken from Table 4.2 on page 125, d) T=58C, other data obtained at 25C.

(b) Bulk Reductive Electrolysis at Platinum Electrodes in DMF The origin of the complex voltammograms at the platinum electrode was probed by exhaustive bulk electrolysis experiments under controlled potential conditions at a large area platinum gauze working electrode at 25C. In initial experiments, the potential was set to 2.0V vs. Fc/Fc+ (more negative than the second (H2-dcbpy)2Ru(NCS)2 process detected at glassy carbon and platinum electrodes) and the electrolysis of 0.52.0mM solutions allowed to go to completion (3040) minutes. Coulometric analysis under these conditions revealed that 3.350.35 electrons per molecule were involved in the complete reaction. Gas chromatographic measurements of the collected gas phase showed molecular hydrogen as being the significant gaseous product. The peak intensity of the hydrogen signal was close to that expected if the majority of electrons transferred were used for the formation of hydrogen. On the basis of the coulometry and hydrogen data, the overall reaction observed on the time scale of bulk electrolysis at 2.0V vs. Fc/Fc+ at a platinum electrode can be said for initial discussion purpose to correspond approximately to an overall deprotonation reaction of the kind (H2-dcbpy)2Ru(NCS)2 + 4e [(dcbpy2)2Ru(NCS)2]4 + 2H2 (5.3)

For the sake of clarity, dcbpy2 and H-dcbpy will be used throughout the remainder of this chapter for the doubly and singly deprotonated H2-dcbpy ligand, respectively. In contrast, H2dcbpy will be used for the singly reduced but fully protonated H2-dcbpy ligand. Mixed forms of the ligand (deprotonated and reduced) will be abbreviated accordingly. The overall charge of the complexes is given outside the square brackets. Clearly, an extensive sequence of electron transfer and chemical reactions can occur at the platinum surface which involves steps where H+ is formed by deprotonation before or after an electron transfer reaction and H2 is generated heterogeneously at the electrode surface or homogeneously by solution phase reactions. One example of heterogeneous generation of H2 that gives the overall reaction in Equation 5.3 would be the reaction sequence in Equations 5.4 to 5.6:

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 141

(H2-dcbpy)2Ru(NCS)2 [(H-dcbpy)2Ru(NCS)2]2

[(H-dcbpy)2Ru(NCS)2]2 [(dcbpy2)2Ru(NCS)2]4 2H2

+ 2H+ + 2H+

(5.4) (5.5) (5.6)

4H+ + 4e

or if hydrogen is generated by homogeneous reactions after electron transfer, then Equations 5.7 to 5.9 would represent one of many examples of this kind of reaction pathway that give the overall Equation 5.3. (H2-dcbpy)2Ru(NCS)2 [(dcbpy2)2Ru(NCS)2]4 + 4H+ (5.7) (5.8) 2H2 (5.9)

2(H2-dcbpy)2Ru(NCS)2 + 4e 2[(H2-dcbpy)2Ru(NCS)2]2 + 4H+

2[(H2-dcbpy)2Ru(NCS)2]2 +

2(H2-dcbpy)2Ru(NCS)2

These sequences of reactions are possible because the protonated (H2-dcbpy)2Ru(NCS)2 sensitizer may exist in an acid-base equilibrium (Eqns. 5.4 and 5.7), the extent to which deprotonation occurs being dependent on solvent (medium) specific pKa-values. Thus, it needs to be noted that in the presence of water, adventitiously present or deliberately added, thermodynamic (and/or kinetic) limitations will be placed on the extent to which reduction can occur via the equilibrium prior of the reaction [(dcbpy2)2Ru(NCS)2]4 + y H2O [(Hy/2-dcbpy(2y/2))2Ru(NCS)2](4y) + y OH (5.10)

Thermodynamically, protons can very easily be reduced to hydrogen, although the suppression of the heterogeneous reaction pathway can be achieved by choice of an appropriate electrode material at which the rate of this reaction is electrochemically negligible at the reversible potential [95]. Glassy carbon is an electrode material where the over-potential for the reduction of protons to hydrogen is very large. In contrast, the over-potential for the generation of hydrogen at platinum electrodes is small. Reduced ruthenium polypyridine compounds are known to catalytically reduce protons or residual water to hydrogen in organic solvents [55]. Thus, the reaction sequence described in Equations 5.8 and 5.9 might be a source of hydrogen generation, at least under long time scale conditions. However, the hydrogen producing step in Eqn. 5.9 is based on a purely homogeneous solution phase reaction and does not predict the experimentally observed influence of electrode material. In contrast, the electrochemical reduction of protons (Eqn. 5.6), made available by the acid-base equilibrium described in Eqns. 5.4, 5.5 and 5.7 occurs at the electrode surface and hence heterogeneous electrode kinetics may be involved in the reaction sequence which could explain the electrode material dependent voltammetric responses observed for the reduction of (H2-dcbpy)2Ru(NCS)2. Overall Equation 5.3 predicts the transfer of 4.0 electrons per molecule if complete deprotonation occurs to quantitatively yield [(dcbpy2)2Ru(NCS)2]4, whereas the number of electrons transferred was only 3.350.35. Hence, this equation and the mechanism given as examples cannot

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 142

provide a complete description of the process. The pKa values of (H2-dcbpy)2Ru(NCS)2 are unknown, although (H2-dcbpy)2Ru(NCS)2 can only be a weak acid since conductivity measurements have shown (see Chapter 4.3.9 and Table 4.3 on page 130) that the complex is present predominantly in its protonated form. However, the first two pKa-values for the (H2dcbpy)2Ru(NCS)2 complex are expected to be relatively low in comparison to those reported for [Ru(H2-dcbpy)2(bpy)]2+ [204] so that at least two protons are likely to be relatively easy to remove. Finally, it needs to be noted that since residual water is present, the reaction sequence (H2-dcbpy)2Ru(NCS)2 + 2 e [(H2-dcbpy)2Ru(NCS)2]2 + 2 H2O (H2-dcbpy)2Ru(NCS)2 + 2 OH would give the overall reaction (H2-dcbpy)2Ru(NCS)2 + 2 H2O + 4 e [(H-dcbpy)2Ru(NCS)2]2 + 2 OH In summary, this reaction with water can be generally written as (H2-dcbpy)2Ru(NCS)2 + (x2)H2O + xe [(H-dcbpy)2Ru(NCS)2]2 + (x2)OH +
x 2

[(H2-dcbpy)2Ru(NCS)2]2

(5.11) (5.12) (5.13)

(H2-dcbpy)2Ru(NCS)2 + 2 OH + H2 [(H-dcbpy)2Ru(NCS)2]2 + H2O

+ 2 H2

(5.14)

H2

(5.15)

This reaction sequence is also consistent with data (vide infra) obtained in this study. Clearly, a range of combination of electron and proton transfer steps are possible (see above) which lead to deprotonation. However, the reaction scheme can be very complex and formation of singly and triply deprotonated (H2-dcbpy)2Ru(NCS)2 complexes also can occur as can potential dependent combinations of reaction pathways. In order to probe the identity of the deprotonated species produced during electrolysis a 1.1mM solution of (H2-dcbpy)2Ru(NCS)2 was reduced in DMF at a platinum gauze electrode using the same conditions as before. However, on this occasion, the course of the reaction was monitored (Figure 5.4) by cyclic voltammetry at a glassy carbon electrode over a more negative potential regime where the existence of an extended series of processes is revealed (Figure 5.4a). When voltammetric monitoring is undertaken over this wider potential range (Figure 5.4a-e), the initial two reduction processes observed above, as well as the other processes in the more negative potential region (Figure 5.4a) ultimately are replaced by two new well defined processes in the very negative potential region (Figure 5.4e). As may be expected, cyclic voltammograms obtained on partially rather than exhaustively reduced solutions (Figure 5.4b, c and d) are complicated. The particular exhaustive bulk electrolysis shown in Figure 5.4 consumed 3.5 electrons per molecule and leads to detection of one chemically and electrochemically reversible process, followed by a partially reversible process at the glassy carbon electrode. However, despite the complexity of the

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 143

initial steps of electrolysis the final form of glassy carbon cyclic voltammogram detected from solutions exhaustively reduced at platinum electrodes at 2.0V simply consists of two processes with Er1/2-values that are about 0.65V more negative than the initial two processes for reduction of (H2-dcbpy)2Ru(NCS)2.

+0.7e

e)

d)
+1.0e
_

5A

c)
+1.0e
_

b)
+1.0e
_

a)

-3.0

-2.5

-2.0

-1.5

-1.0
+

-0.5

potential [V] vs. Fc/Fc

Figure 5.4. Cyclic voltammograms (=1000mVs1, initial five scans, first scan bold) obtained at a glassy carbon working electrode (d=1mm) for reduction of 1.1mM (H2-dcbpy)2Ru(NCS)2 in DMF (0.1M Bu4NPF6) at different stages of reductive bulk electrolysis (Eappl=2.05V) at a platinum gauze electrode. a) before electrolysis commences, b) one electron transferred, c) two electrons transferred, d) three electrons transferred, e) 3.7 electrons transferred, (exhaustive electrolysis).

Detailed analysis of voltammograms at glassy carbon electrodes of the two reduction processes observed after exhaustive electrolysis at Eappl= 2.0V vs. Fc/Fc+ reveals that ipred for the first process (cyclic voltammetry) scaled linearly with 1/2. Furthermore, Levich plots (rotated glassy carbon disk electrode) were linear and passed through the origin for both processes, establishing that they are mass transport controlled. The slope of the Levich plot gave a diffusion coefficient for [(H2-x/2-dcbpyx/2)2Ru(NCS)2]x (x = number of protons removed) of D=2.50.4106cm2s1, which is very similar to the value obtained for the fully protonated form. The Er1/2-value for the initial reduction process (cyclic and rotating disk techniques) was calculated to be 21555mV vs. Fc/Fc+. "Log-plots" (rotating disk voltammograms) gave slopes of 702mV, which are close to the

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 144

theoretical value expected for a reversible one electron process. Scan rates >2000mVs 1 were required to make the second reduction process fully reversible. Under steady-state conditions at the rotating disk electrode, the second process is also close to reversible. Thus, Er1/2-values calculated from "log-plots" for the second process (rotated disk) are in agreement with the value of (Epox + Epred)/2 obtained from cyclic voltammograms (>200mVs1). The reversible half-wave potential for the second reduction process of [(H2-x/2-dcbpyx/2)2Ru(NCS)2]x therefore is established to be 250010mV vs. Fc/Fc+. In another series of bulk electrolysis experiments at the platinum gauze electrode, the solution of electrogenerated [(H2-x/2-dcbpyx/2)2Ru(NCS)2]x was further reduced at a potential of Eappl= 2.3V (more negative than the reversible potential of the first new reduction process for the deprotonated form of (H2-dcbpy)2Ru(NCS)2. Monitoring the course of this experiment by glassy carbon rotating disk electrode measurements indicated that conversion of the desired product back to the originally present [(H2-x/2-dcbpyx/2)2Ru(NCS)2]x form of the ruthenium complex occurred. Moreover, hydrogen was detected by gas chromatographic analysis of the gas phase present above the sample. This reaction pathway would be expected if [(H2-x/2-dcbpyx/2)2Ru(NCS)2](x+1) compound can reduce residual water or even the electrolyte or solvent to regenerate [(H 2-x/2dcbpyx/2)2Ru(NCS)2]x, which is the initial form of the complex [55] and hydrogen gas [65], in a similar form of reaction as postulated for reaction of reduced forms of (H2-dcbpy)2Ru(NCS)2 with water (Equation 5.9) [205]. Alternatively, an even less protonated form might be initially generated with evolution of hydrogen, but then react with residual water to reform the more protonated form. Prolonged reduction periods of >20min at 2.3V resulted in slow decomposition of [(H2-x/2dcbpyx/2)2Ru(NCS)2]x, as determined by voltammetric monitoring at glassy carbon electrodes. Thus, a slow decomposition reaction is available for this species as an alternative to the hydrogen evolution redox reactions considered above. In order to verify that deprotonation and hydrogen evolution could occur at potentials prior to the reversible potential established at glassy carbon electrodes for the [(H2-dcbpy)Ru(NCS)2]0/ process, solutions of (H2-dcbpy)2Ru(NCS)2 were reductively electrolyzed at platinum electrodes at a potential of 1.2V vs. Fc/Fc+. This potential in fact corresponds to the foot of the initial reduction wave observed at platinum electrodes (Figure 5.2d) and therefore is significantly less negative than the reversible potential for the first ligand based reduction process (Er1/2=1.53V, see Table 5.2). Exhaustive electrolysis at this potential required a longer time span of 90 to 180 minutes. Again, hydrogen gas was detected as a product of electrolysis. The number of electrons transferred at this potential was 2.50.5 and a cyclic voltammogram obtained after electrolysis was similar to those observed for partial electrolysis at 2.0V (Figure 5.4bc), implying that only partial deprotonation occurred under these conditions. When the electrolysis was allowed to continue at 2.0V a further 1.00.5 electrons were consumed, so that the sum of both forms of electrolysis give rise to the transfer of as many electrons as when exhaustive electrolysis was carried out at 2.0V in a single step. As expected, a cyclic voltammogram at a glassy carbon electrode of the solution electrolyzed

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 145

in these two stages was identical to that shown in Figure 5.4e, which indicates that the same deprotonated products are formed during electrolysis at positive potentials. Importantly, this last data set leads to the conclusion that reductive electrolysis and deprotonation of (H2dcbpy)2Ru(NCS)2 at platinum electrodes can proceed via the heterogeneous reduction at 1.2V of protons at a platinum electrode surface, since the thermodynamic requirement for the homogeneous formation of hydrogen via ligand reduced forms of (H2-dcbpy)2Ru(NCS)2 (Eqns. 5.75.9) is not met at this potential. (c) Bulk Electrolysis at Glassy Carbon Electrodes When bulk reductive electrolysis of (H2-dcbpy)2Ru(NCS)2 was attempted at 1.2V vs. Fc/Fc+ at a glassy carbon cup electrode, no appreciable reaction could be detected. This result is as expected from cyclic voltammetric data obtained at this electrode material, and also confirms that the heterogeneous reaction pathway for hydrogen production at this electrode material is negligible. However, when the electrolysis was carried out at 2.0V vs. Fc/Fc+, exhaustive electrolysis proceeded as observed when platinum electrodes were used, although the reaction needed longer for completion. Cyclic voltammetric and gas chromatographic monitoring of this reaction, carried out as described above after electrolysis at a platinum basket electrode, showed that the reaction occurred in the same overall manner as that observed at platinum electrodes. That is, deprotonated forms of the (H2-dcbpy)2Ru(NCS)2 complex and hydrogen gas were formed. In contrast to platinum electrodes, at this electrode material hydrogen formation predominately will have occurred via the homogeneous pathway described in Equations 5.75.9. It is noted that after electrolysis minor decomposition products were detected by cyclic voltammetry on glassy carbon electrodes suggesting that additional side reaction pathways are available for the ligand based reduced forms of (H2-dcbpy)2Ru(NCS)2. However, contributions from these side reactions were very small and could only be noted after excessive acidification of the electrolyzed solution, when cyclic voltammograms over the positive potential range of 0.5V to 1.0V (compare Figure 5.6a on page 149) showed slightly different responses after the initial metal centered oxidation process in the potential range of 0.5V to 1.1V. The responses following the initial oxidation process have been assigned to surface based processes in Chapter 4. (d) Molecular Orbital Calculations The 0.65V shift in potential of Er1/2 for the ligand based reduction processes of (H2dcbpy)2Ru(NCS)2 after deprotonation may be attributed to the electronic influence of the carboxylate group on the bpy ligand. In ruthenium polypyridine complexes, the first set of reduction processes are ligand based and are directly related to the energy of the first unoccupied orbital (LUMO or 1*) of the polypyridine ligand [62, 144]. Due to its M (M = mesomeric effect) effect, the protonated carboxylate ligand (-COOH) withdraws electron density from the bpy ring and hence the energy of the LUMO is significantly lowered by about 0.5eV according to MNDO calculations (see below and ref [144]). On the basis of this electronic effect, (R2-dcbpy)2RuX2 (R = C2H5 in Chapter 3 and R = H in this Chapter) should both be easier to reduce than (bpy)2RuX2 [62],

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 146

as is confirmed when data available in literature are examined. Thus, Er1/2 for the first ligand based reduction process of (bpy)2RuX2 (X = halogen or pseudo halogen) compounds lies in the range of 1.95V 2.1V vs. Fc/Fc+, whereas the initial reduction process (R2-dcbpy)2RuX2 occurs between 1.48V and 1.62V vs. Fc/Fc+. However, if the -COOH group is deprotonated, to give -COO, the influence of the electronic effect on redox potentials would be expected to be reversed. That is, a significant increase in LUMO energy is expected to occur upon deprotonation which in turn would be consistent with the observed shift of the reduction processes to more negative potentials. In an endeavor to quantify the effect of deprotonation of the H2-dcbpy ligand, semi-empirical molecular orbital (MO) calculations were undertaken on H2-dcbpy, its deprotonated forms and the bpy ligand. The MO approach has been employed [144] to explain the comparative features of electrochemistry of [Ru(Et2-dcbpy)3]2+ and [Ru(bpy)3]2+, where Et2-dcbpy is the diethlyester of H2dcbpy. Importantly, it is noted that the authors of this particular study [144] could predict the potential of the first ligand based reduction of [Ru(Et2-dcbpy)3]2+ by comparing the LUMO energies for the bpy and H2-dcbpy ligands (H2-dcbpy was used in this calculations as a model ligand for Et2-dcbpy). Prior to undertaking the MO calculations, the ligands of interest were subjected to energy minimization calculations. All bipyridine ligands exhibited an absolute energy minimum with the pyridine nitrogens in a trans position relative to each other. However, in order to maintain a configuration which is more closely related to the manner in which the bipyridine ligands are known [38, 83] to be coordinated to the ruthenium center (see insert in Figure 5.1), the ring systems were minimized into a local energy minimum in which the substituents on both pyridine ring systems are in cis positions relative to each other, although the difference in energy between the trans and cis configurations is very small. In the cis configuration, the pyridine rings are slightly out-of-plane by a tilt angle of 112. In Chapter 3, this non-planarity of the bipyridine ligands has been used to rationalize ESR results obtained from the oxidized [(Et2-dcbpy)2RuCl2]+ complex (see Chapter 3.4.1 on page 72). Furthermore, the carboxylic acid groups (-COOH) were out-of-plane with the pyridine ring by 222, which is in good agreement with crystallographic data of (H2dcbpy)2Ru(NCS)2 where an angle of 30 has been reported [83], whereas the carboxylic anion groups (-COO) were in-plane with the pyridine ring. The calculated MO data are represented schematically in Figure 5.5. The 2*, 3* and 4* orbitals and absolute energy values are included for convenience. The introduction of carboxylic acid groups on the bpy ligand was calculated to lower the 1* orbital by 0.53eV, which is in good agreement with the value of 0.58eV reported by Ohsawa and co-workers [144]. The drastic effect of deprotonation on the energies of the molecular orbitals and hence the effect on the reduction potentials is clearly demonstrated in Figure 5.5.

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 147

1.5

orbital energy [eV]

1.0 0.5 0.0 -0.5 -1.0

* 4 * 3 * 2
* 1

LUMO
bpy

dcbpy

H-dcbpy

H2-dcbpy

Figure 5.5. Orbital energy values obtained from MNDO-PM3 calculations.

Removing one proton from H2-dcbpy increases the energy of the 1* orbital by 0.74eV, which is more then enough to overcome the initial 0.53eV lowering of the electron withdrawing effect caused by introduction of the carboxylic acid groups. Thus, the 1* energy of H-dcbpy lies beyond that of bpy. The 1* orbital of the fully deprotonated ligand, dcbpy2, lies 1.72eV higher in energy than that of H2-dcbpy. Energy density calculations show, that in case of the partly deprotonated ligand (H-dcbpy) the 1* and 2* orbitals are predominantly located on the pyridine moiety containing the protonated acid. Relative values for the reduction potential of protonated and deprotonated forms of (H2dcbpy)2RuX2 may be estimated from the MO calculations. If [(H-dcbpy)(H2-dcbpy)RuX2] is formed, the first reduction process is predicted to be determined by the ligand with the lowest 1* orbital, which therefore remains as H2-dcbpy. In this situation no dramatic change in the value of the redox potential is expected relative to that for (H2-dcbpy)2Ru(NCS)2. In contrast, complete deprotonation, yielding [(dcbpy2)Ru(NCS)2]4, should result in a negative shift of the potential for the first reduction process by 1.7V since the value is now determined by the position of the 1* orbital of dcbpy2. Neither of these scenarios is observed experimentally. However, removal of one proton from each H2-dcbpy ligand to give the [(H-dcbpy)RuX2]2 complex is predicted to give a reduction potential for the first process which is about 0.7V more negative than that of (H2dcbpy)2RuX2. The experimental observation of a potential difference of 0.65V between the protonated and electrochemically deprotonated complex therefore implies that electrochemical reduction of (H2-dcbpy)2Ru(NCS)2 is likely to have generated the doubly deprotonated anion, [(Hdcbpy)RuX2]2, rather than singly, triply or quadruply deprotonated species. No reports on the reduction of metal coordinated bipyridine carboxylic acid ligands are available in the literature. However, extensive polarographic studies on acids of single ring analogues (pyridine derivatives) have been reported in aqueous media [76, 148, 206-213]. For example, the

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 148

polarographic reduction of isonicotinic and picolinic acids have been studied as a function of pH by Volke and Volkov [207, 208]. In this report, changes in the polarographic wave heights and Er1/2values were discussed in terms of the acid-base behavior of the compounds, although proposed reaction products were not isolated. Lund [206] demonstrated that the reduction of isonicotinic acid at a mercury electrode gave high yields of the corresponding aldehyde in acidic solutions. Campanella and coworkers [211, 212] explained the existence of up to five reduction processes for the reduction of dipicolinic and isocinchomeronic acid in terms of acid-base behavior and attributed the origin of some of the processes to a combination of deprotonation, proton reduction and irreversible formation of molecular hydrogen, although the absence of proof of hydrogen gas formation resulted in criticism of some mechanistic aspects of these studies [210]. (e) Voltammetry at Positive Potentials as a Function of Deprotonation The fact that reduction processes are shifted to more negative potentials by deprotonation implies that oxidation will become easier, if the metal d-orbitals and ligand orbitals are affected in an analogous manner. Under low scan rate conditions of cyclic voltammetry at macrodisk electrodes, oxidation of (H2-dcbpy)2Ru(NCS)2 is chemically irreversible and complex in DMF (see Figure 5.6a and Chapter 4). The initial process observed in Figure 5.6a is the metal based oxidation process, whilst oxidation behavior at more positive potentials, not considered further in this Chapter, is related to interaction of (H2-dcbpy)2Ru(NCS)2 and/or products of electrolysis with the electrode surface and has been studied in Chapter 4. Use of microdisk electrodes and use of very fast scan rates (>100Vs1) are in fact necessary to achieve any evidence of chemical reversibility for the oxidation of protonated (H2-dcbpy)2Ru(NCS)2 (see page 127). Despite the irreversibility of some processes and significant interaction with the surface, the influence of deprotonation on the oxidation potential can be determined by periodically recording cyclic voltammograms at a macrodisk glassy carbon electrode over the potential range of 0.6V to +1.0V during the course of reductive bulk electrolysis at a platinum basket mesh electrode (Eappl= 2.0V, experimentally details given in caption to Figure 5.6). When one electron per (H2dcbpy)2Ru(NCS)2 molecule has been transferred, the appearance of a new oxidation process having a peak potential about 150mV less positive than the initial oxidation process may be noted, as may a decrease in the peak height for the initial process (Figure 5.6b). This process has the characteristics of surface interaction. As the bulk electrolysis proceeds further (Figure 5.6c-e) to what is believed to be a deprotonated form of (H2-dcbpy)2Ru(NCS)2, another new process with different character is detected at an even less positive potential (about 0.3V prior to oxidation of protonated complex).

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 149

f)

e)
+20 HBF4

d)
+0.7e
_

5A

c)
+1e
_

b)
+1e
_

a)
+1e
_

-0.5

0.0

0.5

potential [V] vs. Fc/Fc

1.0

Figure 5.6. Cyclic voltammograms (=100mVs1) obtained at a glassy carbon working electrode (d=1mm) over the potential range where the metal based oxidation process occurs of 1.1mM (H2-dcbpy)2Ru(NCS)2 at a platinum gauze electrode in DMF (0.1M Bu4NPF6) during the course of a reductive bulk electrolysis (Eappl=2030mV). a) before commencement, b) one electron transferred, c) two electrons transferred, d) three electrons transferred, e) 3.7 electrons transferred (exhaustive electrolysis), f) 20 equivalents of HBF 4 added to the solution.

After exhaustive deprotonation by reduction, is followed by a series of complex processes (Figure 5.6e). As expected, if the processes following the first oxidation process are associated with oxidation of hydroxyl anions, these new processes were also shown to increase in intensity as Bu4N(OH) is added to solutions of (H2-dcbpy)2Ru(NCS)2 (see below). To further support the hypothesis that reduction of (H2-dcbpy)2Ru(NCS)2 leads to protons being lost by reduction to hydrogen gas, the exhaustively electrolyzed solution was titrated with an HBF4 solution in DMF. It was found that the cyclic voltammogram for the oxidation of the fully protonated (H2dcbpy)2Ru(NCS)2 complex could be regenerated by addition of 3.50.5 mole equivalents of HBF4 solution to the electrolyzed solution. Thus, within experimental error, the amount of added protons required for regeneration of the protonated complex is equivalent to the number of electrons consumed (moles coulomb relationship and Faraday's law) during the course of reductive electrolysis. In the presence of excess acid (4 mole equivalents) the cyclic voltammogram (Figure 5.6f) remains essentially identical to that of the starting solution (Figure 5.6a). The result obtained

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 150

from this acid titration experiment is as expected if the postulated mechanism involving formation of deprotonated forms of (H2-dcbpy)2Ru(NCS)2 via exhaustive electrolysis is correct. The chemical reversibility of the least positive process detected in exhaustively electrolyzed solutions was assessed under conditions of cyclic voltammetry by switching the potential at about 0.27V. As shown in Figure 5.7, the chemical reversibility increases with increasing scan rate. Despite the fact that a peak current ratio (ipox/ipred) of unity is not achieved at scan rates up to 2000mVs1, (Epox + Epred)/2-values calculated from cyclic voltammograms over the scan rate range of 502000mVs1 were constant within experimental error (see Table 5.1), as required for a reversible process. Thus, extent of departure from being an ideally reversible process is small, with respect to calculations of the Er1/2-value. Under steady-state conditions at a rotating disk electrode (f2000min1), Er1/2-values obtained from "log-plots" were in good agreement with Er1/2-values obtained from cyclic voltammograms. However, the slopes of "log-plots" were slightly larger than theoretically predicted for a reversible process (Table 5.1). Thus, this initial process detected after exhaustive electrolysis is assigned to the deprotonated [(H2-x/2-dcbpyx/2)2RuII/III(NCS)2]x/(x+1) metal based oxidation reaction, and the reversible half wave potential is established to be +935mV. From a Levich plot (rotating disk electrode), the diffusion coefficient for [(Lx/2)2Ru(NCS)2](x1) in DMF is D=2.70.4106cm2s1, which is in excellent agreement with the value (D=2.50.4106cm2s1) calculated from voltammograms obtained for the reduction processes (see above).

8 6

current [A]

4 2 0 -2 -4

-0.6

-0.4

-0.2

0.0
+

0.2

potential [V] vs. Fc/Fc

Figure 5.7. Cyclic voltammograms for the oxidation of electrogenerated [(H2x/2-dcbpyx/2)2Ru(NCS)2]x in DMF (0.1M Bu4NPF6) at a glassy carbon electrode as a function of scan rate (=50, 100, 200, 500, 1000 and 2000mVs1).

The shift of 0.3V for oxidation of protonated and deprotonated forms of (H2-dcbpy)2Ru(NCS)2 reflects the altered electronic environment of the ligands and their influence on the ruthenium metal d-orbitals. The fact that oxidation becomes easier after deprotonation and reduction more difficult is consistent with the increased overall negative charge on the complex that accompanies deproton-

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 151

ation. Interestingly, Zaban and co-workers [214] investigated the oxidation of Ru, Mg and Fe photovoltaic sensitizer complexes with ligands containing phosphonic or carboxylic acid groups at platinum and glassy carbon electrodes in aqueous media, but did not observe a pH dependence. In contrast, a pH dependence attributed to electronic changes in the semiconductor electrolyte interface, was reported when these dyes were attached to semiconductor electrode surfaces. (f) Studies on Bu4N-salts of Deprotonated (H2-dcbpy)2Ru(NCS)2 To relate the significance of the above findings to photovoltaic cells, voltammetric studies on "salts" used in photovoltaic cells were investigated., These "salts" are formed by reaction of (H2dcbpy)2Ru(NCS)2 with Bu4N(OH), which is a very strong base [215] in organic solvents. The Bu4N-salt of (H2-dcbpy)2Ru(NCS)2, assigned as [(H-dcbpy)2Ru(NCS)2](Bu4N)2 and assumed to be the deprotonated material used in photoelectrochemical cells [203], is synthesized via reaction of (H2-dcbpy)2Ru(NCS)2 with Bu4N(OH). When the electron transfer process is reversible and protons participate in rapidly established acid-base equilibria, a linear dependence of log(base concentration) on the reversible potential is expected (Nernst relationship). Alternatively, if the acid and base forms of (H2-dcbpy)2Ru(NCS)2 are not in equilibrium on the voltammetric time scale, the peak height of the protonated form will decrease as base is added. Concurrently, an increase in peak height will be observed for the deprotonated form. However, in this non equilibrium situation, the reversible potentials of both processes will remain unaffected by the base concentration. Cyclic voltammograms obtained at glassy carbon electrodes for solutions of (H2-dcbpy)2Ru(NCS)2 to which increasing concentrations of Bu4N(OH) solution was added closely resembled those obtained during the course of reductive bulk electrolysis when deprotonation of (H 2dcbpy)2Ru(NCS)2 was assumed to occur (Figure 5.4b-d and Figure 5.6b-e). Thus, when the number of added equivalents of Bu4N(OH) was 4, cyclic voltammograms were indistinguishable from those obtained over the same negative potential range with bulk electrolyzed solutions (Figure 5.4d and Figure 5.6e). These data imply that deprotonated complexes are formed by bulk reductive electrolysis and that protonated and deprotonated forms of (H2-dcbpy)2Ru(NCS)2 are not in equilibrium on the voltammetric time scale. The Er1/2-value for the [(H2-x/2-dcbpyx/2)2Ru(NCS)2]x/(x1) redox couple, calculated from the deprotonated complex formed by reductive electrolysis, was compared with the values determined for "salts" of (H2-dcbpy)2Ru(NCS)2 by cyclic voltammetry when more than three equivalents of Bu4N(OH) were present. Under these conditions the voltammetric process is clearly defined, whereas for two equivalents, a complex response is evident. Results plotted in Figure 5.8a reveal that the Er1/2-value for the first process for solutions of the "salts" is independent of the concentration of Bu4N(OH), within experimental error, and in agreement with the Er1/2-value determined for the first reduction process of bulk electrolyzed solutions of (H2-dcbpy)2Ru(NCS)2. This result again confirms, that [(Lx/2)2Ru(NCS)2]x is not in a rapid acid-base equilibrium with (H2dcbpy)2Ru(NCS)2 on the voltammetric time scale and that the same species are formed by bulk electrolysis and direct addition of a strong base.

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 152

potential [mV] vs. Fc/Fc

-2140

a)

-2160

b)
-2260

-2280 0.2

0.4

0.6

0.8

1.0

1.2

1.4

log(added TBA(OH)-equivalents)
Figure 5.8. Dependence of Bu4N(OH) concentration on the reversible half-wave potential for the process for reduction of [(H2x/2-dcbpy)x/2)2RuX2]x. Horizontal dashed lines indicate the reversible half-wave potential obtained after exhaustive controlled potential reductive electrolysis. a) (H2-dcbpy)2Ru(NCS)2 b) (H2dcbpy)2Ru(CN)2.

(g) In Situ Reductive OTTLE Experiments As noted in the introduction, electronic spectra are highly indicative of the formation of reduced forms of ruthenium polypyridine compounds. Changes in electronic spectra that occurred during the course of the bulk electrolysis of (H2-dcbpy)2Ru(NCS)2 were obtained in situ by use of an OTTLE (platinum electrode) experiment. In DMF, (H2-dcbpy)2Ru(NCS)2 shows (Figure 5.9) two MLCT bands in the visible region and one dcbpy ligand based * transition band in the UV region [29, 38]. After a potential of 1.95V has been applied at 22C (Figure 5.9), all three bands associated with the occurrence of deprotonation are red shifted to higher energy (Table 5.2) and decrease in intensity (Figure 5.9). No isosbestic points are detected, implying that a range of intermediates may be formed (and reaction pathways occur) during the course of electrolysis. However, the spectrum of the starting solution of (H2-dcbpy)2Ru(NCS)2 could be quantitatively regenerated (no detectable loss of absorbance) when a four fold or greater concentration excess of HBF4 was added to the electrolyzed solution. When the electrolysis experiment is conducted in the OTTLE cell at 58C, an additional band in the near infra-red region is observed at 6300cm1 (see insert in Figure 5.9) in initial stages of the electrolysis. Importantly, it was observed that this band collapses prior to exhaustive electrolysis being achieved, so that the final spectrum observed at 58C is essentially identical to that recorded for the deprotonated sensitizer at 22C. The band at 6300cm1 is characteristic of formation of the reduced dcbpy ligand and has been found in the electronic spectra of the reduced ester analogue (compare Figure 3.22 on page 97). Thus, at 58C the initial reduction process at platinum electrodes incorporates a step including formation of [(H 2dcbpy)(H2-dcbpy)Ru(NCS)2] or [(H2-dcbpy)2Ru(NCS)2]2. Hence, at low temperature, the pathway for generation of deprotonated [(H2x/2-dcbpy)2Ru(NCS)2]x also involves a homogeneous reaction of reduced (H2-dcbpy)2Ru(NCS)2 with H+, H2O or (H2-dcbpy)2Ru(NCS)2. This result

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 153

proves that homogeneous reaction pathways are available for generation of hydrogen gas at platinum electrodes.

molar absorbance [10 M cm ]

-1 4

-1

5.0 4.0 3.0 2.0 1.0 0.0 10000 20000 30000


-1

40000

energy [cm ]
Figure 5.9. Electronic spectra obtained during the course of OTTLE experiments on the reductive electrolysis of 0.5mM (H2-dcbpy)2Ru(NCS)2 in DMF (0.1M Bu4NPF6) at 22C. a) Eappl=1.95V; insert in the near infra-red region shows additional band observed when experiment is conducted at 58C.

In other studies [157], the position of the MLCT bands has been correlated with the reversible potentials of the ligand based reduction and metal centered oxidation processes. It seems reasonable to assume that this argument can be extended to the comparison of protonated and deprotonated complexes by effectively treating them as a different form of ligand [174]. On this basis, it follows that, since the reversible potential for ligand based reduction changes by 0.6V (negative direction) and that of metal based oxidation by only 0.3V (negative direction, vide infra), that deprotonation results in the energy of the ligand *-orbitals being increased to a greater extent than the metal d-orbitals. It therefore also follows that the MLCT band, which formally can be assigned to a Ru(d)H2-dcbpy(*) transition, increases in energy on deprotonation. 5.3.2. Electrochemical Reduction of (H2-dcbpy)2Ru(CN)2 in DMF (a) Voltammetry in DMF As is the case with (H2-dcbpy)2Ru(NCS)2, the initial two voltammetric processes for reduction of (H2-dcbpy)2Ru(CN)2 at a platinum electrode (Figure 5.10a) were significantly different to that observed when glassy carbon was the electrode material (Figure 5.10b). If the switching potential in glassy carbon electrode experiments was set before the onset of the second process (1.8V), the first process exhibited ipred/ipox-values of almost unity for scan rates >1000mVs1, although chemical irreversibility was evident at lower scan rates. This first process was also irreversible when a glassy carbon rotating disk electrode was used at low rotation rates ("Log-plots" were nonlinear and had slopes of 90mV). In contrast, at high rotation rates (3000rpm) the first process

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 154

was close to reversible as slopes of "log-plots" were now 705mV and the E1/2-values coincided with values obtained with cyclic voltammetry at high scan rates (>200mVs1). The estimated reversible Er1/2-value for the [(H2-dcbpy)2Ru(CN)2]0/ (T=+25C) couple is contained in Table 5.2 under a range of conditions.
1A

a)

b)
1A

0.2A

c)

1A

d)

-2.5

-2.0

-1.5

-1.0
+

-0.5

potential [V] vs. Fc/Fc

Figure 5.10. Cyclic voltammograms (=100mVs1) for reduction of (H2-dcbpy)2Ru(CN)2. a) platinum disk electrode, c=1.1mM, T=25C; b) glassy carbon electrode, c0=1.1mM, T=+25C; c) glassy carbon working electrode, c=1.1mM, T=58C; d) electrogenerated [(H2x/2-dcbpy)x/2)2Ru(CN)2]x, c0=1.3mM, glassy carbon disk electrode, T=25C.

The chemical reversibility of the (H2-dcbpy)2Ru(CN)2 reduction processes significantly improved when the temperature was lowered to 58C. The first process was close to reversible at this temperature under conditions of cyclic, rotating disk and microdisk electrode voltammetry when glassy carbon electrodes were used. The slopes of "log-plots" (rotating disk and microdisk electrode) confirmed the one-electron nature of the [(H2-dcbpy)2Ru(CN)2]0/ process. The second process required higher scan rates (>1000mVs1) before an ipred/ipox-value of unity was approached at 58C. The reversible half-wave potential for the [(H2-dcbpy)2Ru(CN)2]/2 process at 58C could be calculated from cyclic voltammograms at scan rates 500mVs1 and from rotating disk electrode experiments (f2000rpm). Data obtained under these conditions are summarized in Table 5.2. Mass transport controlled data at the glassy carbon microdisk electrode could not be obtained at 58C because of significant adsorption. (b) Bulk Reductive Electrolysis in DMF When (H2-dcbpy)2Ru(CN)2 was reduced by bulk electrolysis at a platinum gauze working electrode using a potential of Eappl=2.0V vs. Fc/Fc+, the formation of H2 was detected by gas

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 155

chromatographic analysis of the supernatant gas phase. This result implies, that an overall deprotonation reaction occurs in this long timescale experiment as is the case with reduction of the thiocyanate analogue. After exhaustive electrolysis, the number of electrons per molecule transferred was n=3.20.3. A cyclic voltammogram obtained at a glassy carbon electrode over the negative potential range is shown in Figure 5.10d. In contrast to measurements made with the protonated form of (H2-dcbpy)2Ru(CN)2, both reduction processes of deprotonated [(H2xdcbpyx/2)2Ru(CN)2]x are fully reversible under all voltammetric conditions employed. Er1/2-values calculated from cyclic and near steady-state microdisk and rotating disk electrode measurements after bulk electrolysis and deprotonation coincided for both processes (Table 5.2). Slopes of "Logplots" derived from steady-state measurements confirmed that both reduction steps involved oneelectron charge transfer processes. (c) The oxidation Process as a Function of Deprotonation The Er1/2 value for the oxidation of (H2-dcbpy)2Ru(CN)2 in DMF is readily measured, since the [(H2-dcbpy)2Ru(CN)2]0/+ process requires only moderately high scan rates (500mVs1) to become fully reversible (Table 2). After exhaustive bulk reductive electrolysis at a platinum electrode as described above for the thiocyanate analogue, the metal based oxidation process for deprotonated (H2-dcbpy)2Ru(CN)2 is shifted at glassy carbon electrodes to a greater extent negative (0.36V) than is the case with the thiocyanate analogue (0.30V). As evidenced by examination of data contained in Table 5.2, Er1/2 for the [(H2x-dcbpyx/2)2Ru(CN)2]x/(x+1) process is located at +210mV, which is 360mV less positive than found for oxidation of the protonated form. The process for oxidation of deprotonated (H2-dcbpy)2Ru(CN)2 is followed by a series of complex, possibly surface based, processes at more positive potentials, similar to those shown for the thiocyanate analogue in Figure 5.6. (d) Reductive OTTLE Experiments The electronic spectra obtained during reductive electrolysis (Eappl=2.0V) and deprotonation of (H2-dcbpy)2Ru(CN)2 were monitored with an OTTLE cell arrangement (platinum working electrode) to give the result presented in Figure 5.11a. No isosbestic points were observed and the absorption bands of [(H2x-dcbpyx/2)2Ru(CN)2]x are shifted to higher energy and have a smaller extinction coefficient compared to (H2-dcbpy)2Ru(CN)2. All these electronic spectral data imply that the energy gap between the ruthenium d-orbitals and the ligand -orbital has increased, as also predicted on the basis of voltammetric data. The result of an OTTLE electrolysis experiment on deprotonated (H2-dcbpy)2Ru(CN)2 at a potential dcbpy
x/2

more

negative
x

than

the

first

reduction

process

of

deprotonated

[(H2x-

)2Ru(CN)2] (Eappl=2.4V) is shown in Figure 5.11b. In this case, exhaustive electrolysis

could never be achieved (zero current not attained) with a steady-state condition attained shown in Figure 5.11b. The spectrum obtained after reduction of the deprotonated complex indicates that a dcbpy ligand based reduction process has occurred, since it resembles that of the one-electron

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 156

reduced form of the ester analogue ([(Et2-dcbpy)2Ru(CN)2]) which is shown in Figure 3.22 on page 97. The occurrence of four isosbestic points in this OTTLE experiment is also noteworthy, as is the fact that the electronic spectrum of the starting solution ([(H2x-dcbpyx/2)2Ru(CN)2]x) could be quantitatively regenerated by electrochemical oxidation (Eappl= 1.0V). The NIR absorption band at 8000cm1, is expected for a bpy type ligand based reduction process [122, 123]. In Chapter 3.5.5 on page 104 this band has been assigned to an intra-ligand 1*2* transition of the reduced dcbpy ligand, but occurs at higher energies than for the ester analogue ( 6000cm1, see Table 3.7 on page 98) which again highlights the altered electronic environment of the ligand caused by deprotonation.

molar excitinction [10 M cm ]

-1 -4 -1 -1 -4 -1

6 5 4 3 2 1 0 6

molar excitinction [10 M cm ]

5 4 3 2 1 0 5000

10000 15000 20000 25000 30000 35000

energy [cm ]
Figure 5.11. Electronic spectra obtained during the course of OTTLE experiments on the reductive electrolysis of 0.7mM (H2-dcbpy)2Ru(CN)2 in DMF (0.1M Bu4NPF6) at 22C. a) Eappl=2.0V and formation of [(H2x/2-dcbpyx/2)2Ru(CN)2]x; b) ligand based reduction of [(H2x/2-dcbpyx/2)2Ru(CN)2]x, Eappl=2.4V.

-1

As observed with [(H2x-dcbpyx/2)2Ru(NCS)2]x, no rapid spectral changes could be detected when the applied potential was set to a very negative value of 2.7V, which is more negative than the reversible potential for the second reduction process. Rather after prolonged electrolysis time (t>30min), decomposition was observed and irreversible spectral changes occurred as evidenced by the loss of isosbestic points.

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 157

(e) Bu4N+ "Salts" of (H2-dcbpy)2Ru(CN)2 Bu4N+ "salts" of known stochiometry were prepared (see Experimental). The reversible reduction potentials for Bu4N+ "salts" were measured over a range of 2 to 20 equivalents of Bu4N(OH). As shown in Figure 5.8b, the Er1/2-value for this process is independent of hydroxide concentration over this range and coincides with the value obtained from electrochemically generated [(H 2xdcbpyx/2)Ru(CN)2]x. 5.3.3. Reduction of (H2-dcbpy)2Ru(NCS)2 in Acetone (a) Voltammetry, Adsorption and Bulk Electrolysis in Acetone Voltammetric studies in acetone on the reduction of (H2-dcbpy)2Ru(NCS)2 were substantially more complicated than in DMF. Figures 5.12a and b show the reduction of a 1.1mM solution of (H2-dcbpy)2Ru(NCS)2 in acetone (0.1M Bu4NPF6) at glassy carbon electrodes under conditions of cyclic voltammetry and near steady-state.

20A

5nA

20nA

-2.5

-2.0

-1.5

-1.0
+

-0.5

potential [V] vs. Fc/Fc

Figure 5.12. Voltammetric reduction of (H2-dcbpy)2Ru(NCS)2 in acetone (0.1M Bu4NPF6) at glassy carbon electrodes. a) Cyclic voltammetry, =100mVs1, electrode diameter 2mm, c=1.1mM, b) microdisk electrode, =10mVs1, electrode radius 5.5m, c=1.1mM, c) cyclic voltammetry, =1000mVs1, c=0.05mM, 2mm diameter glassy carbon electrode, consecutive scans, first scan dashed line, background subtracted.

The voltammograms are clearly different to those obtained under similar conditions in DMF (compare Figures 5.12a and c). No reversible mass transport controlled processes are observed, large abrupt changes in current and sudden discharges of the electrode (spikes) indicate the presence of adsorption. Under conditions of cyclic voltammetry the onset of the first cathodic current wave starts about 0.9V, which is over 0.6V more positive than the expected first Faradaic

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 158

reduction process for the protonated acid. However, when the concentration of (H2dcbpy)2Ru(NCS)2 is lowered in the sub-millimolar range (c0.1mM) and moderately increased scan rates 500mVs1 are used, two reversible processes can be observed on consecutive cyclic voltammetric scans, Figure 5.12c. The first cycle is still dominated by adsorption discharge and massive current spikes. However, on the second and third scans these spikes have disappeared and two reversible processes appear at 2.110.01V and 2.420.015V vs. Fc/Fc+, although the second process required higher scan rates ( 2000mVs1) before an ipred/ipox-value of unity was approached. Apparently, under these conditions material is not given enough time to adsorb onto the electrode surface. It is noted that the peak currents obtained for the two reversible processes from the third cycle are smaller than when obtained from the second cycle. This observation may be explained by assuming that material accumulated on the electrode surface prior to the first scan is stripped off on the first cycle, therefore increasing the concentration of material near the electrode surface above the level in bulk solution.

The two reversible processes are due to reduction of a deprotonated form of (H2dcbpy)2Ru(NCS)2 similar to those observed in reductively electrolyzed solutions in DMF. This is confirmed by reductive controlled potential bulk electrolysis of (H2-dcbpy)2Ru(NCS)2 in acetone at a platinum gauze electrode. In an bulk electrolysis experiment the potential was set to 1.9V vs. Fc/Fc+ and the exhaustive electrolysis consumed 51 electrons per molecule and hydrogen gas evolution was detected by gas chromatography. It can be expected that a similar reduction/hydrogen evolution/acid-base scheme as detected in DMF and described in Equations 5.4 to 5.15 is applicable. The larger number of electrons required in acetone for generation of deprotonated forms of (H2-dcbpy)2Ru(NCS)2 than in DMF may be attributable to a different acid-base behavior of acetone and/or of the complex in this solvent. After electrolysis two processes can be observed in the negative potential range at 2.09V and 2.41V when glassy carbon electrodes are used. These values are in excellent agreement with values measured from consecutive cyclic voltammograms on the protonated dye in dilute solutions shown in Figure 5.12c. The first process is reversible at low scan rates whereas the second process requires higher scan rates ( 5000mVs1) to achieve chemical reversibility. Slopes from "logplots" (rotating disk electrode) were 654mV, which confirm the expected one-electron nature of the processes. (b) Adsorption Studies of (H2-dcbpy)2Ru(NCS)2 in Acetone in the Negative Potential Range Further confirmation of (H2-dcbpy)2Ru(NCS)2 adsorbing onto electrode surfaces upon application of negative potentials was obtained from EQCM studies. Figure 5.13a shows the mass change that occurs on a 5mm diameter gold electrode surface during four potential cycles under conditions of cyclic voltammetry. Holding the potential at an initial value of 0.6V does not result in a detectable mass change, however, as the potential is swept toward about 0.8V, a mass increase of

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 159

about 3.5g occurs, which is followed at 1.1V by a mass decrease of 4g. On the reversal and on the following three cycles the mass changes are very small and it appears that a kind of steady-state situation has been reached. It should be noted, that the mass change occurs significantly prior to the expected potential for the reduction of (H2-dcbpy)2Ru(NCS)2 (Er1/2=1.53V in DMF). The mass changes observed in the order of g on a 5mm diameter electrode correspond to several hundred mono layers of (H2-dcbpy)2Ru(NCS)2. Compare Chapter 4.3.4 on page 122 where similar mass changes have been found to occur during an oxidative cycle, but were less expressed. Details on surface interactions that occur during the oxidation process can be found in chapter 4.3.4. Further evidence for the high amount of material precipitating onto the electrode surface in acetone was found during voltammetric experiments. Working electrodes were coated with an intense red film after they were cycled only briefly in the negative potential range. Similarly, after oxidative cycles the counter electrode, a platinum wire, was found to be coated by layers of a dark red solid substance which were occasionally so thick that they could be peeled off by hand, which suggest that a form of polymerization takes place. The large currents measured when the potential is cycled into the negative potential range (see Figure 5.12a) suggest that a conducting polymer may have been formed. Interestingly, this dark red solid product could not be produced under conditions of controlled potential bulk electrolysis. Formation of polymers under similar conditions has been also observed in the following case. Caix-Cecillon and co-workers [216] have investigated the reduction of the Ru(R2-bpy)(CO)2Cl2 series of complexes where R2-bpy stands for 2,2'bipyridine which is substituted in the 4 positions. Reduction of Ru(R2-bpy)(CO)2Cl2 in CH3CN leads to formation of conducting [Ru(R2-bpy)(CO)2]n polymer via reductively induced ligand elimination of the chloride ligands. This is surprising, since in case of the model ester compounds investigated in this work (see Chapter 3.5.4) reductively induced ligand elimination leads to formation of monomeric solvent substituted complexes. Results from double step chronocoulometric experiments agree with those obtained from EQCM studies. In an experiment with 0.9mM (H2-dcbpy)2Ru(NCS)2 in acetone (0.1M Bu4NPF6) the forward potential (Eforward) was set to 1.7V (this potential is more negative than the expected reversible reduction potential for the protonated dye) and the initial potential (Einitial) was stepwise varied between 1.4V and +0.3V. Before the double step experiment was initiated, the freshly polished gold electrode was held at Einitial for precisely 10 seconds to allow material to adsorb in a time controlled manner. From a linear plot (single or double step chronocoulometric experiments) of charge versus square root of time ("Anson-plot" [119, 217], also see Chapter 2.6.1 on page 50) the plot should be linear and pass through zero for a background corrected sample if no adsorption takes place but will be offset by a value corresponding to the amount of substance adsorbed on the electrode surface [218]. A typical "Anson-plot" of (H2-dcbpy)2Ru(NCS)2 in acetone is shown in Figure 5.13b with the insert showing a plot of substance adsorbed (offset in "Anson-plot") versus Einitial. This plot resembles that observed using the EQCM technique, that is, a large amount of material is adsorbed when the initial potential reaches a value below 0.6V. The maximum amount

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Results and Discussion

page: 160

adsorbed (7C offset) was calculated to be about ten times higher than that observed during EQCM studies. However, this is not unexpected, since a twentyfold higher concentration of (H2dcbpy)2Ru(NCS)2 was used with this experiment.

mass change [g]

6 4 2 0 0 50

0 -50 -100

current [A]

a)
100 150 200 250 300

-150 -200 -250

time [s]
20
8 6 current [A] 4 2 0 -1.5 -1.0 -0.5 0.0 + potential [V] vs. Fc/Fc

charge [C]

10 0 -10 -20 -30 0 5 10 15

b)

20
1/2

25

30

sqr(time) [s ]
Figure 5.13. EQCM and chronocoulometric adsorption studies on the reduction of (H2-dcbpy)2Ru(NCS)2 in acetone. a) EQCM measurement, 0.05mM (H2-dcbpy)2Ru(NCS)2 in acetone (0.1M Bu4NPF6), 5mm gold electrode, =50mVs1, dotted line shows potential profile (arbitrary units) Einitial=0.6V, Efinal=2.3V b) "Anson plot" of a double step chronocoloumetric experiment, dashed lines indicate the offset, 0.9mM (H2dcbpy)2Ru(NCS)2 in acetone (0.1M Bu4NPF6), 2mm gold electrode, Einitial=1.1V, Eforward=1.7V, insert shows forward offset as a function of Einitial, Eforward=1.7V.

In the previous chapter, the adsorptive behavior of (H2-dcbpy)2Ru(NCS)2 has been attributed to the thiocyanate ligands which are known [195] to adsorb onto electrode surfaces. However, adsorption studies on the H2-dcbpy ligand also can be found in literature [196]. The H2-dcbpy ligand adsorbs particularly strong onto platinum electrodes which results in very complicated responses when cyclic voltammetric measurements are carried out [196]. Thus, it can be concluded that it is the unfortunate combination of thiocyanate and H2-dcbpy ligands that complicates the voltammetric responses of (H2-dcbpy)2Ru(NCS)2 under certain conditions.

5.4. Conclusions
5.4.1. Summary The reversible potentials for the ligand based [(H2-dcbpy)2RuX2]0/ and [(H2-dcbpy)2RuX2]/2 reduction processes have been determined in DMF. Short time domains, reduced temperatures and glassy carbon electrodes are required in order to obtain chemically and electrochemically reversible

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Conclusions

page: 161

responses for all processes under voltammetric conditions. At platinum electrodes the reduction of both compounds is considerably more complex than at glassy carbon. Data suggest that this complexity is associated with the much lower over-potential for hydrogen ion reduction to hydrogen gas at platinum relative to glassy carbon electrodes. Reduction under long time scale conditions of bulk electrolysis resulted in overall deprotonation and formation of molecular hydrogen at platinum electrodes. However, partial rather than complete deprotonation of (H 2dcbpy)2RuX2 occurs as established from the number of electrons transferred in bulk electrolysis experiments. Spectroscopic data and voltammetric studies undertaken in the presence of deliberately added acid or base imply that the number of protons removed is of the order of two. This number is supported by comparison of data obtained from molecular orbital calculations and measured reversible potentials for reduction of protonated and deprotonated forms of (H2dcbpy)2Ru(NCS)2, which suggests that [(H-dcbpy)2RuX2]2 is formed by reductive electrolysis of (H2-dcbpy)2RuX2. 5.4.2. Relationship of Results to Photovoltaic Cells The potential for the first reduction process is shifted by 0.65V (more negative) when deprotonated [(H2x-dcbpyx/2)2RuX2]x is formed. In contrast the reversible potential for the oxidation of deprotonated complexes is only shifted by about 0.3V. Figure 5.1 suggests that no enhanced cell performance is expected if the reversible potential for the oxidation of the sensitizer is lowered, as occurs upon deprotonation. Indeed the thermodynamic driving force required for the dye regeneration by the electrolyte ion, I, will decrease. The increased performance of deprotonated forms of the sensitizer in photovoltaic cells is therefore likely to be a result of considerations not associated with dark potentials. For example, band potentials of nano-sized TiO2 semiconductor electrodes are very sensitive toward the solution environment [214, 219] (also see Chapter 7) and can shift dramatically when small changes are made to the electrolyte counter ions present in the solution phase [214]. A related effect also may apply, if the TiO2 semiconductor band potentials shift when sensitizers containing different levels of deprotonation and types of counter ions are used. Alternatively, the addition of base to TiO2 photovoltaic cells may also lead to deprotonation of TiO2 surface hydroxyl groups [47] and thus will result in a change of surface states which in turn may have dramatic effects on the overall performance of the solar cell device. Electronic spectra obtained during the course of reductive electrolysis reveal that deprotonation also leads to the MLCT bands being shifted toward the UV region and, also, that lowering of the extinction coefficient occurs. These features would normally be expected to lead to poorer performance when deprotonated forms of (H2-dcbpy)2RuX2 are used as sensitizers in photovoltaic cells. Interestingly, deprotonated forms of the (H2-dcbpy)2Ru(NCS)2 sensitizer are commonly employed in TiO2 photoelectrochemical cells and found to exhibit higher energy conversion efficiencies [203] than the protonated form. On the basis of findings of this study, the explanation for this improved performance must be related to changes in thermodynamics or kinetics rather

Chapter 5: Reduction and Deprotonation of (H2-dcbpy)2RuX2 (X=NCS, CN) Conclusions

page: 162

than spectral properties because it can be concluded that the deprotonation causes the light harvesting MLCT bands to be shifted toward the UV region and also to decrease in absorbance. Importantly, it is noted that both features are generally considered to be detrimental to conversion efficiencies [220]. Consideration must be given to the initial electron excitation and injection processes. Reports on the binding mode of (H2-dcbpy)2RuX2 onto the TiO2 surface [63, 67-69] suggest that predominantly ester formation with TiO2 and two carboxylic acids on one H2-dcbpy ligand occurs, whereas the other two carbon acid groups, located on the second H2-dcbpy ligand, are available for interaction with solvent. If the latter two acid groups are deprotonated, the LUMO of the H2-dcbpy ligand not attached to the TiO2 surface will significantly rise in energy compared to the other surface attached H2-dcbpy ligand. The absorption spectrum of the complex will then consist of transitions into the protonated (TiO2 attached) and deprotonated bipyridine moiety. This is advantageous, since it broadens the absorption bands and increases the light harvesting spectrum of the sensitizer. Electron injection from the TiO2 attached ligand into the TiO2 occurs on the subpicosecond timescale [77] due to the good orbital overlap between the TiO2 manifold and H2-dcbpy ligand [220, 221]. However, it has been shown that direct linkage of the chromophore with the semiconductor is not necessary for efficient electron injection [136, 222]. Thus, the electron injection into the TiO2 from the non surface attached ligand occurs either directly or by an electron cascading effect via the * orbital of the surface attached ligand [136]. A more recent study [83], employing crystallographic data and molecular modeling, confirms the possibility of surface attachment via two acid groups but suggests that for steric reasons these two acid groups are located on different bipyridine ligands. Each dcbpy ligand can then use one carboxylate to form an ester linkage with the TiO2 surface, leaving the second carboxylate group in either the protonated or deprotonated form. Similar considerations apply as pointed out above, deprotonation of the remaining two acid groups leads to a broadening of the MLCT bands. Importantly, in this case a direct pathway exists for electrons promoted into the * orbitals of the deprotonated 4-carboxypyridine ring to cascade into the TiO2 conduction band thereby improving the light harvesting properties of the sensitizer.

Quantitative knowledge concerning the shift of redox potentials upon deprotonation offers the exciting possibility of being able to tune the redox potentials and MLCT bands to a desired value, but also implies that great care has to be taken in studies on (H2-dcbpy)2RuX2 complexes, whether attached to semiconductor surfaces or not, to ensure that a known and stable degree of protonation is maintained throughout the experiments.

Chapter 6

Electrochemical and Photophysical Properties of [(H3-tctpy)Ru(NCS)3]Bu4N.

part of this work is intended to be published in

'Dependence of the voltammetric oxidation of the highly efficient photovoltaic sensitizer [(H3-tctpy)RuII(NCS)3] (H3-tctpy = 2,2':6',2''-terpyridine-4,4',4''-tricarboxylic acid) on the electrode material, solvent and isomeric purity.' Georg Wolfbauer, Alan M. Bond, Glen B. Deacon, Julia Howitt, Douglas R. MacFarlane, Leone Spiccia, J. Electrochem. Soc., 1999, to be submitted

Chapter 6: A Novel Sensitizer: The Black Dye Introduction

page: 164

6.1. Introduction
The initial sensitizers investigated for use in nano-crystalline TiO2 solar cells (also compare Chapter 1.8.1 on page 24) were of the type [Ru(bpy)3]2+ (bpy = 2,2'-bipyridine or derivative). A major leap forward occurred with the use of H2-dcbpy (H2-dcbpy = 2,2'-bipyridine-4,4'-carboxylic acid) as the ligand which allowed effective linkage onto the TiO2 surface by formation of chemical ester bonds [67-69]. A further step forward was the replacement of one H2-dcbpy ligand by two halides or pseudo halides [70-72, 74, 75] which resulted in a red shift of the absorption spectrum of the dye and hence an increased light harvest. This series of advances culminated in the synthesis of cis-(H2-dcbpy)2RuII(NCS)2 [29], which is currently the most widely used sensitizer, allowing solar cells to be constructed with efficiencies of over 10% [29, 36, 72, 138, 139]. Numerous attempts to synthesize superior sensitizers [89, 90, 134, 136, 223] failed to improve the efficiency until 1997 when Nazeeruddin and co-workers [88] reported the synthesis of a dye with the structure [(H3tctpy)RuII(NCS)3] (H3-tctpy = 2,2':6',2''-terpyridine-4,4',4''-tricarboxylic acid, structure is contained in Figure 6.1), which outperformed cis-(H2-dcbpy)2Ru(NCS)2. Due to the fact that solutions of this dye are black in some organic solvents and lead to black coatings when applied to TiO2, it has been named the "Black Dye". This terpyridine complex sensitizes TiO2 effectively over the whole visible spectral range and even into the NIR region of the electronic spectrum.
COOH S C N S C N N C S COOH Ru N N N COOH

Figure 6.1. Structure of [(H3-tctpy)RuII(NCS)3] (Black Dye).

Surprisingly, even though [(H3-tctpy)Ru(NCS)3] is the best performing sensitizer available to date no detailed study describing its photophysical or electrochemical properties is available in the literature. The reversible potential for oxidation of a sensitizer employed in a photoelectrochemical cell is an important thermodynamic value related to characterize its physical/chemical properties. In studies described in the preceding chapters the problems have been verified that need to be overcome with the voltammetric determination of reversible potentials for the surface active cis-(H2dcbpy)2Ru(NCS)2 sensitizer. In this study, the voltammetric oxidation, reduction and effect of deprotonation of [(H3-tctpy)Ru(NCS)3] has been investigated in a variety of solvents and quartz crystal microbalance experiments conducted to estimate the amount of physisorption of the dye on conventional electrode materials. Voltammetric data gives a measure of the stability of the oxidized state in different solvents. High stability is a requirement for a long lived device, despite the

Chapter 6: A Novel Sensitizer: The Black Dye Introduction

page: 165

relatively short period of time (<1s [50]) that the oxidized state exists for in the device before it is reduced back to its original state by a charge mediator (commonly iodide). Finally, the thermodynamic differences associated with oxidation of linkage isomers containing S- bonded thiocyanates are considered and data obtained are compared to that reported for the (H2-dcbpy)2Ru(NCS)2 analogue.

6.2. Experimental
6.2.1. Instrumentation A standard 3-electrode setup has been used (Chapter 3.2 on page 62). All potentials are referenced against the oxidation of ferrocene (Fc/Fc+) under identical conditions as used in the respective experiment. The detailed experimental setup for bulk electrolysis experiments can be found in Chapters 3.2 and 5.2.2. Optically transparent thin-layer electrolysis (OTTLE) experiments, gas chromatographic measurements, and molecular orbital calculations have been carried out as described in Chapter 5.2.2 on page 134. Before molecular orbital calculations were performed all structures were minimized in energy. The terpyridine ligands exhibited an absolute energy minimum with the pyridine nitrogens in the trans position to each other. However, in order to maintain a configuration which is more closely related to terpyridine ligands coordinated to the ruthenium center, the ring systems were minimized into a local energy minimum where all three pyridine ring systems are in the cis position. Energy minimization calculations lead to structures with the pyridine rings slightly out of plane with respect to each other and the carboxylate groups being out of plane with the pyridine ring. The values for these tilt angles are identical to those obtained for the bipyridine analogues reported in Chapter 5.3.1d on page 145. For the partly deprotonated forms (H2-tctpy and H-tctpy2) the symmetrical isomer was chosen for the calculations, although the energy differences between the symmetrical and unsymmetrical isomers are small. The equipment used for electrochemical quartz crystal microbalance (EQCM) studies is described in Chapter 4.2.1 on page 113. 6.2.2. Reagents and Synthesis of Compounds Dimethylformamide (DMF), acetonitrile and acetone used for electrochemical and spectroscopic measurements were of HPLC grade (Mallinckrodt, ChromAR and UltimAR). The description of other reagents used can be found in Chapter 5.2.3 on page 135. For synthesis of samples of [(H3-tctpy)Ru(NCS)3]Bu4N used in this study, 190mg (515mol) 2,2':6',2''-terpyridine-4,4',4''-tricarboxylate and 143mg (545mol) RuCl3 were placed in a 100mL 3neck flask. After the addition of ca. 30mL DMF, the solution was bubbled with nitrogen for several minutes. A constant flow of nitrogen over the solution was maintained throughout the experiment.

Chapter 6: A Novel Sensitizer: The Black Dye Experimental

page: 166

The mixture was then heated to 120C and stirred for 2 hours. The temperature was lowered to 95C and KSCN, dissolved in ~15mL H2O/DMF (1:5), was added. The solution was stirred and the temperature maintained at 95C for 3 days. The DMF solvent was removed by rotary evaporation and the black solid residue was transferred into a beaker using water. The pH (ca. 3) of the solution was then increased to pH 910 by addition of 0.1M Bu4N(OH) solution. The crude product was precipitated as its partially deprotonated Bu4N-salt by titrating to pH 3.5 with 10% HClO4. The precipitate was separated by centrifugation and washed 3 times with acidified water (adjusted to pH 1.8 by 10% HClO4). For final purification, the crude product was chromatographed in its deprotonated form on a 4cm x 65cm column containing Sephadex LH20 gel exclusion resin. For this separation procedure, the black-greenish precipitate was dissolved in a minimal amount (5ml) of water. The pH was increased to ca. pH 79 by addition of a minimal amount of NaOH solution. The dark black solution was then placed onto the column and eluted with pure water. Several bands separated. The first band (reddish-purple) and the second band (black) were rejected and four fractions of the major third (product) band (dark black) were collected. The product obtained from the end of the four fractions was precipitated by titrating the solution pH to 1.6 with 10% HClO4. The precipitate was centrifuged and washed twice with acidified water (pH 1.6). The low pH was necessary to ensure that the fully protonated acid was obtained. Freeze drying gave a fine blackgreenish powder. Yield: 95mg (21%) pure [(H3-tctpy)Ru(NCS)3]Bu4N N,N,N- isomer (see below) which was used throughout this study unless otherwise mentioned. From extensive synthetic work conducted here at Monash University on the purification of the bipyridine analogue, (H2-dcbpy)2Ru(NCS)2, it is known that partial separation of the thiocyanate linkage isomers is possible using Sephadex LH20 gel chromatography (described in Chapter 5.2.3 on page 135). Furthermore, established literature is available on (H2-dcbpy)2Ru(NCS)2 which allows identification of the thiocyanate linkage isomers by NMR spectroscopy [80]. In the case of an isomeric mixture of (H2-dcbpy)2Ru(NCS)2, the generally desired doubly N,N- bonded isomer has the longest retention time and elutes last whereas the S,S- isomer elutes first followed by the mixed bond N,S- isomer. However, complete separation is not possible, rather enrichment of one isomer over the other may be achieved. In accordance with this knowledge I have interpreted the
1

H-NMR spectra shown in Figure 6.2 and obtained from the four fractions separated by chromatog-

raphy in the following way. The last fraction taken contains only one pure product (see Figure 6.2a) which is believed to be the triply N,N,N- bonded [(H3-tctpy)Ru(NCS)3] complex with the signal at 9.04, assigned to the 5,5'' protons of the triply N,N,N- bond isomer. The other fractions, which have eluted earlier, show the presence of a mixture of isomers. The N,N,N- isomer content of fractions 24 was estimated by measuring the relative intensity of the signal at 9.04ppm (marked with an arrow in Figure 6.2a).

Chapter 6: A Novel Sensitizer: The Black Dye Experimental

page: 167

3,3'' 3',5'H 5,5''H 6,6''H

a)

b)

c)

d)
10.0 9.5 9.0 8.5 ppm 8.0 7.5 7.0

Figure 6.2. 1H-NMR spectra recorded in 0.1M NaOD/D2O of fractions collected after chromatographic separation to give variable sulfur bonded thiocyanate isomer content. Assignment of signals tentatively. a) fourth fraction containing pure, N- bond thiocyanate linkage only; b) third fraction containing 42% of other linkage isomers; c) second fraction containing 77% of other linkage isomers and d) first fraction containing 100% other thiocyanate linkage isomers.

Since this compound contains three thiocyanate ligands, the number of possible linkage isomers is six. In the case of (H2-dcbpy)2Ru(NCS)2, the pyridine proton in the 6 position is in very close spatial vicinity to the introduced pseudo halides and hence a very good indicator of their electronic influence. The N- and S- bond isomer can be readily distinguished by a 1H-NMR chemical shift difference in the 6 proton by ca. 0.4ppm. However, the different spatial arrangement of [(H 3tctpy)Ru(NCS)3] means that no proton is in close vicinity to the pseudo halides so that only minor shifts are observed for proton resonances of the different linkage isomers. Having obtained a pure sample of the required N,N,N- bonded isomer, no effort was made to further distinguish or separate the other linkage isomers of [(H3-tctpy)Ru(NCS)3].

6.3. Oxidation: Results and Discussion


Voltammetric investigations described extensively in Chapter 4 on surface active (H 2dcbpy)2Ru(NCS)2 in a variety of solvents revealed the complex nature of the metal centered oxidation process relative to mass transport controlled and highly reversible systems such as [Ru(bpy)3]2+ [150, 224]. The [(H3-tctpy)Ru(NCS)3] dye is also likely to be surface active at electrodes so that measurements of the reversible potential may not be straightforward.

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 168

6.3.1. Studies on the Oxidation Process in Acetonitrile (a) The Metal Centered Oxidation Process Voltammetric studies on (H2-dcbpy)2Ru(NCS)2 in acetonitrile (page 126) are limited by the sparingly soluble nature of this compound. However, significant adsorption is detected in the oxidative voltammetry of this sensitizer when using a saturated solution. [(H3-tctpy)Ru(NCS)3] is considerably more soluble in acetonitrile and voltammetric studies at platinum, glassy carbon and gold electrodes were carried out in this solvent in concentration ranges of 0.1mM to 1.0mM. Depending on the experimental conditions, several processes may be observed under conditions of cyclic voltammetry (data shown later). However, cyclic voltammetric experiments undertaken at macrodisk electrodes in acetonitrile (0.1M Bu4NPF6) by scanning over the positive potential direction, but setting the switching potential to +0.50V, reveal only the presence of a chemically reversible process. Voltammograms obtained for this process as a function of scan rate are shown in Figure 6.3a. This process is assigned to the metal centered oxidation of [(H3-tctpy)Ru(NCS)3] and can be formulated as in Equation 6.1: [(H3-tctpy)RuII(NCS)3]
3 2 1 0 -1 -2 60 40 20 0 -0.4 1.0 -0.2 0.0 0.2 0.4
+

[(H3-tctpy)RuIII(NCS)3]0

+ e

(6.1)

current [A]

current [A]

0.6

potential [V] vs. Fc/Fc


log((ilim-i)/i)
0.5 0.0

-0.5 -1.0 220 240 260 280 300 320 340


+

360

potential [mV] vs. Fc/Fc

Figure 6.3: Voltammetric oxidation of 1.0mM (H3-tctpy)Ru(NCS)3] at 25C in acetonitrile (0.1M Bu4NPF6) at platinum electrodes. a) Cyclic voltammograms, 0.5mm radius electrode, =10, 25, 50, 100, 200mVs1; b) rotating disk electrode voltammogram, 1.5mm radius electrode, =20mVs1, f=500, 1000, 1500, 2000, 2500, 3000rpm; c) "log-plot" obtained from rotating disk voltammogram, f=500rpm.

The process appears to be chemically reversible in the sense that the current ratios (ipox/ipred) between the oxidation peak current (ipox) and the reduction peak current (ipred) approach values of unity when scan rates 50mVs1 are used as required for a reversible process. For a reversible

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 169

diffusion controlled process at a macrodisk electrode, the peak height is given by the RandlesSevcik equation (Eqn. 2.18 on page 43). As expected from this relationship, a linear plot of ip versus square root of the scan rate is obtained at platinum and glassy carbon electrodes which confirms that the [(H3-tctpy)RuII/III(NCS)3]/0 process is diffusion controlled in acetonitrile at these electrode surfaces. However, deviation from diffusion controlled behavior was found at gold electrodes at scan rates 500mVs1 where the oxidation peak currents were larger than predicted from data obtained at lower scan rates and use of Equation 2.18. The enhanced current at high scan rates is consistent with the presence of weak reactant adsorption [119, 189-191]. For a surface controlled process, the peak current scales linearly with the scan rate and hence becomes more evident at higher scan rates. Despite evidence for reactant adsorption at gold, the reversible halfwave potential calculated, is independent of scan rate, electrode material and concentration (see Table 6.1), which suggests that the contributions from adsorption in acetonitrile are negligible with respect to this calculation.
Glassy Carbon d=2mm Platinum d=1mm c=1.0mM scan rate Ep Ep Ep Er1/2 E Epox E Ep 1 [mVs ] [mV] [mV] [mV] [mV] [mV] [mV] [mV] [mV] [mV]
ox red

Gold d=1.5mm

Er1/2 314 316 316 318 319 318 318 318

ox

red

Epred [mV] 277 283 283 279 277 271 261 243

Er1/2 [mV] 311 314 314 316 319 321 323 329

E [mV] 68 62 62 74 84 100 124 172

10 26 50 100 200 500 1000 2000

349 349 353 357 367 369 381 401

279 283 279 279 271 267 255 235

70 66 74 78 96 102 126 166

345 345 345 350 357 365 371 385

279 283 283 283 279 275 267 251

312 314 314 317 318 320 319 318

66 62 62 67 78 90 104 134

345 345 345 353 361 371 385 415

c=0.2mM 10 26 50 100 200 500 1000 2000 349 345 353 357 369 395 385 395 279 283 285 291 289 281 273 257 314 314 319 324 329 338 329 326 70 62 68 66 80 114 112 138 343 337 343 339 349 357 353 395 261 285 279 279 283 283 277 261 302 311 311 309 316 320 315 328 82 52 64 60 66 74 76 134 345 339 345 349 357 377 401 265 281 283 285 283 277 267 305 310 314 317 320 327 334 80 58 62 64 74 100 134

Table 6.1. Cyclic voltammetric data obtained for the [(H3-tctpy)Ru(NCS)3]/0 oxidation process at 25C in acetonitrile (0.1M Bu4NPF6) as a function of electrode material and concentration. All potentials are reported vs. Fc/Fc+. E1/2 was calculated as (Epox + Epred)/2. Uncertainty of reported potentials 2mV.

Whilst not observed at macrodisk electrode experiments, evidence for significant surface interaction was also obtained during the oxidation of [(H3-tctpy)Ru(NCS)3] under slow scan rates near steady-state conditions at platinum and glassy carbon microdisk electrodes in acetonitrile. At potentials more positive than the initial oxidation process and after the current had reached its diffusional plateau, the current started to decay as expected when product adsorption gave rise to electrode blockage. However, this phenomenon did not markedly effect the rising part of the metal centered oxidation process because plots of "log-plots" were linear and had slopes of 652mV which is close to the theoretical value of 59mV at 25C as essential for a reversible one-electron

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 170

diffusion controlled transfer. From radial diffusion controlled microdisk electrode experiments and use of Equation 2.28 (page 48) the diffusion coefficient for [(H3-tctpy)Ru(NCS)3] in acetonitrile was calculated to be D=6.40.4106cm2s1. The oxidation of [(H3-tctpy)Ru(NCS)3] under the hydrodynamic steady-state conditions associated with rotating disk electrode experiments proceeded in an analogous manner to that observed with a microdisk electrode, but with less evidence of electrode blockage in the limiting current region. "Log-plots" obtained for the oxidation process from this technique over the electrode rotation rate of 500 to 3000min1 at glassy carbon and platinum electrodes were also linear and had slopes of 703mV. The limiting current for hydrodynamic voltammograms is given by the Levich equation (Equation 2.29 on page 49). Mass transport control for this process was demonstrated at glassy carbon and platinum electrodes via the linearly dependence of the limiting current versus the square root of the angular velocity and the plot passed through the origin. The slopes of the latter plots and use of Equation 2.29 gave a calculated diffusion coefficient of D=6.40.3106cm2s1 which is in excellent agreement with the value obtained from the microdisk electrode technique. complex [(H3-tctpy)Ru(NCS)3] solventa) acetonitrile acetone DMF DMF DMF DMF acetonitrile acetone DMF +935 oxidation Er1/2 [mV] 3168 2156 3055b) 276 reduction Er1/2 [mV] D [106cm2s1] 6.40.4 7.20.4 1.80.5 10 9.5

[(tctpyx)Ru(NCS)3](1+x) d) tctpy tctpyx d) (H2-tctpy)2Ru(NCS)2 e)

158520 15567c) 221510 202010b,c) 277010b,c) 211010 242015 15254 21545 250010

450 410 390 21545

[(H2x/2-dcbpyx/2)2Ru(NCS)2]x e)

Table 6.2. Reversible potentials (vs. Fc/Fc+) and diffusion coefficients at T=25C. a) 0.1M Bu4NPF6 electrolyte, b) T=55C, c) peak potential at =100mVs1, d) electrogenerated (see text) x=number of protons removed.e) data taken from Chapter 5.

The reversible oxidation potential of the less efficient (H2-tctpy)2Ru(NCS)2 sensitizer in acetonitrile is 450mV (Table 4.2 on page 125), that is 130mV more positive than the value measured in this study for [(H3-tctpy)Ru(NCS)3] (see Table 6.2). Simple thermodynamic considerations show that a negative shift of the potential for the oxidation of a sensitizer in photovoltaic cells should decrease the thermodynamic driving force for dye regeneration by the redox mediator, iodide (see Figure 5.1 on page 134). Clearly, the solar cell efficiency of [(H3-tctpy)Ru(NCS)3] is not solely determined by its thermodynamic properties.

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 171

1A 10A 2A

a)

c) e)

0.4A

5A 2A

b)

d) f)

0.5A

2A

10A

h) g)

i)

-0.5

0.0

0.5

1.0
+

-0.5

0.0

0.5

1.0
+

-0.5

0.0

0.5

1.0
+

potential [V] vs. Fc/Fc

potential [V] vs. Fc/Fc

potential [V] vs. Fc/Fc

Figure 6.4. Cyclic voltammograms on the oxidation of [(H3-tctpy)Ru(NCS)3] at 25C in acetonitrile (0.1M Bu4NPF6) as a function of electrode material, concentration and scan rate over a wide potential range. a) Ptelectrode, r=0.5mm, c*=1.0mM b) as for a, but c*=0.25mM; c) GC-electrode, r=1mm, c*=1.0mM d) as for c, but c*=0.25mM; e) Au-electrode, r=0.75mm, c*=1.0mM f) as for e, but c*=0.25mM g) Pt-electrode, r=0.5mm, c*=1.0mM, =10mVs1 h) as for g, but =200mVs1 i) as for h, but =5000mVs1.

(b) Processes at More Positive Potentials Figure 6.4 shows cyclic voltammograms over a larger positive potential range for the oxidation of [(H3-tctpy)Ru(NCS)3] in acetonitrile as a function of concentration and electrode material. Thus, the initial metal centered oxidation process described above is followed by two additional oxidation processes at about 0.6V and 1.0V, marked as processes B and C in Figure 6.4a and one reduction process on the reverse scan (process D). Since a RuIII/IV oxidation is not expected to occur within the potential window investigated [158], the processes observed in the more positive potential region are proposed to be associated with oxidation of surface attached material, as has been suggested in the case of the bipyridine analogue (H2-dcbpy)2Ru(NCS)2 (see Chapter 4). This assumption is supported by noting that increasing the scan rate (Figure 6.4g,h and i) and decreasing the concentration (Figure 6.4a and b) have similar effects on the relative prominence of the additional process. In case of increased scan rate, less time is available for kinetically controlled adsorption to occur. Similarly, the extent of adsorption is expected to decrease as is the case when the concentration is lowered [189-191]. A thiocyanate ligand based oxidation of [(H 3-tctpy)RuIII-

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 172

(NCS)3]0 does not seem likely, since the processes at positive potentials are absent when the initial metal centered oxidation process becomes reversible, for example at high scan rates or high rotation rates at a rotating disk electrode. However, under long time scale conditions, decomposition of the oxidized complex may occur. For example, in the case of (H2-dcbpy)2Ru(NCS)2 [81] and its ethyl ester (see Chapter 3.4.4 on page 83) it has been shown, that illumination or oxidation of the complexes can lead to loss of sulfur and formation of the cyanide analogue, (H2dcbpy)2Ru(CN)2. Analogous internal redox reactions also might occur after formation of [(H3tctpy)Ru(NCS)3]0 under long time scale conditions. Cyanide complexes are expected to be harder to oxidize than the thiocyanate species (see Chapter 3). Deprotonation after surface attachment is also possible, but this would be expected to give rise to processes at less rather than more positive potentials (Chapter 5.3.1 on page 148) as will be demonstrated below. (c) EQCM Studies of Surface Based Processes To confirm the presence of surface based reactions, studies were carried out at gold electrodes using the electrochemical quartz crystal microbalance (EQCM) technique and assuming that the Sauerbrey equation is valid [193].
1.0 -0.5 0.0 0.5 1.0 0.4

0.5

0.3

mass change [g]

0.0

current [A]

0.2

-0.5

0.1

-1.0

0.0

-1.5 -0.5 0.0 0.5


+

1.0

potential [V] vs. Fc/Fc

Figure 6.5. EQCM experiment at 25C in acetonitrile (0.1M Bu4NPF6) for oxidation of 0.05mM [LRu(NCS)3], =50mVs1, 2.5mm radius gold electrode. Lower curve is the cyclic voltammogram, while upper curve displays the mass change at the electrode surface that accompanies change of potential.

At open circuit potential, 45050ng [(H3-tctpy)Ru(NCS)3] was adsorbed onto a 5mm diameter gold electrode in acetonitrile (0.1M Bu4NPF6) when the dye was present in a concentration of 0.4mM. This equals to a surface coverage of =3109mol cm2 of [(H3-tctpy)Ru(NCS)3]Bu4N, which is similar to the surface coverage measured under similar conditions for the adsorption of (H2-dcbpy)2Ru(NCS)2 onto gold surfaces in acetone (see page 122), where the amount of adsorbed species was calculated to give a surface coverage of 30 monolayers. Thus, a significant amount of interaction with the surface occurs prior to the commencement of a voltammetric experiment when

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 173

a gold electrode is placed in acetonitrile solutions. EQCM experiments undertaken during the course of a cyclic voltammetric experiment at a gold electrode in acetonitrile are shown in Figure 6.5. When the scan is commenced at 0.6V no further mass change occurs until a potential of about 0V is reached when a slight mass loss is observed. Material continues to be lost until the potential for the reversible [(H3-tctpy)Ru(NCS)3]/0 process is reached. The onset of this metal based process prevents further mass loss occurring. However, a significant mass increase commences at 0.5V, which corresponds to a potential which is just beyond the first oxidation process. This mass increase continues until the potential of the third process at 0.9V (process C in Figure 6.4a) is reached, after which no further significant mass change occurs. On reversal of the potential scan direction the mass stays constant until 0.5V which coincides with the onset of the first process and then rapidly decreases again to reach a constant value at potentials less positive than the initial metal centered oxidation process. However, a net mass increase occurs during the course of one cycle. 6.3.2. Oxidation Studies in Acetone (a) The Metal Centered Oxidation Process If the switching potential is set to 0.5V, the initial metal centered oxidation process (Eqn. 6.1) is chemically reversible in acetone at all electrode materials under conditions of cyclic voltammetry or evidenced by the constant peak current ratio ipred/ipox of 1.10.1 over the scan rate range of 505000mVs1. The minor deviation from the expected value of unity for the current ratio is attributed to the presence of adsorption rather than irreversibility. Half-wave potentials were independent of electrode material, concentration, scan rate and rotation rate, as required for the determination of the reversible half-wave potential. Thus, the reversible half-wave potential for the [(H3-tctpy)Ru(NCS)3]/0 process in acetone is Er1/2=2156mV. Confirmation of a close to reversible one-electron process was obtained from noting that "log-plots" were linear and had values of 775mV (microdisk and rotating disk electrode). Levich plots (rotating disk electrode) were linear and enabled a diffusion coefficient of 7.20.4106cm2s1 to be calculated. In this solvent the largest difference (195mV, see Table 6.2) in reversible oxidation potentials between [(H3tctpy)Ru(NCS)3] and (H2-dcbpy)2Ru(NCS)2 is observed. As expected by this large potential shift away from a possible oxidation of the thiocyanate ligands (Chapter 4.3.4 on page 122), the chemical reversibility of [(H3-tctpy)Ru(NCS)3] is significantly increased compared to that observed for oxidation of (H2-dcbpy)2Ru(NCS)2, which is only moderately reversible under similar conditions (Chapter 4.3.1 on page 117). (b) Processes at More Positive Potentials If the switching potential is set at +1.2V, cyclic voltammograms contained two additional processes which are equivalent to processes B and C in Figure 6.4a. However, in this solvent no reductive process D is detected as is the case for acetonitrile. In acetone processes B and C show a very strong dependence on electrode material and concentration. Process B is clearly enhanced on

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 174

gold and glassy carbon electrodes relative to platinum electrodes. However, this process as well as process C are almost completely absent when high scan rates are employed (>1000mVs1), as would be expected when kinetically controlled surface based processes occur.

125 5 100

current [A]

current [A]

75 0 50 25 -5 0 -0.5 0.0 0.5


+

1.0

potential [V] vs. Fc/Fc

Figure 6.6. The effect of addition of an equimolar amount of SS-bpy on cyclic voltammograms and electrode material in acetone (0.1M Bu4NPF6), c*=1.0mM, =100mVs1. Dashed line without modifier; a) 2mm diameter glassy carbon electrode; b) 1mm diameter platinum electrode.

(c) Cyclic Voltammetry in the Presence of SS-bpy The gold electrode modifier 4,4'-bipyridyl disulfide (SS-bpy) was added to acetone solutions of [(H3-tctpy)Ru(NCS)3]. The addition of SS-bpy assists the detection of well defined voltammograms for cytochrome c [197] at gold electrodes and has been employed to suppress surface based processes in the voltammetry of the bipyridine analogue in the previous Chapter 4 (see page 125). SS-bpy forms electroinactive clusters on the gold electrode surface at low modifier concentrations and a monolayer coverage at high concentrations. When an equimolar amount of SS-bpy was added to a 1.0mM [(H3-tctpy)Ru(NCS)3] solution in acetone (0.1M Bu4NPF6), the disappearance of process B was noted in cyclic voltammograms obtained at platinum and gold electrodes (Figure 6.6b). No effect was observed on process B when the glassy carbon electrode was used (Figure 6.6a) and surface based process C remained unchanged on all three electrode materials. That is, SS-bpy successfully competes for some surface sites on the former two electrode materials, but its presence does not suppress all surface based activity. 6.3.3. Oxidation Studies in DMF Nazeeruddin and co-workers [88] provided a value of 0.6V vs. Ag/AgCl for the peak potential of the oxidation process of [(H3-tctpy)Ru(NCS)3](Et3HN) in DMF. However, these authors did not achieve voltammetric conditions that enable the reversible potential to be calculated.

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 175

scan rate [mVs1] 200 500 1000 2000 4000 8000 10 25 50 100 200 500 1000

Epox [mV] a) 280 280 296 304 328 370 336 334 336 340 348 364 396

Epred Er1/2 a) [mV] [mV] a) T=+25C 110 195 134 207 136 216 132 216 128 228 98 234 T=55C 270 303 274 304 274 305 270 305 264 306 244 304 210 303

E [mV] a) 170 146 160 172 200 272 66 60 62 70 84 120 186

Table 6.3. Cyclic voltammetric data for the oxidation of 1.0mM [(H3-tctpy)Ru(NCS)3] in DMF (0.1M Bu4NPF6) at T=+25C and T=55C. a) potential versus Fc/Fc+.

As in acetonitrile and acetone, the initial oxidation process under conditions of cyclic voltammetry was followed by two additional processes (Figure 6.7a). However, in DMF the initial metal centered oxidation process was not chemically reversible at a scan rate of 100mVs1 even when the switching potential was set +0.4V. At higher scan rates (200mVs1), whilst a reverse reduction peak was noticed, the E1/2-values calculated as (Epox + Epred)/2 were not constant over the scan rate range of 2008000mVs1 (shifted by about 50mV). Furthermore, a minor dependence of E1/2 on electrode material was observed. Table 6.3 summarizes some of the data obtained in DMF. These results lead to the conclusion that in DMF the oxidation does not fully correspond to a diffusion controlled [(H3-tctpy)Ru(NCS)3]/0 process as required for determination of the reversible potential of this process, although it is expected to be in the region of 200250mV vs. Fc/Fc+. In order to slow down side reactions that compete with the [(H3-tctpy)Ru(NCS)3]/0 process further voltammetric investigations in DMF were conducted at 55C. At this temperature the reversibility of the initial metal centered oxidation process was significantly increased, so that a reverse peak was observed even at a scan rate of 10mVs1. Cyclic voltammograms as a function of scan rate at 55C are shown in Figure 6.7b. The process becomes fully reversible at scan rates 200mVs1 where ipox/ipred reaches a constant value of 1.050.05. Half-wave potentials are constant over the scan rate range (101000mVs1) investigated and lead to a value of Er1/2=3055mV for the [(H3-tctpy)Ru(NCS)3]/0 process at 55C. This value coincided with Er1/2values calculated from microdisk and rotating disk electrode experiments. "Log-plots" obtained from these latter two techniques had slopes of 494mV (43mV theoretically expected at 55C) confirming the one-electron nature of this process at low temperatures in DMF.

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 176

a)
2A

0.3A

b)

-0.5

0.0

0.5
+

1.0

potential [V] vs. Fc/Fc

Figure 6.7. Cyclic voltammogram for oxidation of [(H3-tctpy)Ru(NCS)3] at 25Cin DMF. a) T=25C, c=0.6mM, 0.5mm radius platinum electrode, =100mVs1 b) T=55C, c*=1.0mM, 1mm diameter glassy carbon electrode, =10, 25, 50, 100, 200mVs1.

It is interesting to note the large chemical reversibility of the [(H3-tctpy)Ru(NCS)3]/0 process compared to that of [(H2-dcbpy)2Ru(NCS)2]0/+, in the latter case very high scan rates of over 100Vs1 are necessary in DMF before the onset of chemical reversibility becomes visible. This suggests a considerably increased lifetime of the oxidized [(H3-tctpy)Ru(NCS)3]0 sensitizer, which is advantageous for stability considerations in photoelectrochemical devices. 6.3.4. Voltammetric Oxidation of Mixtures of Thiocyanate Linkage Isomers The oxidation of solutions containing 42%, 77% and 100% thiocyanate linkage isomers other than the N,N,N- bond isomer (see Experimental) was studied in acetone. Cyclic voltammograms over the positive potential range were similar to those of the pure N,N,N- bonded isomer, except that the prominence of processes B and C, whose origins have been assigned to surface based processes, depended strongly on the isomer content and are shown in Figure 6.8. Whilst process B was absent in all solutions containing thiocyanate isomers (Figure 6.8b, c and d), process C was found to be enhanced in solutions containing 42% and 77% isomers, but also disappeared when the isomer content was increased to 100% linkage isomers (Figure 6.8d). This suggests that surface attachment of species giving rise to process B, and possibly also process C, may occur via electrode interaction with the sulfur of the isothiocyanate groups.

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 177

a)
4A

C A E B

b)

4A

c)
2A

d)
-0.5

1A

0.0

0.5
+

1.0

potential [V] vs. Fc/Fc

Figure 6.8. Cyclic voltammograms of isomeric mixtures of [(H3-tctpy)Ru(NCS)3] in acetone (0.1M Bu4NPF6) recorded at a 1mm diameter Pt disk electrode, =100mVs1. a) pure N,N,N- linkage isomer, * c =1.0mM; b) 42% other isomers, c*=1.2mM; c) 77% other isomers, c*=1.0mM; d) 100% other isomers, c*=0.4mM.

The half-wave potential for the metal based oxidation process was determined by cyclic and differential pulse voltammetry using a 1mM solution of each isomeric mixture in acetone (0.1M Bu4NPF6). Values obtained were 42%: E1/2=+2025mV, 77%: E1/2=+1625mV and 100%: E1/2=+15510mV. Thus, the potential for the initial metal based oxidation process is shifted to a less positive value as the content of sulfur bonded thiocyanate isomers increases. A differential pulse voltammogram of each solution is shown in Figure 6.9. When mixtures of isomers are present, the waveshapes deviate from those theoretically expected for a reversible process and E 1/2values therefore represent an average rather than a discrete value for one isomer. However, a reversible potential difference of 60mV is noted for the pure N,N,N- isomer and the fraction containing the highest content of sulfur bonded thiocyanate linkage isomer. The effect of the coordination mode of thiocyanate ligands in sensitizers on the voltammetry has been shown for (H2-dcbpy)2Ru(NCS)2 in Chapter 4.3.8 on page 129, where the predominant coordination mode of the thiocyanate ligands is also via the nitrogen. However, only solutions containing up to 35% isomers with S- bonded thiocyanate were investigated. No significant change in the reversible oxidation potential was noted in this case, but a decrease of surface based processes was found in cyclic voltammograms as is the case with [(H3-tctpy)Ru(NCS)3].

Chapter 6: A Novel Sensitizer: The Black Dye Oxidation: Results and Discussion

page: 178

Fc/Fc

77% other isomers 42% other isomers pure N,N,Nbonded isomer 100% other isomers

-100

100

200
+

300

potential [mV] vs. Fc/Fc

Figure 6.9. Differential pulse voltammograms for oxidation of [(H3-tctpy)Ru(NCS)3] at 25C in acetone (0.1M Bu4NPF6) as a function of % of linkage isomers other than N,N,N- bonded form. Scan rate 20mVs1, pulse height 50mV, 0.5mm radius platinum macrodisk electrode.

6.4. Reduction: Results and Discussion


6.4.1. Studies on the Reduction Process in DMF Nazeeruddin and co-workers [88] noted that the reduction of [(H3-tctpy)Ru(NCS)3] is complex under conditions of cyclic voltammetry. They reported two irreversible closely spaced waves (100mV apart) in the negative potential range in DMF and no conditions were found to allow the reversible potentials to be calculated. Interestingly, on the basis of comparison with the voltammetry of [Ru(terpy)2]2+ [150] only one reversible tctpy-ligand based reduction process is expected in the negative potential range investigated by Nazeeruddin and co-workers [88]. Studies carried out in the previous chapter on reduced forms of (H2-dcbpy)2Ru(NCS)2 showed that the favored decomposition pathway for the ligand reduced ruthenium complexes in this system involved an overall deprotonation and dihydrogen formation reaction. This reaction pathway was shown to be the origin of irreversible responses observed under standard voltammetric conditions. Clearly, analogous reactions may be anticipated to occur upon reduction of [(H3-tctpy)Ru(NCS)3]. Under conditions of cyclic voltammetry at a platinum electrode, only drawn out irreversible reduction responses were detected in the negative potential range in DMF (see Figure 6.10a). A simplified voltammetric response was observed at more negative potentials when glassy carbon was used as the electrode material instead of platinum (Figure 6.10b). Limiting the switching potential to 1.6V and increasing the scan rate improved the chemically reversibility of this process, but a complete reversible diffusion controlled process was never achieved. Consequently,

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 179

the reversible half-wave potential could only be estimated with a considerable level of uncertainty to be Er1/2=158520mV.

1A

a)

0.5A

b)
0.3A

c)
1A

d)
-2.5 -2.0 -1.5 -1.0
+

-0.5

potential [V] vs. Fc/Fc

Figure 6.10. Reduction of 1.0mM [(H3-tctpy)Ru(NCS)3] (a, b and c) and its deprotonated form (d) in DMF (0.1M Bu4NPF6) as a function of electrode material, temperature and scan rate. a) 1mm diameter platinum electrode, T=25C, =100mVs1 b) 1mm diameter glassy carbon electrode, T=25C, =100mVs1 c) 1mm diameter glassy carbon electrode, T=55C, =10, 25, 50, 100, 200mVs1 d) [(H3-x-tctpyx)Ru(NCS)3]x (deprotonated), c=0.6mM, 1mm diameter glassy carbon electrode, T=25C, =10, 25, 50, 100, 200mVs1.

Reducing the temperature to 55C enabled the reversible reduction of [(H3-tctpy)Ru(NCS)3] to be detected. Under conditions of cyclic voltammetry (Figure 6.10c) at this temperature and when scan rates 50mVs1 and glassy carbon electrodes were used, ipox/ipred was close to 1.0 and as required Er1/2-values were scan rate (cyclic voltammetry) and method (cyclic voltammetry, microdisk and rotating disk electrode) independent. Er1/2 values calculated by the various techniques are found in Table 6.2. Slopes of "log-plots" for steady-state microdisk and rotating disk electrode experiments confirmed the one-electron nature of the process (Table 2). If this initial reduction process is H3-tctpy ligand based, as is known to be the case with the bipyridine analogue, the initial charge transfer processes may be formulated as in Equation 6.2 below: [(H3-tctpy)Ru(NCS)3] + e [(H3-tctpy)Ru(NCS)3]2 (6.2)

6.4.2. Controlled Potential Reduction and Deprotonation of [(H3-tctpy)Ru(NCS)3] in DMF In order to elucidate the nature of reaction pathways involved in the voltammetric reduction of [(H3-tctpy)Ru(NCS)3], controlled potential reductive electrolysis experiments were initially carried out in DMF (0.1M Bu4NPF6) by applying a potential of Eappl=2.0V (25C) at a platinum

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 180

gauze working electrode. The exhaustive reductive electrolysis of [(H3-tctpy)Ru(NCS)3] consumed 2.40.3 electrons under these conditions. Gas chromatographic analysis of the collected gas phase showed molecular hydrogen as being the major gaseous product generated by reductive electrolysis. The peak intensity of the hydrogen signal was close to that expected if the majority of electrons transferred were used for the formation of hydrogen. This information enables the deprotonation reaction in Equation 6.3 to be proposed. [(H3-tctpy)Ru(NCS)3] + 3 e [(tctpy3)Ru(NCS)3]4 + 1.5 H2 (6.3)

For the sake of clarity, tctpy3, H-tctpy2 and H2-tctpy will be used throughout the remainder of this chapter for the triply, doubly and singly deprotonated H3-tctpy ligand, respectively. In contrast, H3-tctpy will be used for the singly reduced but fully protonated H3-tctpy ligand. Mixed forms of the ligand (deprotonated and reduced) will be abbreviated accordingly. The overall charge of the complexes is given outside the square brackets. Clearly, an extensive sequence of electron transfer and chemical reactions can occur at the platinum surface which involves steps where H+ is formed by deprotonation before or after an electron transfer reaction and H2 is generated heterogeneously at the electrode surface or homogeneously by solution phase reactions. One example of heterogeneous generation of H2 that give the overall reaction in Equation 6.3 would be the reaction sequence in Equations 6.4 to 6.7: [(H3-tctpy)Ru(NCS)3] [(H2-tctpy)Ru(NCS)3]2 [(H-tctpy2)Ru(NCS)3]3 [(tctpy3)Ru(NCS)3]4 1.5 H2 + H+ + H+ + H+ (6.4) (6.5) (6.6) (6.7)

[(H2-tctpy)Ru(NCS)3]2 [(H-tctpy2)Ru(NCS)3]3

3H+ + 3e

or if hydrogen is generated by homogeneous reactions after electron transfer, then Equations 6.8 to 6.10 would represent one of many examples of this kind of reaction pathway that give the overall Equation 6.3. [(H3-tctpy)Ru(NCS)3] 3 [(H3-tctpy)Ru(NCS)3] [(tctpy3)Ru(NCS)3]4 + 3H+ (6.8) (6.9) 1.5 H2 (6.10)

+ 3 e

3 [(H3-tctpy)Ru(NCS)3]2 +

3 [(H3-tctpy)Ru(NCS)3]2 + 3 H+

3 [(H3-tctpy)Ru(NCS)3]

These sequences of reactions are possible because the protonated [(H3-tctpy)Ru(NCS)3] sensitizer may exist in an acid-base equilibrium (Eqns. 6.4 and 6.8), the extent to which deprotonation occurs being dependent on solvent (medium) specific pKa-values. Thus, it needs to be noted that in the presence of water, adventitiously present or deliberately added, thermodynamic (and/or kinetic) limitations will be placed on the extent to which reduction can occur via the equilibrium prior of the reaction

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 181

[(tctpy3)Ru(NCS)3]4 + y H2O [(Hy-tctpy(3y))Ru(NCS)3](4y) + y OH

(6.11)

Thermodynamically, protons can be very easily reduced to hydrogen, although the suppression of the heterogeneous reaction pathway can be achieved by choice of an appropriate electrode material at which the rate of this reaction is electrochemically negligible at the reversible potential [95]. Glassy carbon is an electrode material where the over-potential for the reduction of protons to hydrogen is very large. In contrast, the over-potential for the generation of hydrogen at platinum electrodes is small. Furthermore, reduced ruthenium polypyridine compounds are known to catalytically reduce protons or residual water to hydrogen in organic solvents [55]. Thus, the reaction sequence described in Equations 6.9 and 6.10 might be a source of hydrogen generation, at least under long time scale conditions. However, the hydrogen producing step in Eqn. 6.10 is based on a purely homogeneous solution phase reaction and does not predict the experimentally observed influence of electrode material. In contrast, the electrochemical reduction of protons (Eqn. 6.7), made available by the acid-base equilibrium described in Eqns. 6.4, 6.6 and 6.8 occurs at the electrode surface and hence heterogeneous electrode kinetics may be involved in the reaction sequence which could explain the electrode material dependent voltammetric responses observed for the reduction of [(H3-tctpy)Ru(NCS)3]. Overall Equation 6.3 predicts the transfer of 3.0 electrons per molecule if complete deprotonation occurs to quantitatively yield [(tctpy3)Ru(NCS)3]4, whereas the number of electrons transferred was only 2.40.3. Hence, this equation and the mechanism given as examples cannot provide a complete description of the process. The pKa values of [(H3-tctpy)Ru(NCS)3] are unknown, but by comparison to (H2-dcbpy)2Ru(NCS)2 (see last chapter and Chapter 4.3.9 and Table 4.3 on page 130) it is expected that [(H3-tctpy)Ru(NCS)3] is a rather weak acid. Finally, it needs to be noted that since residual water is present, the reaction sequence [(H3-tctpy)Ru(NCS)3] + [(H3-tctpy)Ru(NCS)3]2 + H2O e [(H3-tctpy)Ru(NCS)3]2 [(H3-tctpy)Ru(NCS)3] + OH + 0.5 H2 [(H2-tctpy)Ru(NCS)3]2 + H2O (6.12) (6.13) (6.14)

[(H3-tctpy)Ru(NCS)3] + OH would give the overall reaction [(H3-tctpy)Ru(NCS)3] + 2 H2O + 3 e

+ 1.5 H2 (6.15)

[(H-tctpy)Ru(NCS)3]2 + 2 OH In summary, this reaction with water can be generally written as [(H3-tctpy)Ru(NCS)3] + (x1)H2O + xe

[(H2-tctpy)Ru(NCS)3]2 + (x1)OH +

x 2

H2

(6.16)

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 182

This reaction sequence and predominant formation of the singly deprotonated form of the dye is also consistent with data (vide infra) obtained in this study, although it will emerge later in this chapter that the doubly deprotonated dye also may have been partly formed. Clearly, a range of combination of electron and proton transfer steps are possible (see above) which lead to deprotonation. However, the reaction scheme can be very complex and formation of doubly and triply deprotonated [(H3-tctpy)Ru(NCS)3] complexes also can occur as can potential dependent combination of reaction pathways. (a) Voltammetric Oxidation of [(H3xtctpyx)Ru(NCS)3](1+x) Cyclic voltammograms in DMF of the oxidation process of electrogenerated [(H3xtctpyx)Ru(NCS)3](1+x) as a function of scan rate are shown in Figure 6.11a.

a)
0.5A

b)
5A

-0.6

-0.4

-0.2

0.0
+

0.2

potential [V] vs. Fc/Fc

Figure 6.11. Oxidation of electrogenerated deprotonated [(Hx3-tctpyx)Ru(NCS)3]x in DMF (0.1M Bu4NPF6). a) Cyclic voltammetry, 1mm diameter glassy carbon electrode, =10, 25, 50, 100, 200mVs1 b) rotating disk electrode, 3mm diameter glassy carbon electrode, =20mVs1, f=500, 1000, 1500, 2000, 2500min1.

Under conditions of cyclic voltammetry, the oxidation process is significantly more reversible at T=25C than the oxidation process of the fully protonated complex (see above). Furthermore, the potential for this process is shifted by about 0.3V to more negative potentials (Er1/2=0.03V) which reduces the thermodynamical driving force for a possible thiocyanate ligand oxidation. This shift in potential upon deprotonation is similar to the value observed for the bipyridine analogue (see Chapter 5.3.1e on page 148). The oxidation process of [(H3x-tctpyx)Ru(NCS)3](1+x) in DMF becomes fully reversible at scan rates above 25mV, where constant ipox/ipred-values of 1.2 were obtained. The deviation from unity is again attributed to the presence of adsorption. As required for a reversible process, Er1/2-values obtained from cyclic voltammetric measurements were indepen-

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 183

dent of scan rate over the range investigated (102000mVs1) and coincided with values acquired from microdisk and rotating disk electrode experiments (see Figure 6.11b). Slopes of "log-plots" (microdisk and rotating disk electrode) were 755mV and confirmed the one-electron nature of the process. From rotating disk and microdisk electrode measurements a diffusion coefficient of 1.90.3106cm2s1 was calculated for [(H3x-tctpyx)Ru(NCS)3](1+x) in DMF. (b) Voltammetric Reduction of [(tctpyx)Ru(NCS)3](1+x) Similar to observations made on the oxidation process of [(H3x-tctpyx)Ru(NCS)3](1+x) in DMF, cyclic voltammetric measurements after exhaustive electrolysis at Eappl=2.2V vs. Fc/Fc+ and deprotonation, indicated a reduction process located at more negative potentials and with a higher degree of reversibility than the fully protonated form. Cyclic voltammograms as a function of scan rate are presented in Figure 6.10d. The process assigned to the deprotonated terpyridine ligand based reduction becomes fully reversible at scan rates 500mVs1 where a constant ipox/ipred-value of 1.2 was approached. The deviation from unity may be explainable by the capability of reduced ruthenium polypyridine complexes to catalytically react with residual water to give the less reduced form [55]. This was confirmed by employing the DigiSim 3.0 voltammetric simulation package and by entering a simple catalytic square scheme corresponding to the aforementioned mechanism where almost constant ipox/ipred-values of 1.2 could be readily achieved over a large scan rate range (10 10000mVs1) by using appropriate rate constants. Furthermore, the simulations also showed that shifts of the Er1/2-value calculated as (Epred+Epox)/2 were negligible under these conditions. This was confirmed experimentally by noting that Er1/2-values were constant over the scan rate range investigated (265000mVs1) and "log-plots" obtained from microdisk and rotating disk electrode measurements had slopes of 726mV which is close to the expected theoretical value for an oneelectron process. From a Levich plot (rotating disk electrode) on this process a diffusion coefficient of 1.70.3106cm2s1 was calculated, which is in good agreement with the value obtained using the oxidation process. With an Er1/2-value of 2.22V, this process is located 0.63V more negative than the corresponding process of the fully protonated complex. (c) Molecular Orbital Calculations The shifts of redox potentials of [(H3-tctpy)Ru(NCS)3] upon deprotonation of the terpyridine ligand are readily explained by considering the electronic influence of the carboxylic acid groups on the ligand. This effect has been studied in the previous chapter by molecular orbital calculations on the bipyridine analogue [144] where the LUMO can be lowered by up to 0.9V by introduction of one carboxylic acid group on each pyridine ring (see Figure 5.5 on page 147). When a reduction process occurs, the LUMO represents the redox orbital in to which the electron is placed and thus, the relative position of the potential for the reduction process directly reflects the relative position of the LUMO and allows correlation between theoretical and experimental results. For instance, Haga and co-workers [135] could successfully predict the effect of deprotonation of a ruthenium polypyridyl complex from molecular orbital calculations.

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 184

1.0

6* 5* 4* 3* 2* 1*

orbital energy [eV]

0.5 0.0


-0.5

-1.0

LUMO
tpy

tctpy

H-tctpy

H2-tctpy

H3-tctpy

Figure 6.12. Molecular orbital energies obtained from MNDO-PM3 calculations for tpy, H3-tctpy and its deprotonated forms. tpy = 2,2':6',2''-terpyridine, 1* = LUMO, 2* = LUMO+1,and so forth.

On the basis of these considerations, molecular orbital calculations were conducted on the H3tctpy ligand and its deprotonated forms, which define the reduction orbitals of the ruthenium sensitizer. The results are schematically presented in Figure 6.12. The initial effect of introducing carboxylate groups on the terpyridine ligand results in a decrease of the LUMO energy by 0.55eV which is similar to what has been observed for bipyridine (see Figure 5.5 on page 147 and reference [144]). However, the removal of one proton increases the LUMO of H2-tctpy by 0.73eV already above the value of unsubstituted terpyridine, thus compensating the electron withdrawing effect of the remaining carboxylate groups. Subsequent deprotonation yielding H-tctpy2 and tctpy3 increases the LUMO by 1.23eV and 2.02eV, respectively, compared to H3-tctpy. If these theoretical data are now correlated with experimental results where a potential shift of 0.63V has been measured for the reduction process occurring upon electrochemical deprotonation of [(H 3tctpy)Ru(NCS)3], the conclusion would be reached that only one proton is removed from the complex, thus, yielding [(H2-tctpy)Ru(NCS)3]2 as the final product obtained from exhaustive reductive electrolysis. 6.4.3. Spectroelectrochemical Measurements The electronic spectrum of [(H3-tctpy)Ru(NCS)3] shows two bands in the visible region at 23.1 and 15.7103cm1 which are tentatively assigned to MTLC bands arising from transitions of a ruthenium d electron into a * orbital of the tctpy ligand. The lower energy band has a shoulder reaching down into the near infra red region of the electronic spectrum to about 10000cm1, hence absorbing over the whole visible range. Two additional bands in the UV region of the spectrum are assigned to intra ligand transitions of the tctpy ligand. The free tctpy ligand shows one strong band in this region (Table 6.4) and the shoulder of a second band at higher energies which is not fully detectable in DMF. Hence, it appears that the tctpy intra ligand bands shift to lower energy values in the ruthenium complex.

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 185

absorbance [10 M cm ]

-1

3.0 2.5 2.0 1.5 1.0 0.5 0.0

a)

-1

10000

20000 30000 -1 energy [cm ]

40000

absorbance [10 M cm ]

-1

3.0 2.5 2.0 1.5 1.0 0.5 0.0

-1

b)

10000

20000 30000 -1 energy [cm ]

40000

Figure 6.13. Electronic spectra obtained during the course of OTTLE experiments on the reductive electrolysis of 0.4mM [(H3-tctpy)Ru(NCS)3] in DMF (0.1M Bu4NPF6) at 22C. a) Deprotonation at Eappl=2.2V and formation of [(H3x-tctpyx)Ru(NCS)3](x+1); b) further reduction of deprotonated[(H3xtctpyx)Ru(NCS)3](x+1) at Eappl=2.5V.

When a solution of [(H3-tctpy)Ru(NCS)3] in DMF is reductively bulk electrolyzed at Eappl=2.2V, all bands experience a shift to higher energy values (Figure 6.13a). The electrolysis is accompanied by a distinctive color change of the solution from dark green to sky blue. The spectrum of the starting solution could be regenerated without noticeable loss when a small amount of HBF4 was added to the electrolyzed solution, which furthermore confirms the deprotonation mechanism. It is noted, that no isosbestic points are formed during the electrolysis (Figure 6.13a), although deviations from ideal isosbestic points are small. This would be expected when predominantly the singly deprotonated [(H2-tctpy)Ru(NCS)3]2 complex is formed and deviation from isosbestic points is caused by a change in solution environment accompanying the electrolysis, for example triggered by an increase of the hydroxyl anion concentration during the electrolysis according to Equation 6.16. This assumption is supported by titration experiments. When a solution of [(H3-tctpy)Ru(NCS)3] in DMF was titrated with 4% Bu4N(OH) solution and the course of titration monitored by an UV/VIS arrangement, the results were almost identical to those recorded during a reductive OTTLE experiment (Figure 6.13a). That is, the final spectrum obtained was identical to that obtained from the OTTLE experiment and no isosbestic points were observed during titration. Alternatively, reductive electrolysis of [(H3-tctpy)Ru(NCS)3] also may have given rise to formation of small amounts of doubly deprotonated [(H-tctpy2)Ru(NCS)3]3 leading to mixtures of [(H2-tctpy)Ru(NCS)3]2 and [(H-tctpy2)Ru(NCS)3]3.

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 186

complex

L L*

MLCT MdL*

energy [103cm1] (molar extinction [103M1cm1]) [(H3-tctpy)Ru(NCS)3] [(tctpyx)Ru(NCS)3](1+x) a) tctpy tctpyx a) 34.0 (28.4) 35.8 (28.6) 30.6 (24.2) 30.7 (27.2) 33.2 (17.7) 33.8 (17.4) 28.6 (21.1) 29.6 (29.8) 23.1 (13.6) 25.3 (10.5) 15.7 (8.4) 16.9 (8.4)

Table 6.4. Spectroscopic data recorded in DMF (0.1M Bu4NPF6) at +25C. a) electrogenerated (see text) x=number of protons removed.

The shift of the MLCT bands is as expected from voltammetric data obtained on the shifts of the reversible potentials of the ligand based reduction and the metal centered oxidation upon deprotonation. Since the reduction potentials change by 0.63V and the oxidation potentials only by 0.3V, the ligand *-orbitals have increased to a greater extent than the metal's d-orbitals. Thus, the MLCT, which can formally be written as a Ru(d)H2-dcbpy(*) transition, is expected to increase in energy. Interestingly, deprotonated forms of the [(H3-tctpy)Ru(NCS)3] sensitizer and analogues are commonly employed in TiO2 photoelectrochemical cells and found to exhibit higher energy conversion efficiencies [203] than the protonated form. This fact is surprising, since the spectral properties of the sensitizer are adversely effected when it is deprotonated, the light harvesting MLCT bands are shifted toward the UV region and some bands additionally decrease in absorbance. From the number of electrons per molecule transferred in controlled potential electrolysis experiments (see above) it is known, that the number of protons removed from [(H3-tctpy)Ru(NCS)3] by reductive electrolysis is less than three. To attempt an even more extensive deprotonation of electrogenerated [(H3x-tctpyx)Ru(NCS)3](1+x), an additional controlled potential electrolysis experiment was carried out at more negative potentials at Eappl=2.5V (Er1/2=2.22V for reduction of [(H3x-tctpyx)Ru(NCS)3](1+x)) and the course of the reaction was monitored in an OTTLE arrangement to give the result shown in Figure 6.13b. Distinctive changes were noticed during this electrolysis in the electronic spectrum. For example, bands at 12000cm1 grew in, whereas the intra ligand based transition at 30000cm1 completely disappeared, while the maintenance of seven isosbestic points was noted. The changes observed are consistent with a ligand based reduction rather than further deprotonation, as has been observed for reduction of deprotonated forms of (H 2dcbpy)2Ru(CN)2 (see Figure 5.11 on page 156). However, re-oxidation by applying a potential of Eappl=1.0V did not lead to regeneration of the spectrum of the starting compound, so that reversible ligand based reduction did not occur on this time scale. Experiments conducted at T=55C and at slightly more positive potentials, gave identical data. These results imply that a chemical reaction takes place after the initial charge transfer and that further deprotonation leading to the formation of fully deprotonated [(tctpy3)Ru(NCS)3]4 is not possible by controlled potential electrolysis.

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 187

6.4.4. Reduction and Deprotonation of the H3-tctpy ligand in DMF The uncarboxylated analogue of the tctpy ligand, terpy (2,2':6',2''-terpyridine), exhibits a single one-electron reduction process [155] in acetonitrile. The process has been reported to be partly reversible [150] and occurs at a potential of Er1/2=2.55V vs. Fc/Fc+ (corrected from SSCE reference potential by assuming Fc/Fc+=0.38V vs. SSCE [175]). Voltammograms for the reduction of H3-tctpy in DMF at 25C were broad and highly irreversible at both platinum and glassy carbon electrodes. A cyclic voltammogram recorded at 55C in DMF (0.1M Bu4NPF6) using a glassy carbon electrode is shown in Figure 6.14a. One well defined reduction process is seen, which is irreversible up to scan rates of 5000mVs 1 and has a peak potential of Epred=2.02V (=100mVs1). Controlled potential bulk reductive electrolysis of a 1mM H3-tctpy solution in DMF was undertaken. The potential was set to Eappl=2.1V and a platinum gauze working electrode was used (T=25C). The exhaustive electrolysis consumed 2.50.3 electrons per molecule, which is similar to that obtained for the [(H3-tctpy)Ru(NCS)3] complex (see above). Gas chromatographic measurements of the supernatant gas phase also indicated the formation of dihydrogen during the course of the electrolysis.

a)
0.2A

b)
0.2A

-3.0

-2.5

-2.0

-1.5

-1.0
+

-0.5

potential [V] vs. Fc/Fc

Figure 6.14. Cyclic voltammograms obtained for the reduction of 1.0mM H3-tctpy in DMF (0.1M Bu4NPF6) and its electrogenerated deprotonated form. T=55C, =100mVs1, 1mm diameter glassy carbon electrode a) protonated complex b) deprotonated ligand after exhaustive electrolysis at Eappl=2.0V.

A cyclic voltammogram of the bulk reductively electrolyzed solution recorded at 55C is shown in Figure 6.14b. One process arising from reduction of a deprotonated form is observable at a potential 0.75V more negative than the reduction of the initial H3-tctpy ligand. The reduction process is irreversible (Epred=2.77V, =100mVs1) as was the case with reduction of the starting

Chapter 6: A Novel Sensitizer: The Black Dye Reduction: Results and Discussion

page: 188

compound. Caution is needed when comparing peak potentials of irreversible processes. However, the significant shift in reduction potential that occurs when the H3-tctpy ligand is deprotonated is probably predominantly explicable in terms of the electronic influence of the carboxylate groups on the parent terpy ring system and the data obtained from molecular orbital calculations (see Figure 6.12). The experimentally observed shift of reduction peak potential by about 0.75V is in agreement with molecular orbital data and removal of only one proton during reductive electrolysis yielding [terpy(COOH)2(COO)]. 6.4.5. In Situ Reductive OTTLE Experiments on the H3-tctpy Ligand The course of a reductive electrolysis at Eappl=2.0V in DMF (T=22C) was monitored spectroelectrochemically with an OTTLE arrangement. H3-tctpy exhibits one band in the UV-region at 33.2103cm1 and the course of deprotonation yields a compound with a very similar spectrum which also shows only one band in the UV-region at 33.8103cm1 (see Table 6.4 for more details).

6.5. Conclusions
The reversible potential for the [(H3-tctpy)Ru(NCS)3]/0 oxidation process has been measured in a variety of solvents using different electrochemical techniques and electrode materials. Reactant and product interaction with the electrode surface cause departures from the mass transport controlled oxidation process for this compound. The potential for the oxidation process is solvent dependent, and located 0.150.20V less positive than the reversible potential for oxidation of the (H2-dcbpy)2Ru(NCS)2 sensitizer. Increased reversibility of the [(H3-tctpy)Ru(NCS)3]/0 oxidation process is also observed compared to that of the bipyridine analogue. Table 6.2 summarizes data obtained for the two systems. This improved stability of the oxidized form of this sensitizer is advantageous in terms of its long time stability in photovoltaic cells. The degradation of the oxidized sensitizer that occurs during an electricity producing cycle becomes the limiting factor in determining the life expectation of a dye sensitized solar cell and has commonly been recognized as one of the few shortcomings of these kind of solar cells [225]. However, the "striking performance" [88] of [(H3-tctpy)Ru(NCS)3] compared to that of (H2-dcbpy)2Ru(NCS)2 cannot simply be related to its thermodynamic properties, since no enhanced cell performance is expected if the reversible potential for the oxidation of the sensitizer is lowered. Indeed, the thermodynamic driving force required for the dye regeneration by the electrolyte, I, will decrease. Thus, it appears that in this case the excellent performance of [(H3-tctpy)Ru(NCS)3] in solar cells is related to its photophysical properties and efficient sensitization of TiO2 electrodes over the whole visible spectrum rather than to its electrochemical properties. Furthermore, it is noted that thiocyanate linkage isomers of [(H3-tctpy)Ru(NCS)3] with a higher number of sulfur bonded thiocyanates are easier to oxidize than the pure N,N,N- bonded linkage isomer. The prominence of surface interactions present in voltammetric experiments is also

Chapter 6: A Novel Sensitizer: The Black Dye Conclusions

page: 189

decreased when the thiocyanate ligands are coordinated to the ruthenium center via their sulfur atom rather than via the nitrogen, which suggests that part of the surface attachment occurs by sulfurelectrode surface interactions. However, new synthetic methods need to be developed to provide pure forms of specific isomers to fully confirm the significance of data related to linkage isomers. The reduction of [(H3-tctpy)Ru(NCS)3] in DMF under standard voltammetric conditions was irreversible. Reduced temperature, increased scan rates and glassy carbon electrodes were necessary to obtain one reversible one-electron tctpy ligand based reduction process. The overall reaction pathway for reduced forms of H3-tctpy and [(H3-tctpy)Ru(NCS)3] was found to be deprotonation coupled with dihydrogen formation. Under conditions of controlled potential reductive bulk electrolysis deprotonated forms of H3-tctpy and [(H3-tctpy)Ru(NCS)3] could be produced, although not all protons were removed. Molecular orbital data suggest that [(H2tctpy)Ru(NCS)3]2 should be formed as the major product after reductive bulk electrolysis, whereas electronic spectra obtained during the course of electrolysis suggest that small amounts of the doubly deprotonated complex, [(H-tctpy2)Ru(NCS)3]3, also may have been formed. In general, [(H3-tctpy)Ru(NCS)3] behaves in an analogous manner to (H2-dcbpy)2Ru(NCS)2. Thus, reduction and deprotonation generates new forms of the sensitizer, with the reduction potentials of the deprotonated form shifted to more negative values than the fully protonated complex. The potential for reduction shifts negative by 0.7V and that of the oxidation process negative by 0.3V. However, it still remains unclear as to why performance of deprotonated forms of this sensitizer is superior to the protonated form. Both MLCT bands experience a shift to the UV region and partly decrease in intensity which is commonly considered as being detrimental to cell performance. An extensive discussion on this effect and its relation to performance of sensitizers in real solar cells is provided in Chapter 5.4.2 on page 161 and in Chapter 8.3 on page 218.

Chapter 7

part of this work has been published in

A Channel Flow Cell System Specifically Designed to Study the Efficiency of Redox Shuttles in Dye Sensitized Solar Cells Alan M. Bond, John C. Eklund, Doug MacFarlane, Georg Wolfbauer, Presented at the Twelfth International Conference on Photochemical Conversion and Storage of Solar Energy, 1998, Berlin, Germany, poster 1W36

Chapter 7: The Efficiency of Redox Shuttles Introduction

page: 191

7.1. Introduction
The important solution phase electrochemical properties of sensitizers employed in photo electrochemical cells (PEC) have been investigated in detail in the previous chapters. Knowledge of the thermodynamic properties and decomposition pathways associated with oxidation and reduction processes is essential for describing and understanding the properties of dye sensitized solar cells. However, as shown in Figure 7.1, the electrochemical properties of the redox mediator, which carries (shuttles) the charge between the counter and working electrode to reduce the oxidized dye back to its initial state, is at least of equal importance.

A
e
fermi level

S+/S*

e n e rg y

Eoc R/R

S+/S TiO2
electrolyte solution

Figure 7.1. Simplified diagram showing the electron transitions in a photoelectrochemical solar cell during operation. S .... sensitizer (for structure see right), R .... electrolyte (commonly I /I3), Eoc .... open circuit potential.

According to the simplified scheme presented in Figure 7.1, the open circuit potential (Eoc) is basically only determined by the position of the TiO2 semiconductor fermi level and the chemical potential of the redox shuttle used. Thus, the reversible potential of the redox shuttle determines the overall efficiency of the cell (see chapter 1.6.5(b) on page 14). Since Tsubomura and co-workers first report [28] on nano-crystalline dye sensitized solar cells, no redox electrolyte system other than the iodide/triiodide one appears to have been used in these kind of solar cell.13 In contrast, ferrocene has been successfully used in non-dye sensitized solid state semiconductor solar cells [226]. The iodide system, which consists only of the basic element iodine, might be seen as an intrinsically simple system. This may account for the lack of information available in the literature, although in reality it is highly complicated (vide infra). However, the absence of information available on redox shuttles is more than surprising when the enormous amount of effort spent in designing and tuning sensitizer compounds is considered.

13

Bach and co-workers have recently (1998) reported [51] the construction of a dye sensitzed solar cell where the charge shuttle process was performed by a thin film of a solid organic spiro compound, which acts as a kind of conductor between the counter electrode and the oxidized dye. This approach is clearly different to the solution/redox mediator system and efficiencies reported so far are very low.

counter electrode

Chapter 7: The Efficiency of Redox Shuttles Aim of the Work

page: 192

7.2. Aim of the Work


It is proposed that a range of alternative potential redox shuttles, such as the ferrocene/ferrocenium system need to be identified, and tested for performance under real or cell-like conditions. The manufacturing process required to construct actual sample solar cells is difficult and time consuming. Since, the testing process requires variation of many parameters (i.e. concentration, solvent, etc.), the assembly of many cell units would be required. As an alternative way of efficient testing of redox shuttles, it is proposed that an on-line flow system can be constructed in which solutions of redox shuttles can flow sequentially over a single dye sensitized TiO2 working electrode. This flow-through cell method allows rapid measurements to be conducted without the need to assemble a set of cells each time a parameter needs to be changed.

7.3. Experimental
All solvents used were of HPLC grade (Mallinckrodt, ChromAR and UltimAR). The description of other reagents used can be found in Chapter 3.2 on page 62. Ferrocene, tetraalkylammonium salts and other chemicals were used as received from Aldrich or BDH and were at least of p.a. grade. Conventional voltammetric measurements were carried out using a standard three-electrode cell arrangement. Specific details can be found in Chapter 3.2 on page 62. Voltammetric results obtained with this setup are referenced versus the potential of the ferrocene/ferrocenium process (Fc/Fc+), which was measured by voltammetric oxidation of Fc under identical conditions (solvent, electrolyte, electrode, ....) as used in the relevant experiments. Voltammetric measurements utilizing the channel flow cell were conducted with a platinum quasi reference electrode combined with a MacLab/4e (ADInstruments) potentiostat or PAR 175 waveform generator and an AMEL 551 (Milano, Italy) potentiostat coupled to a MacLab/8e analoguedigital converter. All experiments with the channel flow system were undertaken under potentiostatic control to avoid complications caused by reactions occurring at the counter electrode. Between measurements the working electrode was left at the open circuit potential or at the potential of zero current (PZC). The latter potential was strongly dependent on solution environment. The PZC had to be chosen by careful manual adjustment, since if the potential became too positive, the sensitizer could be oxidized, and if too negative, SnO2 and TiO2 could be reductively dissolved (vide infra). Thus, only a few hundred millivolts of potential is available for stable operation. Problems associated with locating the PZC were enhanced because of the use of a quasi reference electrode. At the end of a measurement, the cell was cleaned by flushing with pure CH3CN. Additional details of the channel flow system experimental arrangement can be found later in the text. The temperature at which all experiments were undertaken was 20(2)C.

Chapter 7: The Efficiency of Redox Shuttles Experimental

page: 193

The flow rate for channel flow measurements was controlled by gravity; the solution reservoir was placed above the cell and was variable in height as was the outlet. Additional control over the flow rate was obtained by incorporating one of three capillaries of different diameter into the flow. Calibration was undertaken for each capillary and each solvent to enable calculation of the flow rate during experiments. An Oriel 1000W lamp equipped with a monochromator (300800nm) and a lens rail bench system were used as the light source. The monochromator was adjusted for maximum intensity resulting in slit widths of ca. 2030nm. The light intensity was estimated by using a calibrated silicon solar cell. Monochromatic light intensities used were in the range of 14mWcm2. For scanning electron microscopy, an Hitachi S5000 in-lens field emission scanning electron microscope (IFESEM) was used. Samples were coated with iridium in a Gatan ion-beam coater to prevent electrostatic charging in the IFESEM electron beam. Powder X-ray diffraction (XRD) measurements of TiO2 samples were performed with a Phillips PW 1729 diffractometer and Cobalt radiation. 10cm 10cm double layer titaniumoxide coated (810m thickness) conductive glass substrates (fluorine doped SnO2) were received from ANSTO, Sydney. TiO2 plates of smaller sizes were obtained by cutting the plates with a diamond cutter. Excess TiO2 was removed with a razor blade, whilst excess conductive layer of SnO2 was removed by scratching with a diamond glass cutting disk. The dye was adsorbed onto the TiO2 by soaking the TiO2 plates in a ca. 15104M solution of (H2-dcbpy)2Ru(NCS)2 in ethanol for 124 hours. Adsorption could be detected visually by a dark purple coloring of the TiO2 surface. Samples of [Cr(CO)5I](C4H9)4N, cis-Mo(CO)2(dpe)(dpm), trans-Cr(CO)2(dpe)2 and cisCr(CO)2(dpm)2 where dpm = methylbis(diphenyl)diphosphane and dpe = ethylbis(diphenyl)diphosphane were kindly donated and their reversible potentials measured by Dr. Graeme Snook, Department of Chemistry, Monash University.

7.4. TiO2 Plates


Many recipes may be found in literature [29, 38, 84, 139] for preparing nano-crystalline TiO 2 layers. Plates used in this study were prepared by a screen printing process, which resulted in highly reproducible surface morphologies. Any redox shuttle needs to penetrate the TiO2 layer in order to react with the sensitizer, which is adsorbed within the TiO2 layer. Hence, knowledge of the structure of this layer is important for the understanding of the transport mechanism of the redox shuttle. The TiO2 plates were characterized by in-field lens emission scanning microscopy (IFESEM). A few typical images are presented in Figure 7.2.

Chapter 7: The Efficiency of Redox Shuttles TiO2 Plates

page: 194

Figure 7.2a is a side view of the glass substrate (black, bottom) showing the coating of a thin film of fluorine doped SnO2 (white) of only a few hundred nanometers thickness. Above this layer is 8m thick layer of TiO2. The sponge like structure of the TiO2 can be seen in the top views of Figures 7.2b and c. The TiO2, consisting predominantly of anatase, forms small particles of 2025nm diameter. All these particles are joined together, forming a rigid network of several hundred particles from top to bottom (back contact). Any redox shuttle needs to travel through the pores for the redox shuttle process to occur in the required manner. It has been reported [227, 228], that the diffusivity of iodide is only decreased by a factor of less than two in the nano-porous network compared to the free solution phase.

Figure 7.2. IFESEM images of TiO2 coated glass slides. a) side view, 30ekV, iridium coated, b) top view, 15keV, iridium coated, c) top view, 2.0keV, uncoated.

The anatase morphology of TiO2 was confirmed by powder XRD. From these measurements the TiO2 average crystallite size was calculated to be 204nm, which is in good agreement with data obtained by IFESEM. The conducting layer of fluorine doped SnO2 has the cassiterite morphology and minor deviations from the theoretical XRD spectrum of pure cassiterite were assigned to the high fluorine doping content of the SnO2. The average crystallite size of SnO2 was 304nm.

Chapter 7: The Efficiency of Redox Shuttles Flow Cell Design

page: 195

7.5. Flow Cell Design


The flow cell was designed with the aim to develop a system which provides: (1) real cell like conditions. To achieve this a thin-layer arrangement is preferable, because in current dye sensitized solar cells, the cell height the distance between the TiO2 layer and the counter electrode is about 50100m. Front illumination (the light hits the TiO2 layer from top, but needs to travel through the counter electrode and solution first) and back illumination (the light hits the TiO2 layer from bottom, so that charge carriers are generated very close to the back contact) should be possible with the flow cell. (2) controlled mass transport. The charge shuttle process is driven by diffusion for a redox shuttle in a real cell. The flow cell should provide conditions of controlled mass transport. That is, the form of mass transport is known, well defined and thus computable. (3) rapid change of solution environment. A real solar cell is assembled with a certain configuration (including the solution with redox shuttle) and conditions cannot be modified after the cell is assembled. The fact that a flow through solution system is developed in this work, will make it easy to study the dependence of the cell response over a wide range of conditions. Care is needed to ensure a minimal amount of handling and that solution carry over phenomena are minimized. (4) easy assembly. Both, the solution and the dye sensitized TiO2 working electrode need to be rapidly exchangeable. Thus, complicated and irreversible sealing procedures such as sealing by wax or polyester resins, need to be avoided. The general channel flow cell design has been briefly mentioned in chapter 2.5.4 on page 50. The mass transport is well defined and is given by the Levich Equation
 x V i L = 0.925nFD 3 c 0 w f2 e h d
2 1 3

(7.1)

where w is the electrode width [cm], xe is the electrode length (in direction of the flow) [cm], 2h is

 is the volume flow rate [cm3s1] and channel height [cm], d is the width of the channel [cm], V f
the other symbols have their usual meaning. Typical cell heights used in channel flow cells are a few hundred micrometers, so that this cell design is ideal because it fulfills the two basic theoretical requirements (points 1 and 2, above). A standard in-house built channel flow cell unit was modified to meet the final two requirements and to allow the accommodation of a TiO2 working electrode. The original channel flow unit was made of a single teflon block (see Figure 7.3) consisting of two tube adapters and one 44mm

Chapter 7: The Efficiency of Redox Shuttles Flow Cell Design

page: 196

platinum electrode, which was embedded into the channel bed. A platinum electrode served as the working electrode under standard channel flow measurements and the counter and reference electrodes were placed outside the cell, down- or upstream, respectively. A plastic plate pressed against this teflon block provided both a seal and produces defined cell heights of 200400m. For the purpose of this study, which required incorporation of a TiO2 semiconductor electrode, the plastic plate was replaced with the conducting glass slide coated with the TiO2. The slide was slightly wider than the cell block to allow the leads to be attached to the potentiostat. The glass slide was prepared to accommodate the working (TiO2) as well as the counter electrode (see Figure 7.3a). The bulk of the TiO2 was removed from the conducting glass slide to leave a 56mm diameter dish acting as the working electrode. A small strip of the electrical conducting SnO 2 layer was removed around the TiO2 electrode in order to electrically isolate it from the rest of the conducting glass slide. A small band of SnO2 was permitted to remain toward the edge of the glass slide to enable contact between the TiO2 and an external electrical circuit. One half of the remaining conducting area adjacent to the TiO2 was then used as the counter electrode. The platinum electrode in the cell block opposite the glass slide was used as a quasi reference electrode. Preliminary results showed that it was crucial to have all three electrodes in close vicinity in order to avoid problems with high resistance encountered when the reference or counter electrode are placed external to the cell in either down- or upstream locations.

Figure 7.3. Custom build channel flow solar cell. a) exploded view, b) assembled cell.

Chapter 7: The Efficiency of Redox Shuttles Flow Cell Design

page: 197

Opposite the TiO2 working electrode, a 4mm diameter quartz window was embedded into the cell block in the same plane as the channel bed. This allows front illumination of the TiO2 working electrode through the cell block. A recess was manufactured around the channel bed to allow sealing of the blockglass slide unit with a silicon O-ring. The presence of the O-ring means that only moderately pressure was required for efficient sealing and this could be achieved by using a standard laboratory clamp. The assembled cell is shown in Figure 7.3b. The flow direction was chosen so that the counter electrode was down-stream of the TiO2 working electrode in order to prevent unwanted decomposition products generated at the counter electrode reaching the working electrode. The cell was calibrated frequently by using voltammograms observed with a solution of ferrocene of known concentration in acetonitrile. Use of the Levich equation (Equation 7.1) enables the cell height to be determined. For calibration purposes the platinum electrode was used as the working electrode since the electrode kinetics for oxidation of ferrocene are known to be very fast at this surface. Plots of limiting current versus (flow rate)1/3 were linear, establishing that mass transport controlled conditions were obtained, at least with the platinum electrode side. Cell heights were determined to be in the range of 400800m, depending on the pressure provided by the clamp. Frequent calibration of the cell height was undertaken. This cell design shown in Figure 7.3 resulted in stable signals being observed from a given TiO 2 working electrode for several days on average. In one instance, reproducible voltammograms were obtained from a single TiO2 electrode for a period of more than three weeks without the need to disassemble the cell and in other cases the electrode needed to be replaced after a few hours of use. The stability of the photo response was tested with a 0.1M Bu4NI solution in CH3CN. In this case the redox mediator reaction of interest is 3 I + 2 dye+ I3 + 2 dye (7.2)

and assuming that for each photo-injected electron one dye+ molecule is formed which is subsequently reduced by the dye regeneration process shown in Equation 7.2. The photocurrent typically obtained with this solution upon irradiation with 540nm (2mWcm2) light was about 3510A under all flow rates used (vide infra). It should be noted, that a current of 35A obtained at a 0.13cm2 electrode (4mm diameter) with 2mWcm2 540nm light calculates to give an IPCE value of 32% (see Equation 1.10 on page 16). This value approaches those obtained in actual sandwich type cells (50100%) [29] and is significantly higher than the IPCE of 8% reported to be achievable employing a standard stationary three electrode setup [82]. As will be shown later in the text, the IPCE value can approach 100% when the iodide counter ion is replaced by an alternative species.

Chapter 7: The Efficiency of Redox Shuttles Flow Cell Design

page: 198

50 40

photocurrent [A]

30 20 10 0 300 400 500 600 700 800 900

wavenlength [nm]
Figure 7.4. Action spectrum of (H2-dcbpy)2Ru(NCS)2 sensitzed TiO2 channel flow cell. 0.06M KI in CH3CN. Corrected for non-linear lamp output.

A typical action spectrum of a dye ((H2-dcbpy)2Ru(NCS)2) sensitized TiO2 cell in a channel flow system is shown in Figure 7.4 and is very similar to those presented in the literature [29, 82, 86]. The maximum photocurrent is obtained at a wavelength of about 540nm. The waveform at the lower energy side is defined by the absorption characteristics of the dye (compare Figure 3.22 on page 97). The low photocurrents measured at the higher energy side of the spectrum have usually been attributed to absorption by triiodide present in solution as redox shuttle. However, in this particular experiment, no deliberately added triiodide was present and the amount of triiodide formed at the TiO2 electrode surface must have been small, since any formed triiodide would be swept away by the flowing solution. Another possibility of the low photon-to-current conversion efficiency at this wavelength might be due to the fact that (H2-dcbpy)2Ru(NCS)2 shows two absorption bands in the visible spectrum (compare Figure 5.9 on page 153) corresponding to d1* (540nm) and d2* (400nm) transitions (see chapter 3.5.5 on page 104) and that electron injection from the 2* orbital of the H2-dcbpy ligand into the TiO2 manifold is unfavorable, possibly because symmetry reasons.

7.6. Results and Discussion


7.6.1. Electrochemistry at TiO2 and SnO2 Electrodes The electrochemical properties of the nano-crystalline TiO2 and SnO2 electrodes were studied by investigating the voltammetric response obtained for a well defined model system. The oxidation of ferrocene in acetonitrile was chosen as the model system. Voltammetric oxidation of ferrocene under standard conditions is generally considered to be reversible and the standard heterogeneous charge transfer at standard electrode surfaces (gold, platinum, glassy carbon, Hg, etc.) is large (48cm s1 [229]). Furthermore, since the conductivity of acetonitrile in the presence of electrolyte is very high [230], close to ideal voltammetric responses can be expected.

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 199

Uncoated and coated TiO2 conducting (SnO2) glass slides were used as working electrodes by placing them in a standard small volume voltammetric cell. The cell had a hole (ca. 7mm diameter) at the side and the slides were pressed against this hole. Sealing was achieved by placing an O-ring between both components so that conditions prevailed for what the O-ring/hole diameter then defined the electroactive area of the working electrode (6mm diameter).
0.9
current [mA]
0.4 0.2 0.0 -0.2 -0.4 -0.6 -0.8

0.6

a)
-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5

current [mA]

0.3

potential [V] vs. Fc/Fc+

0.0
0.4

peak current [mA]

0.3

-0.3

0.2

0.1

b)
0 10 20 30

-0.6 -2.0 -1.5 -1.0 -0.5 0.0

0.0

(scan rate)

1/2

[mV s

40 1/2 -1/2

50

0.5
+

1.0

1.5

potential [V] vs Fc/Fc

Figure 7.5. Cyclic voltammetric response obtained at a 6mm diameter nano-crystalline TiO2 electrode in CH3CN (0.1M Bu4NI), =100mVs1. Responses are due to reduction of TiO2 (Ti4+ + e Ti3+) and oxidation of iodide. Vertical dashed lines indicate "usable" potential range. a) as before, except 0.1M Bu4NPF6 as electrolyte. b) Plot of reduction peak current versus square root of scan rate.

SnO2 and the TiO2 working electrodes are generally considered to be chemically and electrochemically inert, particularly TiO2, which is so highly abundant in nature. However, voltammetric experiments in acetonitrile solutions containing only the electrolyte (Bu4NPF6) showed that reduction of electrode material was possible at relatively positive potentials. When SnO2 was used as the electrode material (voltammogram not shown), reduction processes starting at about 1.0V vs Fc/Fc+ were observed which probably correspond to the reduction of SnO2 to metallic tin (Sn4+ + 4e Sn0). This hypothesis was supported by noting the presence of complex oxidation processes at +0.7V vs. Fc/Fc+ when the potential was scanned in the positive oxidative direction. These anodic processes were enhanced when the switching potential for the preceding reductive scan was made more negative and are attributable to the oxidation of the previously formed metallic tin corresponding to Sn0 Sn2+ + 2e. These suggestions are supported by Fletcher and co-workers' [231] studies on the electrochemistry of SnO2 electrodes in aqueous solution. They state: "SnO2 electrodes are not inert at negative potentials. When the electrode potential is sufficiently negative a reduction process is initiated. This process involves SnIV Sn0 by an unspecified mechanism. There is strong

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 200

evidence that complete reduction to Sn0 occurs .... .... the oxidation leads to soluble SnII hydroxy species, ...." Small analogous processes at positive potentials were also detected on freshly prepared SnO2 electrodes rinsed with concentrated HNO3. TiO2 material also could be reduced, but at potentials 0.5V more positive than was the case with SnO2. A typical cyclic voltammogram obtained in CH3CN (0.1M Bu4NPF6) is displayed in Figure 7.5a. The responses in the positive potential range are due to tin chemistry (see above) which is used as the underlying substrate, whereas the complex processes in the negative potential range can be assigned to reduction of Ti4+ to Ti3+ according to: 2TiVIO2 + 2e Ti2IIIO6 + O22 (7.3)

The electrochemical reduction of TiO2 electrode material at relatively positive potentials in aqueous solution (0.1M Na2SO4, pH 2) has been also reported at single crystal anatase [79] and nano-crystalline TiO2 [232]. Cao and co-workers [232] reported a linear dependence of the anodic peak current versus the square root of the scan rate, suggesting that the TiO2 reductionreoxidation process occurs under diffusion mass transport controlled conditions, although it is not clear how this could be the case in this solid state reaction. A plot of anodic peak current (measured from zero current) versus square root of the scan rate obtained in this study is shown in Figure 7.5b and is not linear, as expected for a complex mass transport mechanism. Figure 7.5 shows a cyclic voltammogram obtained from an acetonitrile solution containing Bu4NI as the electrolyte, which is commonly used in this study as an internal standard for testing the reproducibility of the photoresponse and stability of the dye sensitized TiO2 electrode. The potential range available for photochemical measurements under potentiostatic control is then defined by the onset of the TiO2 reduction and the iodide oxidation process as indicated by the vertical dashed lines in Figure 7.5. Great care had to be taken on all subsequent experiments to specifically avoid accidental reduction of the TiO2 layer, which could significantly reduce the photovoltaic response and thus the lifetime of the cell. An intrinsic semiconductor is non-conducting (see Chapter1.6.2 on page 9). Passage of current when a semiconductor is used as an electrode requires the flatband potential to be within the conduction band or within the valence band in the space-charge region. Thus, at an n-type semiconductor, electrode electrons are the majority carriers, which allow reduction of solution species negative of the flatband potential to occur, whereas the electrode is passivated toward oxidation. From these considerations it is predicted that for the ferrocene/ferrocenium redox process, reduction (Fc+ + e Fc) will proceed as expected (at least thermodynamically), whereas oxidation will require positive potentials to be applied. This may result in large deviations of the measured E1/2-value from the true formal potential. However, the physical properties of nano-sized

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 201

semiconductor material can be significantly different to macro-sized material (see chapter 1.7.4 on page 20) so that deviations from results predicted on the basis of these rules can be observed.
100

a)

current [A]

50 0 -50

-100

b)
100

current [A]

50 0 -50 -1.0 -0.5 0.0 0.5


+

1.0

potential [V] vs. Fc/Fc

Figure 7.6. Oxidation of 1.0mM ferrocene in CH3CN (0.2M Bu4NPF6) at a a) 6mm diameter SnO2 coated glass electrode; b) 6mm diameter nano-crystalline TiO2 electrode (on SnO2 coated glass) under condition of cyclic voltammetry, =10, 25, 200, 1000mVs1. All scans background corrected. Vertical dashed lines indicate peak potetnials

SnO2 electrode material scan rate [mVs1] 1000 500 200 100 50 26 10 Epox Epred [mV] 674 584 456 364 330 274 198 TiO2 electrode material scan rate [mVs1] 1000 500 200 100 50 26 10 Epox Epred [mV] 1400 1304 1030 878 806 754 646 khet [cm s1] 0.0053 0.0043 0.0043 0.0042 0.0035 0.0029 0.0024 khet [cm s1] 0.022 0.020 0.021 0.024 0.021 0.022 0.026

Table 7.1. Heterogeneous rate constant for the oxidation of 1.0mM ferrocene in CH3CN (0.2M Bu4NPF6) obtained from cyclic voltammograms.

The oxidation of 1mM ferrocene in CH3CN (0.2M Bu4NPF6) was investigated at a 6mm diameter SnO2 electrode. A series of cyclic voltammograms obtained over the scan rate range of 101000mVs1 is shown in Figure 7.6a. The peak-to-peak separation (Ep) is clearly larger than

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 202

the 57mVexpected for an electrochemically and chemically reversible one-electron transfer process. The resistivity of acetonitrile is reasonable small [230] so that the contribution to Ep from uncompensated resistance should not be significant. Large peak-to-peak separations are expected when the heterogeneous electron transfer is slow. Bard and Faulkner have presented a working curve (Table 6.5.2 in reference [113]) that allows calculation of the heterogeneous rate constant from peak-to-peak separations. The data are summarized in Table 7.1 and allow a heterogeneous rate constant for the ferrocene oxidation process of khet=0.0220.002cms1 to be estimated, which is more than two orders of magnitude lower than the value measured on "normal" electrode material (see above). Similar experiments to those described at SnO2 were undertaken on TiO2 electrode material and the resulting cyclic voltammograms for the oxidation of ferrocene are shown in Figure 7.6b. The peak-to-peak separation on this electrode material is even larger than on SnO2. In case of the TiO2 electrode a reaction with the underlying SnO2 is theoretically possible. However for this to occur, ferrocene needs to diffuse through the whole 10m TiO2 layer without prior reaction with TiO2. Since voltammetric responses are obviously different at the TiO2 electrode, a contribution from the SnO2 layer must be small. The cyclic voltammetric data obtained for the oxidation of ferrocene at TiO2 are summarized in Table 7.1 and analysis gives an average heterogeneous rate constant of khet=3.51103cms1. 7.6.2. Calibration and Mass Transport in the Channel Flow Cell A typical photoresponse obtained with the channel flow cell arrangement and Bu4NI as the electrolyte is displayed in Figure 7.7.
9 8 7 6

light off

current [A]

5 4 3 2 1 0 0 5 10 15 20

light on

time [s]
Figure 7.7. Typical photoresponse obtained from the dye sensitized TiO2 electrode when using a channel flow cell. CH3CN (20mM Bu4NI), =540nm, flow rate: 0.01cm3s1.

For this and other related experiments, the initial potential applied to the TiO2 working electrode was chosen to be the potential of zero current under dark conditions. As pointed out above, the selected potential range was restricted to the range 0.5V0.7V in the presence of iodide. Additionally, the use of a platinum quasi reference electrode required that the relevant applied

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 203

potential had to be selected each time the solution environment was changed. Finally, the potential was fine tuned to obtain the maximum photoresponse. In order to probe the nature of the mass transport of the electroactive compound (iodide) in the channel flow cell, the magnitude of photocurrent was investigated as a function of added iodide concentration. As expected, a linear relationship between photocurrent and iodide concentration was found at low concentrations (100mM, see Figure 7.8 insert), but linearity was lost at higher concentrations, probably due to an ion pairing or other form of saturation mechanisms. It should be noted that the maximum current measured was smaller than that theoretically predicted when 100% photon-to-light conversion efficiency prevails (100%IPCE 0.1mA).

110 100

calculated measured

photocurrent [A]

40 30
15

20 10 0 0 50 100

10 5 0 0 20 40 60

150

200

250

300

iodide concentration [mM]


Figure 7.8. Calibration of a dye sensitzed TiO2 electrode channel flow cell with CH3CN/Bu4NI under constant flow rate of 0.23cm3s1 and irradiation of 2mWcm2 (=540nm). Measured slope:0.26AmM1, calculated slope: 16AmM1.

Use of the Levich Equation (Eqn. 7.1) enables the theoretically expected mass transport controlled limiting current to be calculated. Assuming a square electrode with the area 2r (r=2mm) and use of the other parameters given in Figure 7.8, enables the limiting current to be calculated as a function of iodide concentration and the data are displayed in Figure 7.8 (blue line). The calculated slope of 16AmM1 is almost two orders of magnitude larger than the measured slope of 0.26AmM1. In fact, the limiting current (photocurrent) was found to be independent of flow rate. Thus, even when the flow was stopped, the photocurrent remained unchanged for a considerable period of time ( 5s) before a decay was noticeable. Such a behavior would be expected in a CE type reaction where a rate determining chemical reaction takes place before the electrochemical step. If this chemical reaction is sufficiently slow, a flow rate independent response is expected over a large flow rate range. An example for such a mechanism is shown in Equations 7.4 and 7.5 and where the identity of X is unknown. I X (7.4)

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 204

X + dye+

X+ + dye

(7.5)

The theoretically expected concentrationphotocurrent characteristic should be defined by the slope obtained from the Levich equation to the left (blue line in Figure 7.8) assuming that channel flow mass transport conditions are achieved and the maximum number of electrons available (defined by the light intensity, 100% IPCE-value, ca. 0.1mA in case of Figure 7.8, magenta line) should define the top plateau. The discrepancy between theory based on mass transport control and experiment suggests that the reaction of the redox mediator iodide and the oxidized dye is limited by processes which are not restricted to transport, such as kinetic limitations. 7.6.3. Phototransients and Double Layer Capacitance When the light source is switched on in a stationary solution experiment, the electrode potential should change from zero to a more positive value which is defined by the reversible potential for the sensitizer. Thus, in principle, a light-offlight-on cycle, corresponds to the potential step experiment as described in chapter 2.6.1 on page 50. The expected current decay for this kind of experiment should be governed by the Cottrell equation (Eqn. 2.31), if the process is diffusion controlled. In case of the channel flow cell it would have been expected that the current would reach a maximum value immediately after the light is turned on and then decay, according to the Cottrell equation linearly with the square root of time (see Figure 2.12 on page 51) until it reaches the limiting or steady-state value defined by the Levich equation.

photocurrent [A]

30 20 10 0 1.0

a)

photocurrent [A]

b)

0.5

0.0 10 20

time [s]
Figure 7.9. Phototransients obtained at different iodide concentrations (=540nm (2mWcm2), CH3CN, flow rate: 0.01cm3s1). a) 100mM Bu4NI, b) 3mM Bu4NI.

In contrast with this prediction (see Figure 7.9), after the light was switched on, the current started to increase slowly until it reached a steady-state value. Furthermore, the rise time for the photocurrent was found to depend on the concentration of iodide present in solution; at low iodide concentrations the rise time was much larger than at higher iodide concentrations.

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 205

The observed behavior would be explained if the TiO2 working electrode behaves as a capacitor that needs to be charged before faradaic or other electrochemical processes can be observed. Due to its nano-porous character, the actual surface area of the TiO2 working electrode is several hundred times larger than its geometric area [49]. When an electrode is immersed into a medium that contains charged species, e.g. the electrolyte, and a potential is applied, these charges will move to compensate for the electric field, and thus, the electrode(s) will act as a capacitor. The simplest model describing these effects is the Helmholtz model, which proposes the formation of the so called Helmholtz layer, where it is assumed that there are two sheets of charge, having opposite polarity and derived from the ions in solution, separated by a distance of molecular order. Under these circumstances, the double layer capacitance could be calculated by assuming it is derived from an electronic capacitor (see Equation 2.3). However, the Gouy-Chapman theory provides a more realistic treatment of the double layer problem and involves a diffuse layer of charge in the solution where the greatest concentration of charge would be adjacent to the electrode and progressively lesser concentrations would be found at greater distances as electrostatic forces are weakened. Consequently, the average distance of species from the electrode is a function of potential and electrolyte concentration. This theory predicts a diffuse layer thickness of only a few angstroms for high electrolyte concentrations (3 thickness at c=1M) and a much larger layer thickness at lower concentrations (304 thickness at c=0.1mM). The double layer capacitance is proportional to the square root of the concentration and given by the following equation [233]
C d = 228zc 02 cosh( 19.5 0 )
1

(7.2)

where c0 is the concentration of the electrolyte, z is the charge of the electrolyte and 0 is the potential of the electrode relative to the bulk of the solution. Stern's modification of the GouyChapman theory corrects some shortcomings of this approach, but maintains the proportionality of the double layer capacitance with the square root of the electrolyte concentration.

photocurrent x rise time [C]

2.0

1.5

1.0

0.5 0 2 4 6 8 10
1/2

12

14
1/2

16

(electrolyte concentration)

[mM ]

Figure 7.10. Double layer capacitance of TiO2 electrode versus square root of electrolyte concentration. CH3CN, 3mM Bu4NI, flow rate: 0.01cm3s1, Bu4NPF6 added, TiO2 sensitized with (H2-dcbpy)2Ru(NCS)2, =540nm (2mWcm2).

To ascertain if the phototransients are consistent with a large surface area and a large double layer capacitance the following experiment was undertaken. Phototransients of dye sensitized TiO2

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 206

electrodes were measured in the presence of a small amount of redox mediator, 3mM Bu 4NI, and as a function of electrolyte (Bu4NPF6, c=0200mM) concentration. The product of photocurrent and rise time14 has the unit of Coulomb and is directly proportional to the capacitance since the electrode potential is maintained constant throughout the experiment. A typical plot of product of photocurrent and rise time versus square root of total electrolyte concentration is shown in Figure 7.10. Clearly, the linear relationship expected, if phototransients had obeyed the diffuse double layer theory, is not found. "Slow" photoresponses have been observed at dye sensitized and unsensitized nano-porous TiO2 in other studies and a variety of explanations are available in the literature [39, 40, 45, 47]. Schwarzburg and Willig [47] have shown that charge trapping in semiconductor defect states can give rise to very slow phototransient processes, as found in polycrystalline TiO 2. This material is characterized by a high density of trap sites, and consequently, charge transport is often dominated by the properties of the traps. Cao and co-workers [40] were able to mathematically model the photoresponse at dye sensitized TiO2 electrodes which resulted in a light intensity dependent diffusion coefficient of TiO2 conduction band electrons. The authors also report rise times of several seconds and further state that the electron diffusion coefficient is expected to be a function of the sample morphology and preparation methods. Another explanation of the effect may be the charging of the space-charge layer that develops at the semiconductorelectrolyte interface. 7.6.4. The Effect of Cation Size The iodide redox shuttle has been used in dye sensitized solar cells as different iodide salts. Most commonly, lithium (Li+) [29, 49, 78, 138] has been used as the cation, but K+ [28, 81, 225, 234], Na+ [67, 73, 75, 75], (i-C3H7)4N+ [69, 139, 235], (n-C4H9)N+ [84] and recently the 1-hexyl-3methylimidazolium cation [227] have also been used. No comments on the relevance of choice of salt appear to be available in the literature. Merely O'Regan and Grtzel [72] have commented briefly that slightly higher IPCE values can be achieved when LiI is used instead of (C3H7)4NI, but no explanation on this unusual effect has been provided. The effect of the iodide counter cation on the photocurrent was investigated by using the following iodide salts as redox mediators: LiI, KI, NH4I, (CH3)4NI, (C2H5)4NI and (C4H9)4NI. In order to make it possible to obtain the same solubility for all of these salts, the experiments were carried out in a 95:5 CH3CN/H2O mixture. The results of these experiments are shown in Figure 7.11 as a plot of photocurrent versus cation radius. As observed with experiments conducted in CH 2Cl2, the photocurrent achieved in the CH3CN/H2O mixture with the "standard" iodide salt Bu4NI was only 3010% of the value measured in pure CH3CN. Interestingly, a linear dependence of photocurrent versus cation radius was found. Photocurrents measured with the small lithium cation being present

14

The rise time was measured as the time that was needed for the current to reach 95% of the steady-state value.

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 207

in solution resulted in a five fold increase of the photocurrent compared to when the relatively large tetrabutylammonium cation was used as the counter ion (see Table 7.3).

Figure 7.11. Photocurrent as a function of iodide cation size. Iodide concentration: 50mM, solvent: CH3CN + 5% H2O, =540nm, cation size from reference [236] and [237].

A second interesting result was obtained when solutions containing mixtures of small and large cations were investigated. In one experiment, the small ammonium cation (r=1.5) was successively added as NH4ClO4 to a CH3CN/0.1M Bu4NI solution containing the large tetrabutylammonium cation (r=5.0) and the result is shown in Figure 7.12. The addition of very small concentrations of ammonium cations results in a large increase in photocurrent. For instance, the addition of only 0.05mol% NH4+ cations to the 0.1MBu4NI solution doubles the photocurrent! It is noted, that the use of Li+ ions in pure CH3CN leads to IPEC values of about 100%!
120 100

photocurrent [A]

80
photocurrent [A]

100

60 40 20 0

80

60

40

20 0 1 2 3 4

conc. added NH4NCS [mM]

10

15
+ 4

20

25

30

conc. added NH [mM]


Figure 7.12. Photocurrent as a function of added NH4+ cations. 0.1M Bu4NI, =540nm, solvent: CH3CN, ammonium cations added as NH4ClO4.

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 208

The influence of the redox shuttle cations on the photocurrent may be rationalized by two hypotheses. Firstly, results in Figure 7.11 suggest that the photocurrent is dominated by the geometric size of the cation. The TiO2 particles have an average diameter of about 22nm, whereas the tetrabutylammonium cation has a diameter of 1nm. Furthermore, Papageorgiou, Barb and Grtzel [227] report a 30% decrease in TiO2 porosity after dye adsorption onto the TiO2 surface. Thus, free channels in the TiO2 network, which are available for mass transport of cations, have atomic dimensions. The height of the TiO2 layer is commonly a few m (410m), which is several orders of magnitudes larger than any cation diameter. Dye sensitization occurs across this TiO2 layer, hence the iodide anion has to diffuse into the TiO2 layer to the sensitization site. In order to maintain charge neutrality, the counter ion must co-diffuse with the iodide anion into the TiO2. Charge neutrality cannot be neglected over this distance, if the counter ion could not diffuse together with the iodide anion into the TiO2 layer an unfavorably high electric field would be needed to allow this process to take place. The hindrance of free diffusion of cations into the TiO2 layer would possibly explain the origin of the decreased photocurrents measured with the larger ions and the results shown in Figure 7.11. The second possible explanation may be related to the special properties of nano-sized TiO2 semiconductor particles. For example, the absence of a well defined space charge region characterizes nano-sized semiconductor particles (Chapter 1.7.4 on page 20). The existence of a space charge region requires a large region of "bulk" semiconductor, with bulk semiconductor properties, with respect to which changes near the surface can be observed. These bulk semiconductor regions are absent in nano-sized particles and thus, any physical changes in the semiconductor properties occur across the whole particle and not only at the surface. Therefore, the semiconductor bands of nano-sized particles "float" and are more sensitive to the solution/electrolyte environment than bulk semiconductor material. Several presentations at the 12th International Conference of Solar Energy Conversion and Storage in Berlin 1998 [238] addressed the dependence of cation and led to the suggestion that the effect of the redox shuttle cation depends on the cation charge density rather than on geometrical size. Thus, when the Helmholtz layer is formed, ions are in direct contact with the semiconductor surface and thereby influence and alter the semiconductor properties. Enright and co-workers [219] have reported that cations are flatband potential determining in nano-crystalline TiO 2. Zaban, Ferrere and Gregg [214] have correlated results obtained at dye sensitized semiconductor electrodes in the presence of lithium or tetrabutylammonium cations with the thickness of the Helmholtz layer giving rise to a cation dependent potential gradient at the semiconductor surface.

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 209

7.6.5. Efficiency of Redox Shuttles (a) Thermodynamic Requirements and other desirable Properties of Redox Shuttles Redox Potential The theoretical thermodynamic properties required for a redox shuttle are clearly defined. In order to be able to reduce the oxidized dye back into its initial oxidation state, the redox shuttle reversible potential has to be negative of the dye reversible potential. The more negative the potential for the redox mediator, the larger the thermodynamic driving force for the dye regeneration. However, as illustrated in Figure 7.1, the more negative the potential or the higher the energy for this process, the smaller the open circuit potential and hence the efficiency of the solar cell. Thus, the reversible potential for a redox shuttle must be negative of the reversible potential for the dye but as positive as possible to avoid unnecessary loss of usable energy.

High So lubility
Concentration ranges commonly used for redox shuttles are 0.1M0.5M to ensure sufficient supply of the redox mediator and to minimize the possibility of diffusion limited situations arising, which would result in an undesirable extended lifetime of the oxidized form of the dye and consequently would increase the chance of dye decomposition.

High D iffusion Coefficient


Mass transport of the redox shuttle occurs solely by diffusion, therefore a high diffusion coefficient is essential for enhanced mass transport.

No spec tral Characteristics in the Visible Region


Absorbance of visible light by the redox shuttle system results in less light being available for the light-to-electricity conversion and thus low energy conversion. Since the redox shuttle is present in a relatively large concentration, even small extinction coefficients will affect the solution absorbance characteristics.

High St ability of both Reduced and Oxidized Form of the Couple


For efficient redox "shuttling" in solar cells, both forms of the redox shuttle (oxidized and reduced) must be present in solution and both forms must have high stability. In the case of the iodide/triiodide system, the reduced (iodide) form is in excess [81, 234]. However, extremely high stability for both forms is required.

Highly Reversible Couple


The oxidation of the reduced form and reduction of the oxidized form of the redox shuttle must be electrochemically and chemically reversible to guarantee fast electron transfer and to avoid side reactions, which may lead to decomposition. (b) Results Iodide The theoretical thermodynamic efficiency of redox shuttles is illustrated in Figure 7.13 where the voltammetry of the dye (H2-dcbpy)2Ru(NCS)2 is compared to the voltammetry of several

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 210

potential sensitizers, such as iodide, bromide and ferrocene. The oxidation potential for the sensitizer (red dashed vertical line) determines the thermodynamic threshold for the oxidation potential of the redox shuttle. The cyclic voltammograms presented in Figure 7.13, provide approximate redox potentials, since complete reversibility is not found under conditions of cyclic voltammetry in acetonitrile. Additionally, the reversible potential for solid (H2-dcbpy)2Ru(NCS)2 attached onto the TiO2 surface may be different to the solution phase reversible potential, although the potential difference is expected to be small, since solid (H2-dcbpy)2Ru(NCS)2 attached to carbon electrodes exhibits a reversible half-wave potential which is identical to that found for the dissolved compound (see chapter 4.3.7 on page 129). The oxidation of halides occurs in two steps [239] according to the following equations (also see page 76): 6X 2X3 + 4e 2X3 3X2 + 2e (7.6) (7.7)

Firstly, the halide anion (X) is oxidized to the trihalide anion (X3) and in a second step the trihalide anion is oxidized to the halide (X2). Thus, two processes are observed for the voltammetric oxidation of iodide and bromide corresponding to the above processes. Deviations from the expected reversible voltammetric behavior have been explained in terms of high surface activity of the products [170]. The voltammetric responses for the oxidation of iodide and bromide are almost identical in shape but have a potential difference of 0.38V. The potential difference between the X/X3 and X3/X2 processes is approximately 0.35V for both halides.
thermodynamically available potential range for redox shuttles

Dye/Dye

(H2-dcbpy)2Ru(NCS)2

Br /Br3

Br3 /Br2

Bromide

I /I3

I3 /I2

Iodide

Fc/Fc

Ferrocene

-0.5

0.0

0.5
+

1.0

1.5

potential vs. Fc /Fc [V]


Figure 7.13. Cyclic voltammograms of the sensitizer and some potential redox shuttles in CH3CN (0.1M Bu4NPF6). =100mVs1, c=1.0mM except (H2-dcbpy)2Ru(NCS)2 c=3.5105M.

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 211

The commonly used iodide/triiodide redox shuttle fails on many of the above criteria. Triiodide absorbs strongly below 500nm (400nm=8103M1cm1) even in ultra thin cell designs and its redox potential is more than 0.5V more negative than the sensitizer. Open circuit voltages achieved with the (H2-dcbpy)2Ru(NCS)2 sensitizer are around 500mV [29, 240]. Thus, the "unnecessary" potential difference of 0.5V between the dye and the redox shuttle decreases the efficiency of the cell by 50%, which means that the efficiency of this kind of dye sensitized solar cell could be doubled by a better match of the sensitizer and redox shuttle potentials!

Triiodid e
Another interesting feature shown in Figure 7.13 is the thermodynamic ability of triiodide to reduce the oxidized form of dye according to Equation 7.8. 2 I3 + 2 dye+ 3 I2 + 2 dye (7.8)

In fact, triiodide should be a significantly preferred reductant since its oxidation potential is only separated by ca. 0.2V from the reversible potential of (H2-dcbpy)2Ru(NCS)2. In dye sensitized solar cell applications, the oxidized and reduced form of the redox mediator are present; in case of iodide, triiodide is also present (see above). It has not been clarified yet which process (Eqn 7.2 or 7.8) is the electroactive process reducing the oxidized dye, since both compounds posses a sufficiently large thermodynamic driving force to act as the reductant.
percentage [%]
100 25 20 15 10 5 0 80 60 40 20 0

photocurrent [A]

20

40

60

80

100

percentage [%]
Figure 7.14. Photocurrent of (H2-dcbpy)2Ru(NCS)2 sensitized TiO2 as a function of triiodide concentration in CH3CN. 100% I = 100mM Bu4NI, =650nm, flow rate: 0.01cm3s1.

In an experiment iodine (I2) was successively added to a solution containing CH3CN/0.1M Bu4NI, thereby consecutively converting the iodide to triiodide.15 The photocurrent recorded in the channel electrode as a function of iodide/triiodide concentration at a flow rate of 0.01cm3s1 is shown in Figure 7.14. To avoid light absorbance by triiodide, the experiment was conducted at a

15

It can be assumed that quantitative conversion into the trihalide occurs because the formation constant for triiodide in acetonitrile is Kf=106.6 [170].

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 212

wavelength of 650nm where triiodide under the most unfavorable conditions (cell height 500m, 0.1M I3) exhibits an absorbance of 0.045 (650nm10M1cm1). With increasing triiodide concentration, the photocurrent decreased with no photocurrent being detectable when 100% I3 (c=100mM) was present. This result is as expected if triiodide cannot effectively reduce the oxidized dye and demonstrates that it is indeed the iodide anion which reduces the oxidized sensitizer. Related results have been obtained by Lindstrm and co-workers [82] who showed that experiments under potentiostatic control resulted in a decrease of IPCE values over the whole visible spectrum when the triiodide concentration was increased [82]. The adverse effects of triiodide on the cell operation indicate that the concentration of triiodide should be reduced to the minimum amount necessary for operating solar cell devices.

Bromid e, Thiocyanate and Ferrocene


The 0.38V potential difference between the redox processes of iodide and bromide, should make bromide the preferred redox shuttle. However, when Bu4NBr was used instead of Bu4NI, the detected photoresponses were very small and less than 2% of those observed when the same amount of iodide was used.
p +)

compound Bromide (Br) Triiodide (I3) Thiocyanate (NCS) Ferrocene (Fc) LiI (I) trans-[Cr(CO)5I](C4H9)4N cis-Mo(CO)2(dpe)(dpm) trans-Cr(CO)2(dpe)2 cis-Cr(CO)2(dpm)2

Er1/2 or E ox [V] vs. Fc+/Fc +0.32 *) +0.27 *) +0.2 *) 0.000 0.06 *) 0.06 0.32 0.53 0.59

photocurrent (compared to Bu4NI) [%] <2% none <2% none ~550% #) <1% yes #) yes #) yes #)

Table 7.3. Potentials of redox mediator systems and their performance under potentiostatic control. +) Potentials measured in CH3CN (0.1M Bu4NPF6). *) Irreversible process.#) See text.

Thiocyanate is a pseudo halide which, in an analogous manner to iodide and bromide, also can form trithiocyanate ((NCS)3) and thiocyanagen ((NCS)2) when oxidized [172, 241]. Its oxidation potential lies between those of iodide and bromide (see page 126) and therefore, based on the redox potentials, thiocyanate is preferable to iodide. However, the stability of thiocyanate oxidation products is relatively low and similarly to bromide only very low photocurrents (<2% of Bu4NI) were measured when Bu4N(NCS) was used as the electrolyte (see Table 7.3). Ferrocene has successfully been used in solid state cells as a redox mediator [226]. The ferrocene/ferrocenium couple is highly reversible and its oxidation potential is similar to that of iodide which suggests that it may show similar performance. However, when ferrocene in CH3CN (0.1M

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 213

Bu4NPF6)16 was used no photocurrents could be measured even at very high (100mM) ferrocene concentrations.

Chromi um and Molybdenum Complexes


[Cr(CO)5I](C4H9)4N, cis-Mo(CO)2(dpe)(dpm), trans-Cr(CO)2(dpe)2 and cis-Cr(CO)2(dpm)2 were tested for performance as redox shuttles in the channel flow solar cell system. Due to their relatively negative potentials (see Table 7.3), low solubilities in CH3CN (0.17mM) and toxicity, these compounds would be unsuitable for real solar cell applications but they provide additional information on the dye regeneration process. Experiments on these complexes were conducted in CH2Cl2 (0.1M Bu4NPF6) where all compounds were at least 2mM soluble. Photoresponses were generally less than observed in CH3CN. For Bu4NI the photocurrent in CH2Cl2 was only 2030% of that measured at a similar concentration level in CH3CN. Photoresponses obtained for the four molybdenum and chromium complexes at a 2mM concentration level are presented in Figure 7.15. For comparison, when 2mM Bu4NI was used as the redox shuttle an average photocurrent of 0.60.2A was measured. The shape and intensity of the photoresponses shown in Figure 7.15 are clearly different to those observed with Bu4NI (see Figure 7.7).
25 20

a)

trans-Cr(CO)2dpe2
photocurrent [A]

b)
1.5

cis-Mo(CO)2(dpe)(dpm)
light off

photocurrent [A]

15 10 5 0 0 10

light off light on

1.0 light on 0.5

20

time [s]
8

10

20

30

40

time [s]

c)

cis-Cr(CO)2dpm2
light off 0.25

d) [Cr(CO)5I](Bu4N)
light off

photocurrent [A]

6 4 light on 2 0 0 10

photocurrent [A]

0.00 light on -0.25

time [s]

20

30

40

10

20

30

40

time [s]

Figure 7.15. Photoresponses obtained for various Mo and Cr complexes at a (H2-dcbpy)2Ru(NCS)2 sensitized TiO2 electrode in CH2Cl2 (0.1M Bu4NPF6). c=2mM, =540nm, a)+b) flow rate: 0.024cm3s1 c) flow rate: 0.033cm3s1 d) flow rate: 0.036cm3s1.

The photoresponses obtained with trans-Cr(CO)2(dpe)2 are most interesting (Figure 7.15a). In contrast to measurements conducted with iodide, the photoresponses with trans-Cr(CO)2(dpe)2 are closer to those theoretically predicted for a mass transport controlled channel flow system as
16

Photocurrents were uneffected when Bu4NPF6 was added to CH3CN/Bu4NI solutions

Chapter 7: The Efficiency of Redox Shuttles Results and Discussion

page: 214

discussed in Section 7.6.2 on page 202. In case of iodide, the photocurrents were always significantly smaller than predicted by the mass transport calculations on the flow cell. With transCr(CO)2(dpe)2 as the redox shuttle, the photocurrents were 2030 times larger than when a similar concentration of Bu4NI was used! Furthermore, Figure 7.15a shows a phototransient which decays in Cottrell like manner, as expected for initially diffusion controlled conditions. Indeed, the initial decay of the photocurrent was linearly dependent on the square root of time for the first few seconds before the steady-state condition was reached. Additionally, the photocurrent showed a strong dependence on flow rate. This feature was not observed with any other redox shuttles used. However, a Levich plot of current versus (flow rate)1/3 was not linear, suggesting that completely mass transport controlled conditions have not been fully achieved. Photocurrents and transients acquired with cis-Cr(CO)2(dpm)2 (Figure 7.15c) and cisMo(CO)2(dpe)(dpm) (Figure 7.15b) as redox shuttles are similar to those obtained with iodide. That is, photocurrents are independent of flow rate and phototransients did not show a Cottrell like behavior at short times. However, photocurrents were higher (by a factor of ~10 and ~2, respectively) than measured with Bu4NI. The case of [Cr(CO)5I] demonstrates that sometimes results can be very complicated and cannot always be explained in a straight forward manner. Unusual phototransients with "opposite polarity" were measured with no or little overall photocurrent (see Figure 7.15d). That is, when the light was switched on, a negative photocurrent was detected (electrons being withdrawn from the TiO2 electrode) and consequently when the light was switched off a positive photocurrent was obtained. All four complexes are highly colored so that a reaction with the excited state complex molecule could occur, although only little dependence on the shape and magnitude of the photoresponse was found when the wavelength of the light source was varied over the visible range.

7.7. Conclusions
A flow dye sensitized solar cell test system has been successfully constructed and tested and provides a significant improvement over present available procedures for measurement of photocurrents under a wide range of different solvent/electrolyte conditions. Dye sensitized TiO2 electrodes used with this system were stable for several weeks which enabled 100 different solvent/electrolyte solutions to be studied without the need for disassembling the flow cell. The most reproducible and stable conditions are achieved in acetonitrile solvent. Measurements made in other solvents or mixed solvents (i.e. CH2Cl2, DMF, H2O and mixtures with CH3CN) resulted in reduced lifetimes of the sensitized TiO2 working electrode. Despite the fact that no redox shuttle has been found that could replace iodide, this should be a reasonable goal since the iodide/triiodide redox shuttle system possesses many undesirable properties which lead to a loss of the theoretical overall cell efficiency of up to 50%. Alternative redox shuttles tested, such as bromide and ferrocene, were over two orders of magnitude less

Chapter 7: The Efficiency of Redox Shuttles Conclusions

page: 215

efficient than iodide. Furthermore, this study has shown, when the iodide/triiodide redox couple is used for the redox shuttling process, only the iodide anion is capable of regenerating the oxidized dye. The performance of redox shuttles clearly is not dominated by their thermodynamically properties. Thus, kinetic properties of the redox shuttle system would appear to determine their efficiency. An important side reaction lowering the efficiency of redox shuttles in real cells is the back reaction between the photogenerated TiO2 conduction band electron and the oxidized form of the redox shuttle (eTiO2 + I3), in which case the triiodide acts as an electron scavenger. This assumption is supported by results obtained from measurements of the triiodide anion shown in Figure 7.14. If the triiodide anion is only electroinactive toward the dye generation and does not act as an TiO2 electron scavenger, a linear slope of a plot of triiodide concentration versus photocurrent would have been expected, mirroring the results found for calibration of the flow system shown in Figure 7.8. However, the large deviation from linearity suggests, that I 3 is not only electroinactive toward the dye generation but also scavenges conduction band electrons. Interestingly, some molybdenum and chromium carbonyl complexes have been found to significantly outperform iodide in respect of efficiency under certain conditions. Unfortunately, due to their unfavorable physical and chemical properties (solubility, oxidation potential, absorbance, etc.), these complexes would not be useful for use in real cells, but may provide the basis for discovery of alternatives to iodide. Specifically trans-Cr(CO)2(dpe)2 appears to have considerably less kinetic limitations than iodide; under low concentration conditions this complex produced photocurrents to be measured that are 2030 times higher than encountered with Bu4NI under identical conditions.

Chapter 8

Chapter 8: Conclusions The Thermodynamic Importance of the Oxidation Potential

page: 217

8.1. The Thermodynamic Importance of the Oxidation Potential


In this thesis, the electrochemical properties of sensitizers commonly employed in dye sensitized solar cells have been studied in great detail. In particular, the oxidation potentials for the sensitizer cis-(H2-dcbpy)2Ru(NCS)2 and [(H3-tctpy)Ru(NCS)3], and their deprotonated forms have been meticulously measured in the solution phase. Surprisingly, the observed trend of lower reversible potential having improved cell performance contradicts what is commonly believed to represent a simplified, but qualitatively correct, schematic representation of energy transitions (see Figure 8.1). The results of these studies thus question the use of such reaction schemes for this purpose. On the other hand, the observed trend that low potentials lead to improved cell performance suggests that the thermodynamic value for the oxidation of the sensitizer is important and may be related to cell performance, although the mechanistic aspects of this relationship are still unclear. This trend has interesting consequences and for example, would imply that S- bonded thiocyanate linkage isomers of sensitizers may give superior performance to their N- bonded counter parts, since it was shown (Chapter 6.3.4 on page 176) that a change of the thiocyanate coordination from the N- to S- bonded mode can shift the reversible potential less positive by as much as 60mV.

A
e

S+/S*
fermi level

e n e rg y

Eoc R/R

TiO2

electrolyte solution

Figure 8.1. Simplified diagram showing the electron transitions that occur during the operation of a photoelectrochemical solar cell. S is the sensitizer, R is the electrolyte (commonly I/I3) and Eoc is the open circuit potential.

8.2. Surface Based Processes During the Voltammetric Oxidation


Perhaps not surprisingly, the voltammetric oxidation of the sensitizers can be accompanied by surface based processes. That is, the starting material adsorbs onto electrode surfaces prior to commencing a voltammetric scan as do products formed after oxidation of the sensitizer. These surface interactions may lead to deviation of the responses from those theoretically predicted for pure mass transport control and great care has to be taken to measure the true thermodynamically important value for the oxidation process, ideally under conditions where the effects of adsorption

counter electrode

S+/S

Chapter 8: Conclusions Electrochemical Reduction and the Effect of Deprotonation

page: 218/247

are minimal. From studies carried out on an extensive series of sensitizers and their derivatives, it has emerged that the thiocyanate ligands present in the sensitizers are the origin of substantially increased surface activity relative to compounds where the thiocyanate ligands are replaced by other halides or pseudo halides such as chloride, iodide or cyanide (see Chapter 3 and reference [195]) where minimal adsorption is observed. In principle, part of the surface activity may be also attributable to the polypyridyl carboxylic acid ligands (H2-dcbpy and H3-tctpy) which are also known to adsorb onto certain electrode surfaces [196]. However, surface interaction of this ligand via its acid functionalities, which is the dominant mode of interaction of these sensitizers with a TiO2 surface [63, 67-69], must be small at conventional electrodes since esterification and thereby blocking the acid proton does not too significantly alter the voltammetric appearance of surface based processes (see chapter 3.4).

8.3. Electrochemical Reduction and the Effect of Deprotonation


Avoidance of an acid-base equilibrium by esterification of the acid groups, leads to the observation of rich cathodic electrochemistry. Four one-electron reduction processes are accessible for the (Et2-dcbpy)2RuX2 (Et2-dcbpy = ethyl ester of H2-dcbpy) series of complexes. Detailed spectroscopic studies reveal electron "hopping" between ligands which confirms a ligand localized electron model as found for related complexes in work by other scholars. In contrast, when the acid functionalities of the sensitizers are in their protonated form, the voltammetric reductive scans are highly complex. Electrochemical reduction of the acid form of the complexes under "standard" and long time-scale voltammetric conditions does not produce ligand based radicals, as expected from studies on the ester analogues and similar compounds, but instead reduces protons provided by acid-base equilibria in which the acid complexes participate. This proton reduction leads to hydrogen gas formation and simultaneously to deprotonation of the ruthenium complexes. Interestingly, not all protons can be removed by electrochemical reduction or by addition of a strong base, such as Bu4N(OH). In case of (H2-dcbpy)2Ru(NCS)2, voltammetric and spectroscopic measurements, as well as molecular orbital calculations, indicate formation of the doubly deprotonated sensitizer [(H-dcbpy)2Ru(NCS)2]2 (H-dcbpy is the singly deprotonated H2-dcbpy ligand) as being the major product formed upon electrochemical reduction of the initial protonated complex. When [(H3-tctpy)Ru(NCS)3] is electrochemically reduced, the singly deprotonated complex, [(H2-tctpy)Ru(NCS)3]2, is probably formed as the major product, but some doubly deprotonated complex, [(H-tctpy2)Ru(NCS)3]3, also may have been formed. The electrochemical deprotonation of dye sensitizers is of particular interest, since deprotonated forms of sensitizers are currently employed [203] in solar cell devices. Importantly, the kind of electrochemical experiment (coulometric titration) allows stepwise deprotonation, with concurrent spectroscopic monitoring of the complex. Most interestingly, the absorption bands decay in

Chapter 8: Conclusions Redox Shuttles

page: 219

intensity and experience a blue shift, and the oxidation potentials for the metal centered oxidation of the complexes shift by 0.30 0.36V to less positive potentials during exhaustive deprotonation. Both effects are generally considered to be detrimental to cell performance. Once again the role of the oxidation potential of the sensitizers in regard to cell performance remains unclear and the validity of the commonly employed reaction scheme model is questioned.

8.4. Redox Shuttles


The iodide/triiodide system is the only redox shuttle reported in the literature which yields satisfactory results when used in dye sensitized solar cells. A customized flow system has been constructed that allows easy and rapid measurement and testing of redox shuttles. This work confirms the dominant role of the iodide/triiodide redox shuttle system, as no other alternative has been found that can match its performance in solar cell devices. However, this work also clearly shows that the iodide/triiodide system has several significant shortcomings and does not provide efficient redox shuttling under many conditions so that improvement in this aspect of solar cells is desirable. The reason for the superiority of the iodide couple remains unsolved but is clearly not simple related to the thermodynamic properties of the redox mediator systems. Presumably, kinetic properties of the reactions are of greater importance than the thermodynamic properties generally considered. Specifically the reaction of the reduced redox mediator with the oxidized dye (dye regeneration) and the reaction of the conduction band electron with the oxidized form of the redox mediator (charge recombination), are noted in this context. A recent paper on the "The limiting role of iodide oxidation in cis-Os(dcb)2(CN)2/TiO2 photoelectrochemical cells" by Alebbi and coworkers [73] supports this view. Interestingly, the kinetic properties of [Cr(CO)5I](C4H9)4N, cisMo(CO)2(dpe)(dpm), trans-Cr(CO)2(dpe)2 and cis-Cr(CO)2(dpm)2 have been found to be superior to iodide, unfortunately other properties limit these complexes from being proposed for use in real cell applications.

8.5. Future Outlook


Two of the three major components of dye sensitized solar cells have been studied in this work, these are the sensitizer and the redox shuttle system. The third major component, the semiconductor, and more specifically the TiO2/dye/solution interface, has only been briefly mentioned. For a holistic understanding of the physical and chemical properties of a solar cell, clearly more studies on the semiconductor material and its interface and interaction with the solution environment are required. When better understanding of these three major areas has been achieved it should be possible to relate the value for the reversible potential of the sensitizer to its performance in solar cells. Clearly, additional work will be required in the future to understand the kinetic aspects of chemical reactions occurring during the operation of a solar cell.

Chapter 8: Conclusions Redox Shuttles

page: 220

Despite extensive studies, it still remains unclear what exactly the "desirable" properties of photovoltaic dye sensitizers are in the thermodynamic sense. It is intrinsically clear that a good sensitizer must absorb over a large fraction of the visible spectrum. From long term stability considerations of solar cells, the stability of the oxidized sensitizers should be high. This already is achieved, perhaps empirically, by employing deprotonated forms of the sensitizers and by replacing the (H2-dcbpy)2Ru(NCS)2 sensitizer by the [(H3-tctpy)Ru(NCS)3]. Current overall efficiencies of high-performance dye sensitized solar cells are now slightly in excess of 10%, which implies that further improvements can still be achieved by basic research in this relatively new field of technology. In comparison with the 50 year history of solid state solar cells, the technology and science of dye sensitized solar cells is very young, although the speed of development is very rapid. Nevertheless, it can now be envisaged that the technology will eventually advance which will ultimately allow progress from the R&D stage to the production stage to be achieved.

Reference List

page: 221

Reference List

1. "Solar Energy - Power for the new Millennium." K. Lovegrove and K. Weber, Australian National University - Centre for Sustainable Energy Systems, http://online.anu.edu.au/engn/solar/solarth/ct.html, 23. July 1999 2. "Renewable Energy" B. Srensen, 1st edn., Academic Press, London, 1979. 3. "Solar Power Plants" C.-J. Winter, R.L. Sizmann and L.L. Vant-Hull, Springer-Verlag, Berlin, Heidelberg, 1991. 4. "Photochemical Hydrogen Production" T. Ohta, in "Solar-Hyrdogen Energy Systems", chapter 6, 115-169, edited by N. Kamiya, Pergamon Press Ltd., Oxford, 1979. 5. "The Use of Functionalized Polymers as Photosensitizers in an Energy Storage Reaction." R.R. Hautala, J. Little and E. Sweet, Solar Energy, 1977, 19, 503-508. 6. "Rhodium(I) Catalysis of Vinylcyclopropane Epimerization and Ring Cleavage Rearrangements " R.G. Salomon, M.F. Salomon and J.L.C. Kachinski, J. Am. Chem. Soc., 1977, 99(4), 1043-1054. 7. "Thermal Storage for Solar Power Plants" M.A. Geyer, in "Solar Power Plants", chapter 6, 199-214, editor: C.-J. Winter, R.L. Sizmann and L.L. Vant-Hull, Springer-Verlag, Berlin, Heidelberg, 1991. 8. "Photoelectrochemical and Photocatalytic Methods of Hydrogen Production: A Short Review. " N. Getoff, Int. J. Hydrogen Energy, 1990, 15(6), 407-417. 9. "Hydrogen-Evolving Solar Cells" A. Heller, Science, 1984, 223, 1141-1147. 10. "Solarenergie und Wasserstoff" K. Forster and M. Fuchs, VGB Kraftwerkstechnik 68, 1988, 2, 108-113. 11. "Solar Radiation Conversion" R.L. Sizmann, P. Kpke and R. Busen, in "Solar Power Plants", chapter 2, 17-83, editor: C.-J. Winter, R.L. Sizmann and L.L. Vant-Hull, Springer-Verlag, Berlin, Heidelberg, 1991. 12. "Applied Photovoltaics" S.R. Wenham, M.A. Green and M.E. Watt, Bridge Printery, Sydney, 1994. 13. "Photovoltaic Power Generation" D.L. Pulfrey, 1st edn., Van Nostrand Reinhold Company, New York, 1978. 14. "Crystalline Silicon Solar Cells" A. Goetzberger, J. Knobloch and B. Vo, John Wiley & Sons Ltd., Chichester UK, 1998. 15. "High Efficiency Silicon Solar Cells" M.A. Green, Trans Tech SA, Switzerland, 1987. 16. "Solar Energy Fundamentals and Applications" T.N. Vezirolu, Hemisphere Publishing Corporation, 1987. 17. "Solar Energy Conversion. A Photoelectrochemical Approach" Y.V. Pleskov, Springer Verlag, Berlin Heidelberg, 1990. 18. "Dynamic Aspects of Semiconductor Photoelectrochemistry" L.M. Peter, Chem. Rev., 1990, 90, 753-769. 19. "Thermodynamic Inefficiency of Conversion of Solar Energy to Work" A.W. Adamson, J. Namnath, V.J. Shastry and V. Slawson, J. Chem. Education, 1984, 61(3), 221-224. 20. "Mmoire sur les effets lectriques produits sous l'influence des rayons solaires." E. Becquerel, C. R. Acad. Sci. Paris, 1839, 9, 561-567. 21. "A New Silicon p-n Junction Photocell for Converting Solar Radiation into Electrical Power " D.M. Chapin, C.S. Fuller and G.L. Pearson, J. Appl. Phys., 1954, 25, 676-677. 22. "Photovoltaics: Unlimited Electrical Energy from the Sun" J.L. Stone, Physics Today, 1993, 9, 22-29. 23. "Solar Energy Conversion. The Solar Cell." R.C. Neville, 2nd edn., Elsevier Science B.V., Amsterdam, 1995.

Reference List

page: 222

24. "19.8-Percent efficient honeycomb textured multicrystalline and 24.4-Percent monocrystalline silicon solar cells" J.H. Zhao, A.H. Wang, M.A. Green and F. Ferrazza, Appl. Phys. Lett., 1998, 73(14), 1991-1993. 25. "Sensitisation in photochemistry and photovoltaics." A.J. Mcevoy and M. Grtzel, Sol. Energy Mater. Sol. Cells, 1994, 32(3), 221-227. 26. "Notiz ber Verstrkung photoelektrischer Strme durch potische Sensibilisirung." J. Moser, Monatshefte fr Chemie, 1887, 8, 373-373. 27. "Photochemical Conversion and Storage of Solar Energy" J.S. Connolly, Academic Press, 1981. 28. "Dye sensitised zinc oxide: aqueous electrolyte: platinum photocell" H. Tsubomura, M. Matsumura, Y. Nomura and T. Amamiya, Nature (London), 1976, 261, 402-403. 29. "Conversion of light to electricity by cis-X2bis(2,2'-Bipyridyl-4,4'-dicarboxylate)Ruthenium(II) Charge-transfer sensitizers (X = Cl-, Br-, I-, CN-, and SCN-) on nanocrystalline TiO2 electrodes." M.K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphry-Baker, E. Muller, P. Liska, N. Vlachopoulos and M. Grtzel, J. Am. Chem. Soc., 1993, 115(14), 6382-6390. 30. "Photoelectrochemistry at semiconductors" A.J. Bard and L.R. Faulkner, in "Electrochemical methods", chapter 14.5, 629-642, John Wiley & Sons Inc., 1980. 31. "Long-lived photoinduced charge separation across nanocrystalline TiO2 interfaces." R. Argazzi, C.A. Bignozzi, T.A. Heimer, F.N. Castellano and G.J. Meyer, J. Am. Chem. Soc., 1995, 117(47), 11815-11816. 32. "Insitu ftir study of anodic photoreactions at the n-TiO2 (Anatase) Electrode in aprotic electrolyte solutions." L. Kavan, P. Krtil and M. Grtzel, J. Electroanal. Chem., 1994, 373(1-2), 123-131. 33. "Highly efficient semiconducting TiO2 photoelectrodes prepared by aerosol pyrolysis." L. Kavan and M. Grtzel, Electrochim. Acta, 1995, 40(5), 643-652. 34. "Study of nanocrystalline TiO2 (Anatase) Electrode in the accumulation regime." L. Kavan, K. Kratochvilova and M. Grtzel, J. Electroanal. Chem., 1995, 394(1-2), 93-102. 35. "Nanocrystalline TiO2 (Anatase) Electrodes - surface morphology, adsorption, and electrochemical properties." L. Kavan, M. Grtzel, J. Rathousky and A. Zukal, J. Electrochem. Soc., 1996, 143(2), 394-400. 36. "Photoelectrochemical cells based on dye sensitized colloidal TiO2 layers." R. Kndler, J. Sopka, F. Harbach and H.W. Grnling, Sol. Energy Mater. Sol. Cells, 1993, 30(3) , 277281. 37. "Effect of electron and hole acceptors on the photoelectrochemical behaviour of nanocrystalline microporous TiO2 electrodes" S.K. Poznyak, A.I. Kokorin and A.I. Kulak, J. Electroanal. Chem., 1998, 442(1-2), 99-105. 38. "Structure of nanocrystalline TiO2 powders and precursor to their highly efficient photosensitizer." V. Shklover, M.K. Nazeeruddin, S.M. Zakeeruddin, C. Barbe, A. Kay, T. Haibach, W. Steurer, R. Hermann, H.U. Nissen and M. Grtzel, Chem. Mater., 1997, 9(2), 430-439. 39. "FTIR spectro-photoelectrochemical cell with adjustable solution layer thickness - photocurrent transients at photoexcited TiO2 polycrystalline electrodes" J. Klma, K. Kratochvilov and J. Ludvk, J. Electroanal. Chem., 1997, 427(1-2), 57-61. 40. "Electron transport in porous nanocrystalline TiO2 photoelectrochemical cells." F. Cao, G. Oskam, G.J. Meyer and P.C. Searson, J. Phys. Chem., 1996, 100(42), 17021-17027. 41. "Laser induced photocurrent transients and capacitance measurements on nanocrystalline TiO 2 electrodes" A. Hagfeldt, Sol. Energy Mater. Sol. Cells, 1995, 38, 339-341. 42. "Spectroscopic and excited-state properties of titanium dioxide gels." F.N. Castellano, J.M. Stipkala, L.A. Friedman and G.J. Meyer, Chem. Mater., 1994, 6(11), 2123-2129. 43. "The Isoelectric Points of Solid Oxides, Solid Hydroxides, and Aqueous Hydroxo Complex Systems" G. Parks, Chem. Rev., 1965, 65, 177-198. 44. "Oxyfluoride Photoelectrodes for the Photodecomposition of Water by Solar Energy" A. Wold and K. Dwight, in "Solid State Chemistry: A Contemporary Overview", chapter 9, 161-175, editor: S.L. Holt, J.B. Milstein and M. Robbins, Advances in Chemistry Series, American Chemical Society, Washington, D.C., 1980.

Reference List

page: 223

45. "Photo Responses of Pure and Doped Rutile" J.B. Goodenough, in "Solid State Chemistry: A Contemporary Overview", chapter 6, 113-137, editor: S.L. Holt, J.B. Milstein and M. Robbins, Advances in Chemistry Series, American Chemical Society, Washington, D.C., 1980. 46. "Electrical properties of TiO2 thin films formed on self-assembled organic monolayers on silicon" H. Shin, M.R. Deguire and A.H. Heuer, Journal of Applied Physics, 1998, 83(6), 3311-3317. 47. "Formation of biologically active bone-like apatite on metals and polymers by a biomimetic process " T. Kokubo, Thermochimica Acta, 1996, 280, 479-490. 48. "Light-induced redox reactions in nanocrystalline systems [Review]." A. Hagfeldt and M. Grtzel, Chem. Rev., 1995, 95(1), 49-68. 49. "Nanocrystalline titanium oxide electrodes for photovoltaic applications" C.J. Barbe, F. Arendse, P. Comte, M. Jirousek, F. Lenzmann, V. Shklover and M. Grtzel, Journal of the American Ceramic Society, 1997, 80(12), 3157-3171. 50. "Vectorial Electron Injection into Transparent Semiconductor Membranes and Electric Field Effects on the Dynamics of Light-Induced Charge Separation" B. O'Regan, J. Moser, M. Anderson and M. Grtzel, J. Phys. Chem., 1990, 94, 8720-8727. 51. "Solid-state dye-sensitized mesoporous TiO2 solar cells with high photon-to-electron conversion efficiencies" U. Bach, D. Lupo, P. Comte, J.E. Moser, F. Weissortel, J. Salbeck, H. Spreitzer and M. Grtzel, Nature (London), 1998, 395(6702), 583-585. 52. "Panel Discussion" Twelfth International Conference on Photochemical Conversion and Storage of Solar Energy, Berlin , 1998. 53. "Number of dyes synthesized for solar cell applications." S. Jenkins, 1998, personal communication. 54. "Testing of dye sensitized TiO2 solar cells .1. Experimental photocurrent output and conversion efficiencies." G. Smestad, C. Bignozzi and R. Argazzi, Sol. Energy Mater. Sol. Cells, 1994, 32(3), 259-272. 55. "Reduction of Water to Hydrogen by Reduced Polypyridine Complexes of Ruthenium" H.D. Abrua, A.Y. Teng, G.J. Samuels and T.J. Meyer, J. Am. Chem. Soc., 1979, 101(22), 6745-6746. 56. "Photochemical production of hydrogen and oxygen from water: A review and state of the art " E. Amouyal, Sol. Energy Mater. Sol. Cells, 1995, 38, 249-276. 57. "Electrochemical properties of [(C5me5)RhIII(L)Cl]+ Complexes (L=2,2'-bipyridine or 1,10phenanthroline derivatives) In solution and in related polypyrrolic films - application to electrocatalytic hydrogen generation." S. Chardonnoblat, S. Cosnier, A. Deronzier and N. Vlachopoulos, J. Electroanal. Chem., 1993, 352(12), 213-228. 58. "Some aspects of photochemical systems for direct light-induced hydrogen production" C. Knigstein, J. Photochem. Photobio. A:Chem, 1995, 90, 141-152. 59. "Electron injection by photoexcited Ru(Bpy)32+ Into colloidal SnO2 - analyses of the recombination kinetics based on electrochemical and auger-capture models." W.E. Ford, J.M. Wessels and M.J. Rodgers, J. Phys. Chem. B, 1997, 101(38), 7435-7442. 60. "Photophysical properties of ruthenium polypyridyl photonic SiO2 gels." F.N. Castellano, T.A. Heimer, M.T. Tandhasetti and G.J. Meyer, Chem. Mater., 1994, 6(7), 10411048. 61. "Dynamic electron transfer in aquo- and alco-SiO2 gels." F.N. Castellano and G.J. Meyer, J. Phys. Chem., 1995, 99(40), 14742-14748. 62. "Ru(II) Polypyridine Complexes: Photophysics, Photochemistry, Electrochemistry, and Chemiluminescence" A. Juris, V. Balzani, F. Barigelletti, S. Campagna, P. Belser and A. Zelewsky, Coord. Chem. Rev., 1988, 84, 85-277. 63. "Molecular-level electron transfer and excited state assemblies on surfaces of metal oxides and glass." T.J. Meyer, G.J. Meyer, B.W. Pfennig, J.R. Schoonover, C.J. Timpson, J.F. Wall, C. Kobusch, X.H. Chen, B.M. Peek, C.G. Wall, W. Ou, B.W. Erickson and C.A. Bignozzi, Inorg. Chem., 1994, 33(18), 3952-3964. 64. "Localization of Electronic Excitation Energy in Ru(2,2'-Bipyridine)2(2,2'-Bipyridine-4,4'Dicarboxylicacid)2+ and Related Complexes" J. Ferguson, A.W.H. Mau and W.H.F. Sasse, Chem. Phys. Lett., 1979, 68(1), 21-24.

Reference List

page: 224

65. "Light-Induced Electron-Transfer Reactions Involving the Tris(2,2'-bipyridine)ruthenium Dication and Related Complexes. III. Improved Synthesis of 2,2'-Bipyridine-4,4'-dicarboxylicacid and Photoreduction of Water by Bis(2,2'-bipyridine)(2,2'-bipyridine-4,4'-dicarboxylicacid)ruthenium(II) " A. Launikonis, P.A. Lay, A.W.H. Mau, A.M. Sargeson and W.H.F. Sasse, Aust. J. Chem., 1986, 39, 1053-1062. 66. "Proton Transfer in the Excited State of Carboxylic Acid Derivatives of Tris(2,2'-bipyridineN,N')ruthenium(II)" P.A. Lay and W.H.F. Sasse, Inorg. Chem., 1984, 23, 4123-4125. 67. "Enhanced spectral sensitivity from ruthenium(II) polypyridyl based photovoltaic devices. " R. Argazzi, C.A. Bignozzi, T.A. Heimer, F.N. Castellano and G.J. Meyer, Inorg. Chem., 1994, 33(25), 5741-5749. 68. "Vibrational spectroscopic study of the coordination of (2,2'-Bipyridyl-4,4'-dicarboxylic acid)Ruthenium(II) Complexes to the surface of nanocrystalline titania" K.S. Finnie, J.R. Bartlett and J.L. Woolfrey, Langmuir, 1998, 14(10), 2744-2749. 69. "Importance of binding states between photosensitizing molecules and the TiO2 surface for efficiency in a dye-sensitized solar cell." K. Murakoshi, G. Kano, Y. Wada, S. Yanagida, H. Miyazaki, M. Matsumoto and S. Murasawa, J. Electroanal. Chem., 1995, 396(1-2), 27-34. 70. "cis-Diaquabis(2,2'-bipyridyl-4,4'-dicarboxylate)-ruthenium(II) Sensitizes Wide Band Gap Oxide Semiconductors Very Efficiently over a Broad Spectral Range in the Visible" P. Liska, N. Vlachopoulos, M.K. Nazeeruddin, P. Comte and M. Grtzel, J. Am. Chem. Soc., 1988, 110, 3686-3687. 71. "Electronic Coupling in Cyano-Bridged Ruthenium Polypridine Complexes and Role of Electronic Effects on Cyanide Stretching Frequencies" C.A. Bignozzi, R. Argazzi, J.R. Schoonover, K.C. Gordon, R.B. Dyer and F. Scandola, Inorg. Chem., 1992, 31, 5260-5267. 72. "A low-cost. high-efficiency solar cell based on dye-sensitized colloidal TiO2 films" B. O'Regan and M. Grtzel, Nature (London), 1991, 353, 737-740. 73. "The limiting role of iodide oxidation in cis-Os(dcb)2(CN)2/TiO2 photoelectrochemical cells" M. Alebbi, C.A. Bignozzi, T.A. Heimer, G.M. Hasselmann and G.J. Meyer, J. Phys. Chem. B, 1998, 102(39), 7577-7581. 74. "Molecular level photovoltaics - the electrooptical properties of metal cyanide complexes anchored to titanium dioxide." T.A. Heimer, C.A. Bignozzi and G.J. Meyer, J. Phys. Chem., 1993, 97(46), 11987-11994. 75. "Electron Transfer in Sensitized Nanostructured TiO2 Photovoltaic Cells" T.A. Heimer and G.J. Meyer, Electrochemical Society Proceedings, 1995, 95-8, 167-179. 76. "Efficient light-to-electrical energy conversion with dithiocarbamate ruthenium polypyridyl sensitizers" R. Argazzi, C.A. Bignozzi, G.M. Hasselmann and G.J. Meyer, Inorg. Chem., 1998, 37(18), 4533-4537. 77. "Dynamics of electron injection in nanocrystalline titanium dioxide films sensitized with [Ru(4,4'dicarboxy-2,2'-bipyridine)2(NCS)2] by infrared transient absorption" R.J. Ellingson, J.B. Asbury, S. Ferrere, H.N. Ghosh, J.R. Sprague, T.Q. Lian and A.J. Nozik, J. Phys. Chem. B, 1998, 102(34), 6455-6458. 78. "Charge recombination in dye-sensitized nanocrystalline TiO2 solar cells." S.Y. Huang, G. Schlichthorl, A.J. Nozik, M. Grtzel and A.J. Frank, J. Phys. Chem. B, 1997, 101(14), 2576-2582. 79. "Electrochemical and photoelectrochemical investigation of single-crystal anatase." L. Kavan, M. Grtzel, S.E. Gilbert, C. Klemenz and H.J. Scheel, J. Am. Chem. Soc., 1996, 118(28), 6716-6723. 80. "Ruthenium(II) Charge-transfer sensitizers containing 4,4'-dicarboxy-2,2'-bipyridine - synthesis, properties, and bonding mode of coordinated thio- and selenocyanates." O. Kohle, S. Ruile and M. Grtzel, Inorg. Chem., 1996, 35(16), 4779-4787. 81. "The photovoltaic stability of bis(Isothiocyanato)Ruthenium(II)-Bis-2,2'-bipyridine-4,4'-dicarboxylic acid and related sensitizers." O. Kohle, M. Grtzel, A.F. Meyer and T.B. Meyer, Adv. Mater., 1997, 9(11), 904. 82. "Electron transport properties in dye-sensitized nanoporous-nanocrystalline TiO2 films." H. Lindstrm, H. Rensmo, S. Sdergren, A. Solbrand and S.E. Lindquist, J. Phys. Chem., 1996, 100(8), 3084-3088.

Reference List

page: 225

83. "Structure of organic/inorganic interface in assembled materials comprising molecular components crystal structure of the sensitizer bis[(4,4'-carboxy-2,2'-bipyridine)(thiocyanato)]ruthenium(II) " V. Shklover, Y.E. Ovchinnikov, L.S. Braginsky, S.M. Zakeeruddin and M. Grtzel, Chem. Mater., 1998, 10(9), 2533-2541. 84. "The Dark Current at the Dye-Sensitized TiO2 Electrode" A. Stanley, B. Verity and D. Matthews, Proceedings of the Tenth Australasian Electrochemistry Conference, 1997, 159-171. 85. "Efficient photosensitization of nanocrystalline TiO2 films by a new grass of sensitizer - cisdithiocyanato bis(4,7-dicarboxy-1,10-phenanthroline)ruthenium(II)" H. Sugihara, L.P. Singh, K. Sayama, H. Arakawa, M.K. Nazeeruddin and M. Grtzel, Chemistry Letters, 1998, (10), 1005-1006. 86. "Subpicosecond interfacial charge separation in dye-sensitized nanocrystalline titanium dioxide films. " Y. Tachibana, J.E. Moser, M. Grtzel, D.R. Klug and J.R. Durrant, J. Phys. Chem., 1996, 100(51), 20056-20062. 87. "Spectral characterization of the one-electron oxidation product of cis-bis(Isothiocyanato)Bis(4,4'Dicarboxylato-2,2'-bipyridyl) Ruthenium(II) Complex using pulse radiolysis " S. Das and P.V. Kamat, J. Phys. Chem. B, 1998, 102(45), 8954-8957. 88. "Efficient panchromatic sensitization of nanocrystalline TiO2 films by a black dye based on a trithiocyanato-ruthenium complex." M.K. Nazeeruddin, P. Pechy and M. Grtzel, J. Chem. Soc. , Chem. Commun., 1997, (18), 1705-1706. 89. "Molecular engineering of photosensitizers for nanocrystalline solar cells - synthesis and characterization of ru dyes based on phosphonated terpyridines." S.M. Zakeeruddin, M.K. Nazeeruddin, P. Pechy, F.P. Rotzinger, R. Humphry-Baker, K. Kalyanasundaram, M. Grtzel, V. Shklover and T. Haibach, Inorg. Chem., 1997, 36(25), 5937-5946. 90. "Redox regulation in ruthenium(II) Polypyridyl complexes and their application in solar energy conversion" M.K. Nazeeruddin, E. Muller, R. Humphry-Baker, N. Vlachopoulos and M. Grtzel, J. Chem. Soc. , Dalton Trans., 1997, (23), 4571-4578. 91. "Anorgansiche Chemie: Eine Einfhrung" V. Gutmann and E. Hengge, 4th edn., VCH Verlagsges.m.b.H., Weinheim, 1985. 92. "Inorganic Chemistry" D.F. Shriver, P.W. Atkins and C.H. Langford, 2nd edn., Oxford University Press, 1996. 93. "Measurement of ultrafast photoinduced electron transfer from chemically anchored Ru-dye molecules into empty electronic states in a colloidal anatase TiO2 film." T. Hannappel, B. Burfeindt, W. Storck and F. Willig, J. Phys. Chem. B, 1997, 101(35), 6799-6802. 94. "The electronic structure of the cis-bis(4,4'-dicarboxy-2,2'-bipyridine)-bis(isothiocyanato) ruthenium(II) complex and its ligand 2,2'-bipyridyl-4,4'-dicarboxylic acid studied with electron spectroscopy." H. Rensmo, S. Sodergren, L. Patthey, K. Westermark, L. Vayssieres, O. Kohle, P.A. Bruhwiler, A. Hagfeldt and H. Siegbahn, Chem. Phys. Lett., 1997, 274(1-3), 51-57. 95. "Electrochemical methods" A.J. Bard and L.R. Faulkner, John Wiley & Sons Inc., 1980. 96. "Electrochemistry: Principles, Methods and Applications" C.M.A. Brett and A.M.O. Brett, Oxford University Press, Oxford, 1998. 97. "Elektrochemie I: Leitfhigkeit, Potentiale, Phasengrenzen" C.H. Hamann and W. Vielstich, 2nd edn., VCH Verlagsges.m.b.H., Weinheim, 1985. 98. "Elektrochemie II: Elektrodenprozesse, Angewandte Elektrochemie" C.H. Hamann and W. Vielstich, Verlag Chemie G.m.b.H., Weinheim, 1981. 99. "Electrochemistry" P. Rieger, Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 1987. 100. "Fundamentals of electrochemical science" K.B. Oldham and J.C. Myland, Acadamic Press Inc., 1993. 101. "Laboratory Techniques in Electroanalytical Chemistry" P.T. Kissinger and W.R. Heineman, 2nd edn., Marcel Dekker Inc., 1996. 102. "Chemie" V.W. Schrter, K.-H. Lautenschlger, H. Bibrack and A. Schnabl, 17th edn., VEB Fachbuchverlag, Leipzig, 1986. 103. "Famous Chemists: The Men and Their Work" W.A. Tilden, George Routledge & Sons, Ltd., London, 1935.

Reference List

page: 226

104. "Physikalische Chemie: Allgemeine Grundlagen" M. Suard, B. Praud and L. Praud, Georg Thieme Verlag, Stuttgart, 1976. 105. "The Absolute Electrode Potential: An Explanatory Note" S. Trasatti, Pure & Appl. Chem., 1986, 58(7), 955-966. 106. "On-line determination of vanadium by adsorptive stripping voltammetry" G.M. Greenway and G. Wolfbauer, Anal. Chim. Acta, 1995, 312(1), 15-25. 107. "Amperometric and voltammetric electroanalysis." C.M.A. Brett and A.M.O. Brett, in "Electrochemistry: Principles, Methods and Applications", chapter 14.6, 322-325, Oxford University Press, Oxford, 1998. 108. "Rotating Disk Electrode" A.J. Bard and L.R. Faulkner, in "Electrochemical methods", chapter 8.3, 283-298, John Wiley & Sons Inc., 1980. 109. "Practical Problems in Voltammetry 3: Reference Electrodes for Voltammetry" A.W. Bott, Current Separations, 1995, 14(2), 64-68. 110. "Conditions Under Which the Oxidation of Ferrocene Can Be Used as a Standard Voltammetric Reference Process in Aqueous Media" A.M. Bond, E.A. McLennan, R.S. Stojanovic and F.G. Thomas, Anal. Chem., 1987, 59, 2853. 111. "Nonaqueous Solvents for Electrochemical Use" C.K. Mann, Electroanal. Chem., 1969, 3, 57-134. 112. "Solvents and Supporting Electrolytes" A.J. Fry, in "Laboratory Techniques in Electroanalytical Chemistry", chapter 15, 469-486, 2nd edn., editor: P.T. Kissinger and W.R. Heineman, Marcel Dekker Inc., 1996. 113. "Controlled Potential Microelectrode Techniques - Potential Sweep Methods" A.J. Bard and L.R. Faulkner, in "Electrochemical methods", chapter 6, 213-248, John Wiley & Sons Inc., 1980. 114. "Theory of Stationary Electrode Polarography: Single Scan and Cyclic Methods Applied to Reversible, Irreversible, and Kinetic Systems" R.S. Nicholson and I. Shain, Anal. Chem., 1964, 36(4), 706-723. 115. "Theory of ultramicroelectrodes [Review]." K. Aoki, Electroanalysis, 1993, 5(8), 627-639. 116. "The Application of Microdiskelectrodes to the Study of Homogeneous Processes Coupled to Electrode Reactions Part I. EC' and CE Reactions" M. Fleischmann, F. Lasserre, J. Robinson and D. Swan, J. Electroanal. Chem., 1984, 177, 97-114. 117. "Steady-State Voltammetry" K.B. Oldham and J.C. Myland, in "Fundamentals of electrochemical science", chapter 8, 263-308, Acadamic Press Inc., 1993. 118. "Physicochemical hydrodynamics" V.G. Levich, Prentice-Hall, Englewood Cliffs, New Jersey, 1962. 119. "Innovations in the Study of Adsorbed Reactants by Chronocoulometry" F.C. Anson, Anal. Chem., 1966, 38(1), 54-57. 120. "The Application of Double Potential-Step Chronocoulometry to the Study of Chemical Reactions Following the Electrode Reaction. Theory" J.H. Christie, J. Electroanal. Chem., 1967, 13, 79-89. 121. "Determination of inorganic ionic mercury down to 5x10-14 mol l-1 By differential pulse anodic stripping voltammetry" S. Meyer, F. Scholz and R. Trittler, Fresenius Journal of Analytical Chemistry , 1996, 356(3-4), 247252. 122. "Spectroelectrochemical Studies on Tris-bipyridyl Ruthenium Complexes; Ultraviolet, Visible. and Near-Infrared Spectra of the Series [Ru(bpy)3]2+/+/0/1-" G.A. Heath, L.J. Yellowlees and P.S. Braterman, J. Chem. Soc. , Chem. Commun., 1981, 287-289. 123. "Spectroelectrochemical Studies on Tris(bipyridyl)iridium Complexes: Ultraviolet, Visible. and NearInfrared Spectra of the Series [Ir(bpy)3]3+/2+/+/0" V.T. Coombe, G.A. Heath, A.J. MacKenzie and L.J. Yellowlees, Inorg. Chem., 1984, 23, 3423-3425. 124. "Case Histories" E.A.V. Ebsworth, D.W.H. Rankin and S. Cradocj, in "Structural Methods in Inorganic Chemistry", chapter 10, 395-460, 2 edn., Blackwell Scientific Publications, 1994. 125. "Structural Methods in Inorganic Chemistry" E.A.V. Ebsworth, D.W.H. Rankin and S. Cradocj, 2nd edn., Blackwell Scientific Publications, 1994.

Reference List

page: 227

126. "Spektroskopische Methoden in der organischen Chemie" M. Hesse, H. Meier and B. Zeeh, 3 edn., Georg Thieme Verlag Stuttgart, New York, 1987. 127. "Thin Polymer and Phospholipid Films for Biosensors" T. Tjrnhage, PhD Thesis, Department of Analytical Chemistry, Ume University, Sweden (http://www.anachem.umu.se/~tte/Thesis.html), 1996. 128. "Verwendung von Schwingquarzen zur Wgung dnner Schichten und zur Mikrowgung." G. Sauerbrey, Zeitschrift fr Physikalische Chemie, 1959, 155, 206-222. 129. "Piezoelectric Crystals as Detectors in Liquid Chromatography" P.L. Konash and G.J. Bastiaans, Anal. Chem., 1980, 52, 1929-1931. 130. "Single-drop method for determination of cyanide in solution with a piezoelectric quartz crystal." T. Nomura, Anal. Chim. Acta, 1981, 124, 81-84. 131. "Solvent effect on the photoinduced electron-transfer reactions between dicyanobis(polypyridine)ruthenium(II) complexes and tris(-diketonato)ruthenium(III) complexes." N. Sonoyama, O. Karasawa and Y. Kaizu, J. Chem. Soc. , Faraday Trans., 1995, 91(3), 437-443. 132. "Absorption, Magnetic Circular Dichroism and Magnetic Circular Polarization of Luminescence Studies of Ru(bpy)32+ and Complexes with a Di(ethoxycarbonyl)-substituted Bipyridine Ligand as a Probe of Rigid Environments" E. Krausz, J. Chem. Soc. , Faraday Trans. 1, 1988, 84(3), 827-840. 133. "Solid state electrochemically generated luminescence based on serial frozen concentration gradients of RuIII/II and RuII/I couples in a molten ruthenium 2,2'-bipyridine complex." K.M. Maness, H. Masui, R.M. Wightman and R.W. Murray, J. Am. Chem. Soc., 1997, 119(17), 39873993. 134. "Novel sensitisers for photovoltaic cells - structural variations of Ru(II) Complexes containing 2,6bis(1-Methylbenzimidazol-2-yl)Pyridine." S. Ruile, O. Kohle, P. Pechy and M. Grtzel, Inorg. Chim. Acta, 1997, 261(2), 129-140. 135. "Proton-induced tuning of electrochemical and photophysical properties in mononuclear and dinuclear ruthenium complexes containing 2,2'-bis(Benzimidazol-2-yl)-4,4'-Bipyridine - synthesis, molecular structure, and mixed-valence state and excited-state properties." M.A. Haga, M.M. Ali, S. Koseki, K. Fujimoto, A. Yoshimura, K. Nozaki, T. Ohno, K. Nakajima and D.J. Stufkens, Inorg. Chem., 1996, 35(11), 3335-3347. 136. "Remote interfacial electron transfer from supramolecular sensitizers." R. Argazzi, C.A. Bignozzi, T.A. Heimer and G.J. Meyer, Inorg. Chem., 1997, 36(1), 2-3. 137. "Structural investigations of the catalytic mechanisms of water oxidation by the [(Bpy)2Ru(OH2)]2O4+ Ion." Y.B. Lei and J.K. Hurst, Inorg. Chem., 1994, 33(20), 4460-4467. 138. "Verification of high efficiencies for the Grtzel-Cell - A 7-percent efficient solar cell based on dyesensitized colloidal TiO2 films." A. Hagfeldt, B. Didriksson, T. Palmqvist, H. Lindstrom, S. Sodergren, H. Rensmo and S.E. Lindquist, Sol. Energy Mater. Sol. Cells, 1994, 31(4), 481-488. 139. "Testing of dye sensitized TiO2 solar cells .2. Theoretical voltage output and photoluminescence efficiencies." G. Smestad, Sol. Energy Mater. Sol. Cells, 1994, 32(3), 273-288. 140. "Independant Control of Charge-Transfer and Metal-Centered Excited States in Mixed-Ligand Polypyridine Ruthenium(II) Complexes via Specific Ligand Design" W.F. Wacholtz, R.A. Auerbach and R.H. Schmehl, Inorg. Chem., 1986, 25, 227-234. 141. "Stability and Response Studies of Multicolor Electrochromic Polymer Modified Electrodes Prepared from Tris(5,5'-dicarboxyester-2,2'-bipyridine)ruthenium(II)" C.M. Elliott and J.G. Redepenning, J. Electroanal. Chem., 1986, 197, 219-232. 142. "Electrochemical and Spectral Investigations of Ring-Substituted Bipyridine Complexes of Ruthenium" C.M. Elliott and E.J. Hershenhart, J. Am. Chem. Soc., 1982, 104, 7519-7526. 143. "Electrochemistry and Near Infrared Spectroscopy of Tris(4,4'-dicarboxyethyl-2,2'bipyridine)ruthenium(II)" C.M. Elliott, J. Chem. Soc. , Chem. Commun., 1980, 261-262. 144. "Molecular Orbital Study of the Substituent Effect on the Redox Properties of Disubstituted Bipyridines and Their Ruthenium(II) Complexes" Y. Ohsawa, M.H. Whangbo, K.W. Hanck and M.K. DeArmond, Inorg. Chem., 1984, 23, 3426-3428. 145. "Spatially Isolated Redox Orbitals: Evidence from Low-Temperature Voltammetry" Y. Ohsawa, M.K. DeArmond, K.W. Hanck, D.E. Morris, D.G. Whitten and P.E. Neveux Jr, J. Am. Chem. Soc., 1983, 105(21), 6522-6524.

Reference List

page: 228

146. "Spectroscopic Study of the Parent and Reduction Products of Some Substituted Bipyridine Complexes of Iron(II) and Osmium(II). 1. Substitution at the 5,5' Positions" R.J. Donohoe, C.D. Tait, M.K. DeArmond and D.W. Wertz, J. Phys. Chem., 1986 , 90, 3923-3926. 147. "Spectroscopic Study of the Parent and Reduction Products of Some Substituted Bipyridine Complexes of Iron(II) and Osmium(II). 2. Substitution at the 4,4' Positions" R.J. Donohoe, C.D. Tait, M.K. DeArmond and D.W. Wertz, J. Phys. Chem., 1986 , 90, 3927-3930. 148. "Excited state redox potentials of ruthenium diimine complexes - correlations with ground state redox potentials and ligand parameters" A.A. Vlcek, E.S. Dodsworth, W.J. Pietro and A.B.P. Lever, Inorg. Chem., 1995, 34(7), 1906-1913. 149. "Medium effects on the redox properties of tris(2,2'-bipyridyl)ruthenium complexes" M.R. McDevitt and A.W. Addison, Inorg. Chim. Acta, 1993, 204, 141-146. 150. "Electrogenerated Chemiluminescence. XIII. Electrochemical and Electrogenerated Chemiluminescence Studies of Ruthenium Chelates" N.E. Tokel-Takvoryan, R.E. Hemingway and A.J. Bard, J. Am. Chem. Soc., 1973, 95(20), 6582-6589. 151. "Ligand-Ligand inter-valence Charge-Transfer Absorption in Reduced Ruthenium(II) Biypridine Complexes" G.A. Heath, L.J. Yellowlees and P.S. Braterman, Chem. Phys. Lett., 1982, 92(6), 646-648. 152. "ESR studies of the Redox Orbitals in Diimine Complexes of Iron(II) and Ruthenium(II)" D.E. Morris, K.W. Hanck and M.K. DeArmond, J. Am. Chem. Soc., 1983, 105, 3032-3038. 153. "ESR of Reduction Products of [Fe(bpy)3]2+ and [Ru(bpy)3]2+" A.G. Motten, K.W. Hanck and M.K. DeArmond, Chem. Phys. Lett., 1981, 79(3), 541-546. 154. "Comprehensive Coordination Chemistry" M. Schrder and T.A. Stephenson, in "Ruthenium", chapter 45, 277-499, edited by G. Wilkinson, R.D. Gillard and J.A. McCleverty, Pergamon Press, 1989. 155. "Electrochemical and ESR Studies of [Fe(terpy)2]2+ and [Ru(terpy)2]2+ and their Reduction Products" D.E. Morris, K.W. Hanck and M.K. DeArmond, J. Electroanal. Chem., 1983, 149, 115-130. 156. "Configuration-dependent electron hopping between equivalent ligands in isomeric ruthenium(II) Complexes." M. Heilmann, F. Baumann, W. Kaim and J. Fiedler, J. Chem. Soc. , Faraday Trans., 1996, 92(21), 4227-4231. 157. "Stepwise Ligand-Additvity Effects on Electrode Potentials and Charge-Transfer Spectra in Hexahalide, Mixed Halide / Nitrile, and Hexakis(nitrile) Complexes of Ruthenium(IV), -(III), and -(II) " C.M. Duff and G.A. Heath, Inorg. Chem., 1991, 30, 2528-2535. 158. "Mononuclear ruthenium(III) Complexes of the type LRuX3 (X=Cl-, NCO-, NCS-, N3- L=1,4,7trimethyl-1,4,7-triazacyclononane) [German]." R. Schneider, T. Justel, K. Wieghardt and B. Nuber, Z. Naturforsch. , B, 1994, 49(3), 330-336. 159. "Electronic spectra of complexes" D.F. Shriver, P.W. Atkins and C.H. Langford, in "Inorganic Chemistry", chapter 14, 579-617, 2 edn., Oxford University Press, 1996. 160. "Synthesis, characterization and electron transfer properties of some picolinate complexes of ruthenium." N. Ghatak, J. Chakravarty and S. Bhattacharya, Polyhedron, 1995, 14(23-24), 3591-3597. 161. "Magnetic Resonance Studies of Some Low-Spin d5 Tri Diimine Complexes" R.E. DeSimone and R.S. Drago, J. Am. Chem. Soc., 1970, 92(8), 2343-2352. 162. "Mssbauer Studies of Iron Organometallic Complexes-III Octahedral Comples " R.L. Collins, R. Petit and W.A. Baker, J. Inorg. Nucl. Chem., 1966, 28, 1001-1010. 163. "Molecular orbital Interpretation of X-Ray Absorption Edges" W. Seka and H.P. Hanson, J. Chem. Phys., 1969, 50(1), 344-350. 164. "Electrochemical and X-ray Diffraction Study of the Redox Cycling of Nanocrystals of 7,7,8,8tetracyanoquinodimethane. Observation of a Solid-solid Phase Transformation Controlled by Nucleation and Growth" A.M. Bond, S. Fletcher, F. Marken, S.J. Shaw and P.G. Symons, J. Chem. Soc. , Faraday Trans., 1996, 92(20), 3925-3933. 165. "The relationship between the electrochemistry and the crystallography of microcrystals - the case of tcnq (7,7,8,8-Tetracyanoquinodimethane)" A.M. Bond, S. Fletcher and P.G. Symons, Analyst , 1998, 123(10), 1891-1904.

Reference List

page: 229

166. "Oxidative Electrochemistry of Compounds of the Type (5-C5H5)2MX2 Where M = Ti(IV), Mo(IV), and W(IV) and X = Halide, Thiolate, or Ferrocenyl " J.C. Kotz, W. Vining, W. Coco, R. Rosen, A.R. Dias and M.H. Garcia, Organometallics, 1983, 2, 6879. 167. "Redox behavior of edge-shared bioctahedral [Re-2(-NCS)2(NCS)8]3- And its relationship to the direct metal-metal-bonded ion [Re2(NCS)8]2- - A spectroelectrochemical study" R.H. Clark and D.G. Humphrey, Inorg. Chem., 1996, 35(7) , 2053-2061. 168. "Oxidation of the Ligand in Nitro Complexes of Ruthenium(III)" F.R. Keene, D.J. Salmon, J.L. Walsh, H.D. Abrua and T.J. Meyer, Inorg. Chem., 1980, 19, 18961903. 169. "Solvent and Electrode Kinetic Effects on Cathodic Reduction of I-3+2e->3I- in Some Pure and Mixed Dipolar Aprotic Solvents" S. Bhattacharya and K.K. Kundu, Bull. Chem. Soc. Jpn., 1989, 62(8), 2676-2683. 170. "Voltammetric Evaluation of the Stability of Trichloride, Tribromide, and Triiodide ions in Nitromethane, Acetone, and Acetonitrile" I.V. Nelson and R.T. Iwamoto, J. Electroanal. Chem., 1964, 7, 218-221. 171. "Carbon: Group IVA(14)" F.A. Cotton and G. Wilkinson, in "Advanced Inorganic Chemistry", chapter 8, 234-264, 5 edn., John Wiley & Sons Inc., 1988. 172. "Kinetics and equilibria in the nitric acid nitrous acid sodium thiocyanate system. " E. Jones, C.G. Munkley, E.D. Phillips and G. Stedman, J. Chem. Soc. , Dalton Trans., 1996, (9), 19151920. 173. "Electrochemical Parametrization of Metal Complex Redox Potentials, Using the Ruthenium(III)/Ruthenium(II) Couple To Generate a Ligand Electrochemical Series" A.B.P. Lever, Inorg. Chem., 1990, 29, 1271-1285. 174. "Spectro-Electrochemistry and Electrochemical Spectroscopy" G.A. Heath, in "Molecular Electrochemistry of Inorganic, Bioinorganic and Organometallic Compounds", 533-547, edited by A.J.L. Pombeiro and J.A. McCleverty, Kluwer Academic Publishers, The Netherlands, 1993. 175. "Chemical redox agents for organometallic chemistry [Review]." N.G. Connelly and W.E. Geiger, Chem. Rev., 1996, 96(2), 877-910. 176. "Electrochemical formation of dianions of 2,2'-bipyridine and related compounds" M. Krejcik and A.A. Vlcek, J. Electroanal. Chem., 1991, 313, 243-257. 177. "Electrochemical reduction of pyridine- and benzene-substituted N-alkyl esters and thioic S-esters in acetonitrile" R.D. Webster and A.M. Bond, J. Org. Chem., 1997, 62(6), 1779-1787. 178. "Electrochemical reduction of some 2,6-disubstituted pyridine-based esters and thioic S-esters in acetonitrile" R.D. Webster, A.M. Bond and T. Schmidt, J. Chem. Soc. , Perkin Trans., 1995, (7), 1365-1374. 179. "Electrochemical reduction of pyrethroid insecticides in non-aqueous solvents" D.C. Coomber, D.J. Tucker and A.M. Bond, J. Electroanal. Chem., 1997, 426(1-2), 63-73. 180. "Voltammetric determination of the insect repellent dipropyl pyridine-2,5-dicarboxylate in nonaqueous solvents" D.C. Coomber, D.J. Tucker and A.M. Bond, Analyst, 1997, 122(12), 1587-1592. 181. "Chemistry of Highly Reduced Polypyridyl-Metal Complexes. Anion Subtitution Induced by LigandBased Reduction" B.P. Sullivan, D. Conrad and T.J. Meyer, Inorg. Chem., 1985, 24(22), 3640-3645. 182. "ESR Characterization of the Redox Orbitals in [Os(bpy)3]2+" D.E. Morris, K.W. Hanck and M.K. DeArmond, Inorg. Chem., 1985, 24, 977-979. 183. "Multiple State Emission and Related Phenomena in Transition Metal Complexes" M.K. DeArmond and C.M. Carlin, Coord. Chem. Rev., 1981, 36, 325-355. 184. "The lower excited electronic states of singly and doubly reduced 2,2'-bipyridine" E. Knig and S. Kremer, Chem. Phys. Lett., 1970, 5(2), 87-90. 185. "The Transition Metal Coordination Chemistry of Anion Radicals" W. Kaim, Coord. Chem. Rev., 1987, 76, 187-235. 186. "Reaction mechanism of d-block complexes" D.F. Shriver, P.W. Atkins and C.H. Langford, in "Inorganic Chemistry", chapter 15, 618-658, 2 edn., Oxford University Press, 1996.

Reference List

page: 230

187. "Comparative bonding and photophysical properties of 2,2'-bipyridine and 2,2'-bipyrazine in tetracyano complexes containing ruthenium and osmium." M.G. Posse, N.E. Katz, L.M. Baraldo, D.D. Polonuer, C.G. Colombano and J.A. Olabe, Inorg. Chem., 1995, 34(7), 1830-1835. 188. "Intervalence-transfer absorption. Part2. Theoretical considerations and spectroscopic data " N.S. Hush, Prog. Inorg. Chem., 1967, 8, 391-444. 189. "Effects of Adsorption of Electroactive Species in Stationary Electrode Polarography" R.H. Wopschall and I. Shain, Anal. Chem., 1967, 39(13), 1514-1527. 190. "Adsorption Characteristics of the Methylen Blue System Using Stationary Electrode Polarography" R.H. Wopschall and I. Shain, Anal. Chem., 1967, 39(13), 1527-1534. 191. "Adsorption Effects in Stationary Electrode Polarography with a Chemical Reaction Following Charge Transfer" R.H. Wopschall and I. Shain, Anal. Chem., 1967, 39(13), 1535-1542. 192. "Beitrag zur Konsitution anorganischer Verbindungen. ber rhodanatokobaltiake und strukturisomere Salze." A. Werner, H. Mller, R. Klien and F. Brunlich, Anorg. Chem., 1900, 91-157. 193. "Measurement of Interfacial Processes at Electrode Surfaces with the Electrochemical Quartz Crystal Microbalance" D.A. Buttry and M.D. Ward, Chem. Rev., 1992, 92, 1355-1379. 194. "Ru(2,2'-bpy)2(NCS)2.X [X = CH3CN, (CH3)2SO] and Related Compounds: Crystal Structure, VTFTIR, and NMR Study" R.H. Herber, G. Nan, J.A. Potenza, H.J. Schugar and A. Bino, Inorg. Chem., 1989, 28, 938-942. 195. "Cyanate and thiocyanate adsorption at copper and gold electrodes as probed by in situ infrared and surface-enhanced raman spectroscopy." M. Bron and R. Holze, J. Electroanal. Chem., 1995, 385(1), 105-113. 196. "Adsorption of Bipyridyls and Structurally Related Compounds at Pt(111) Electrodes: Studies by Vibrational Spectroscopy (EELS), Auger Spectroscopy, and Electrochemistry" S.A. Chaffins, J.Y. Gui, B.E. Kahn, C.H. Lin, F. Lu, G.N. Salita, D.A. Stern, D.C. Zapien, A.T. Hubbard and C.M. Elliott, Langmuir, 1990, 6, 957-970. 197. "Investigations of the Mass Transport Process in the Voltammetry of Cytochrome c at 4,4'-Bipyridyl Disulfide Modified Stationary and Rotated Macro-and Microdisk Gold Electrodes" A.M. Bond, H.A.O. Hill, S. Kamorsky-Lovric, M. Lovric, M.E. McCarthy, I.S. Psalti and N.J. Walton, J. Phys. Chem., 1992, 96(20), 8100-8105. 198. M.J. Eddows and H.A.O. Hill, J. Chem. Soc. , Chem. Commun., 1977, 771-772. 199. D.A. Palmer and D.W. Watts, Inorg. Chem., 1971, 10, 281. 200. "Photoanation of the Tris(2,2'-bipyridine)ruthenium(II) Cation by Thiocyanate" P.E. Hoggard and G.B. Porter, J. Am. Chem. Soc., 1978, 100(5), 1457-1463. 201. "Handbook of Chemistry and Physics" 70th edn., editors R.C. Weast and D.R. Lide, CRC Press, 1989. 202. "Low cost and efficient photovoltaic conversion by nanocrystalline solar cells." M. Grtzel, Chem. -Ing. -Tech., 1995, 67(10), 1300-1305. 203. "Ruthenium 535 TBA" Solaronix Inc., Aubonne, CH, http://www.solaronix.ch/products/ruthenium535tba.html, 21. June 1999 204. "Acid-Base Behavior in the Ground and Ecited States of Ruthenium(II) Complexes Containing Tetraimines or Dicarboxybipyridines as Protatable Ligands" M.K. Nazeeruddin and K. Kalyanasundaram, Inorg. Chem., 1989, 28, 4251-4259. 205. "Electrochemistry of Symmetrical and Asymmetrical Dinuclear Ruthenium, Osmium, and Mixed-Metal 2,2'-Bipyridine Complexes Bridged by 2,2'-Bibenzimidazolate" M.A. Haga and A.M. Bond, Inorg. Chem., 1991, 30, 475-480. 206. "Electrorganic Preparations" H. Lund, Acta Chem. Scand., 1963, 17(4), 972-978. 207. "Polarographie aromatischer heterocyclischer Verbindungen II. Polarographisches Verhalten der Isonicotinsure und Picolinsure" J. Volke and V. Volkov, Coll. Czech. Chem. Comm., 1955, 20, 1332-1339. 208. "Polarographie aromatischer heterocyclischer Verbindungen II. Oszillographische Unterscheidung einiger Pyridinederivate" J. Volke and V. Volkov, Coll. Czech. Chem. Comm., 1955, 20, 908-916.

Reference List

page: 231

209. "Influence de l'acidit du milieu sur la rduction polarographique des acides pyridinedicarboxyliques" C. Tissier and M. Agoutin, J. Electrochem. Soc., 1973, 47, 499-508. 210. "Reduction of picolinic acid and dipicolinic acid at mercury cathodes" O.R. Brown, J.A. Harrison and K.S. Sastry, J. Electroanal. Chem., 1975, 58, 387-391. 211. "Polarographic and Oscillopolarographic Behavior of Dipicolinic Acid" L. Campanella, E. Chiacchierini and M. Palchetti, Revue Roumaine de Chimie, 1972, 17(4), 647-660. 212. "Polarographic Behavior of Isochinchomeronic Acid" L. Campanella, P.L. Cignini and G. de Angelis, Revue Roumaine de Chimie, 1973, 18(9), 1649-1657. 213. "Determination of Nicotinic and Isochinchomeronic Acids in Mixtures" V.A. Serazetdinova and B.V. Suvorov, Zhurnal Analiticheskoi Khimii, 1978, 33(1), 964-965. 214. "Relative energetics at the semiconductor sensitizing dye electrolyte interface" A. Zaban, S. Ferrere and B.A. Gregg, J. Phys. Chem. B, 1998, 102(2), 452-460. 215. "Titrations in Nonaqueous Solvents" C.A. Streuli, Anal.Chem., 1964, 36(5), 363-369 216. "Electrochemical formation and spectroselectrochemical characterization of organometallic [Ru(L)(CO)2]n polymers; L = disubstituted-2,2'-bipyridine." C. Caix-Cecillon, S. Chardon-Noblat, A. Deronzier, M. Haukka, T.A. Pakkanen, R. Ziessel and D. Zsoldos, J. Electroanal. Chem., 1999, 466(1), 187-196. 217. "Measurement of Charge Passed Following Application of a Potential Step-Application to the Study of Electrode Reactions and Adsorption" J.H. Christie, G. Lauer and R.A. Osteryoung, J. Electroanal. Chem., 1964, 7, 60-72. 218. "Effect of Adsorption of Electroactive Species on the Electrochemical Response" A.J. Bard and L.R. Faulkner, in "Electrochemical methods", chapter 12.5, 519-538, John Wiley & Sons Inc., 1980. 219. "Spectroscopic determination of flatband potentials for polycrystalline TiO2 electrodes in mixed solvent systems" B. Enright, G. Redmond and D. Fitzmaurice, J. Phys. Chem., 1994. 220. "Applications of functionalized transition metal complexes in photonic and optoelectronic devices [Review]" K. Kalyanasundaram and M. Grtzel, Coord. Chem. Rev., 1998, 177, 347-414. 221. "Molecular photovoltaics" J.E. Moser, P. Bonnote and M. Grtzel, Coord. Chem. Rev., 1998, 171, 245-250. 222. "An acetylacetonate-based semiconductor-sensitizer linkage" T.A. Heimer, S.T. Darcangelis, F. Farzad, J.M. Stipkala and G.J. Meyer, Inorg. Chem., 1996, 35(18), 5319-5324. 223. "Designed synthesis of mononuclear tris(heteroleptic) Ruthenium complexes containing bidentate polypyridyl ligands." P.A. Anderson, G.B. Deacon, K.H. Haarmann, F.R. Keene, T.J. Meyer, D.A. Reitsma, B.W. Skelton, G.F. Strouse, N.C. Thomas, J.A. Treadway and A.H. White, Inorg. Chem., 1995, 34(24), 6145-6157. 224. "Voltammetric Determination of the Reversible Redox Potential for the Oxidation of the Highly Surface Active Polypyridyl Ruthenium Photovoltaic Sensitizer RuII(dcbpy)2(NCS)2 (dcbpy = 2,2'bipyridine-4,4'-dicarboxylic acid)" A.M. Bond, G.B. Deacon, J. Howitt, D.R. MacFarlane, L. Spiccia and G. Wolfbauer, J. Electrochem. Soc. , 1999, 146(2), 648-656. 225. "Mechanisms of instability in Ru-based dye sensitization solar cells." R. Grnwald and H. Tributsch, J. Phys. Chem. B, 1997, 101(14), 2564-2575. 226. "A plasticized polymer-electrolyte-based photoelectrochemical solar cell" D. Mao, M.A. Ibrahim and A.J. Frank, J. Electrochem. Soc., 1998, 145(1), 121-124. 227. "Morphology and adsorbate dependence of ionic transport in dye sensitized mesoporous TiO2 films" N. Papageorgiou, C. Barbe and M. Grtzel, J. Phys. Chem. B, 1998, 102(21), 4156-4164. 228. "An iodine/triiodide reduction electrocatalyst for aqueous and organic media." N. Papageorgiou, W.F. Maier and M. Grtzel, J. Electrochem. Soc., 1997, 144(3), 876-884. 229. "Applications of Microelectrodes in Kinetics" M.I. Montenegro, in "Research in Chemical Kinetics, Volume 2", 1-80, editor: R.G. Compton and G. Hancock, Elsevier, 1994. 230. "Resistance of Nonaqueous Solvent Systems Containing Tetraalkylammonium Salts. Evaluation of Heterogeneous Electron Transfer Rate Constants for the Ferrocene Couple" K.M. Kadish, J.Q. Ding and T. Malinski, Anal. Chem., 1984, 56, 1741-1744.

Reference List

page: 232

231. "Nucleation and Charge-Transfer Kinetics at the Viologen/SnO2 Interface in Electrochromic Device Applicatons." S. Fletcher, L. Duff and R.G. Barradas, J. Electroanal. Chem., 1979, 100, 759-770. 232. "Electrical and optical properties of porous nanocrystalline TiO2 films" F. Cao, G. Oskam, P.C. Searson, J.M. Stipkala, T.A. Heimer, F. Farzad and G.J. Meyer, J. Phys. Chem., 1995, 99(31), 11974-11980. 233. "Models for Double-Layer Structure" A.J. Bard and L.R. Faulkner, in "Electrochemical methods", chapter 12.3, 500-514, John Wiley & Sons Inc., 1980. 234. "Photocurrent losses in nanocrystalline/nanoporous TiO2 electrodes due to electrochemically active species in the electrolyte." H. Rensmo, H. Lindstrom, S. Sodergren, A.K. Willstedt, A. Solbrand, A. Hagfeldt and S.E. Lindquist, J. Electrochem. Soc., 1996, 143(10), 3173-3178. 235. "The dark current at the TiO2 electrode of a dye-sensitized TiO2 photovoltaic cell" A. Stanley and D. Matthews, Aust. J. Chem., 1995, 48(7) , 1293-1300. 236. "The Nature of the Chemical Bond and the Structure of Molecules and Crystals; An Introduction to Modern Structural Chemistry." L.C. Pauling, 3rd edn., Cornell University Press, 1960. 237. "Electrolyte Solutions : The Measurement and Interpretation of Conductance, Chemical Potential and Diffusion in Solutions of Simple Electrolytes" R.A. Robinson and R.H. Stokes, Butterworths Scientific Publications, London, 1959. 238. "Book of Abstracts, Twelfth International Conference of Solar Energy Conversion and Storage" H. Tributsch and F. Willig, Hahn-Meitner-Institute, Berlin, 1998. 239. "Relative Standard Electrode Potentials of I-3/I-, I2/I-3, and I 2/I- Redox Couples and the related Formation Constants of I-3 in Some Pure and Mixed Dipolar Aprotic Solvents" J. Datta, S. Bhattacharya and K.K. Kundu, Bull. Chem. Soc. Jpn., 1987, 61(5), 1735-1742. 240. "An efficient dye-sensitized photoelectrochemical solar cell made from oxides of tin and zinc " K. Tennakone, G.A. Kumara, I.M. Kottegoda and V.S. Perera, Chemical Communications, 1999, (1), 15-16. 241. "Electrochemical oxidation of thiocyanate in a two-phase electrolyte." P. Krishnan and V.G. Gurjar, J. Appl. Electrochem., 1995, 25(8), 792-796. 242. "Physikalische Chemie" P.W. Atkins, 2nd edn., VCH Verlagsges.m.b.H., Weinheim, 1990.

Curriculum Vitae: Georg Wolfbauer

page: 233

Curriculum Vitae

Georg Wolfbauer
Leifhelmgasse 11b, A-1140 Wien, Austria Tel: ++43-1-9145265

Name: Nationality: Marital status:

Georg Wolfbauer, Born on 23.7.68 in Vienna / Austria Austrian, Australian permanent resident. Married to Claudia Wolfbauer, intellectual and psychiatric disability worker

Education:
78 - 86 86 - 88 88 - 94 93 - 94 Gymnasium BRG VII, Kandlgasse, 1070 Vienna, mathematical direction. University Vienna, Computer Science and Business Administration. Technical University Vienna, Technical Chemistry, Specialized in Organic Chemistry. Diploma Thesis at the University of Hull/England, School of Chemistry, Title OnLine Determination of Vanadium by Adsorptive Stripping Voltammetry under the supervision of Dr. G.M.Greenway, Prof. A. Townshend and Prof. M. Grasserbauer (Vienna). Diploma examination passed with highest distinction. Graduation to "Diploma Engineer" (M.Sc.). Vienna Institute of Commerce, 'Quality Management', accredited Internal Auditor and ISO 9000 Manager according to EN 45013. PhD on the "Electrochemistry of Dye Sensitized Solar Sells, Their Sensitizers and Redox Shuttles" at Monash University in Melbourne under the supervision of Prof. A.M. Bond and Prof. D.R. MacFarlane.

Nov. 94 Nov.96 - Jan. 97 Feb. 97 - July 99

Work Experience:
July 91 Research Assistant: University Budapest/Hungary, Institute for General and Analytical Chemistry. Synthesis and measurement of new plasticized PVC H+-selective electrode membranes. Environmental Technical Adviser: ISS Vienna, division for waste management. Adviser on toxic waste, identification and supervision of correct handling of toxic waste from private households. (permanent part time position) Research Assistant: Central Laboratory, Wilhelminen Hospital Vienna, Department for Flow Cytometry. Performance of routine analysis (immune status determinations and classifications of leukemia subtypes) and research on new methods for determination of leukocytes sub populations. Supervisors: Doz.Dr. W. Hbl and Prof. P.M. Bayer.

Aug. 92 - Oct. 93

Sep. 95 - Aug. 96

Curriculum Vitae: Georg Wolfbauer

page: 234

Awards:
1993 1997 1998 1998 ERASMUS Research Scholarship, performed at the University of Hull in England. Monash University Doctorate Scholarship. Monash University Travel Award. Co-winner of the National (Austria) IFCC-AVL Award for Significant Advances in Critical Care Testing 1998 for the paper entitled "Differential Expression of Tumor Necrosis Factor Receptor Subtypes on Leukocytes in Systemic Inflammatory Response Syndrome". Monash University Publication Award

1999

Publications:
-Journal Papers: 1. 2. 'On-Line Determination of Vanadium by Adsorptive Stripping Voltammetry' G.M. Greenway, G. Wolfbauer, Anal. Chim. Acta, 1995, 312(1), 15-25 'Interferenzen der automatisierten Leukozytenzhlung: Abklrung durch die Durchfluzytometrie' W. Hbl, G. Wolfbauer, P.M. Bayer, Ber. d. GKC, 1996, 19, 17-22 'Toward a new Reference Method for the Five Part Differential' W. Hbl, G. Wolfbauer, S. Andert, G. Thum, J. Streicher, C. Hbner, A. Lapin, P.M. Bayer, Cytometry, 1997, 30(2), 72-84 'Differential Expression of TNF-Receptor Subtypes in Patients with Systemic Inflammatory Response Syndrom' W. Hbl, G. Wolfbauer, J. Streicher, S. Andert, G. Stanek, S. Fitzal, P.M. Bayer, Critical Care Medicine, 1999, 27(2), 319-324 'Voltammetric Determination of the Reversible Redox Potential for the Oxidation of the Highly Surface Active Polypyridyl Ruthenium Photovoltaic Sensitizer RuII(dcbpy)2(NCS)2 (dcbpy = 2,2-bipyridine-4,4-dicarboxylic acid)' A.M. Bond, G. Deacon, J. Howitt, D. MacFarlane, L. Spiccia, G. Wolfbauer, J. Electrochem. Soc., 1999, 146(2), 648-656 'Spectroscopic and electrochemical characterisation of the photo-catalysed reduction and proton transfer reactions that occur at the interface between solid [S2Mo19O62]4-, water and triphenylphosphine.' John C. Eklund, Alan M. Bond, Keith S. Murray, Truc Vu, Anthony G. Wedd and Georg Wolfbauer, J. Chem. Soc., Dalton Trans., 1999, submitted. 'Electrochemical and spectroscopic studies on the oxidation of the cis-(Et2dcbpy)2RuX2 series of photovoltaic sensitizer precursor complexes (Et2-dcbpy =2,2'bipyridine-4,4'-diethoxydicarboxylic acid, X = Cl, I, NCS, CN).' G. Wolfbauer, A.M. Bond, D.R. MacFarlane, Inorganic Chemistry, 1999, in press. 'Electrochemical and spectroscopic studies on the reduction of the cis-(Et2dcbpy)2RuX2 series of photovoltaic sensitizer precursor complexes (Et2-dcbpy =2,2'bipyridine-4,4'-diethoxydicarboxylic acid, X = Cl, I, NCS, CN).' G. Wolfbauer, A.M. Bond, D.R. MacFarlane, J. Chem. Soc., Dalton Trans., 1999, submitted. 'Experimental and theoretical investigations of the effect of deprotonation on electronic spectra and reversible potentials of photovoltaic sensitizers: Deprotonation of cis-L2RuX2 (L = 2,2'-bipyridine-4,4'-dicarboxylic acid, X = CN, NCS) by electrochemical reduction at platinum electrodes.' Georg Wolfbauer, Alan M. Bond, Glen B. Deacon, Douglas R. MacFarlane, Leone Spiccia, J. Am. Chem. Soc., 1999, submitted 'Dependence of the voltammetric oxidation of the highly efficient photovoltaic sensitizer [(H3-tctpy)RuII(NCS)3] (H3-tctpy = 2,2':6',2''-terpyridine-4,4',4''-tricarboxylic acid) on the electrode material, solvent and isomeric purity.' Georg Wolfbauer, Alan M. Bond, Glen B. Deacon, Julia Howitt, Douglas R. MacFarlane, Leone Spiccia, J. Electrochem. Soc., 1999, to be submitted

3.

4.

5.

6.

7.

8.

9.

10.

Curriculum Vitae: Georg Wolfbauer

page: 235

-Conference Papers and Posters: 1. 'On-Line Determination of Vanadium by Adsorptive Cathodic Stripping Voltammetry' G.M.Greenway, G.Wolfbauer, presented at the 6th international Conference on Flow Analysis, 1994, Toledo, Spain 'Verifying interferences of impedance-based leukocyte counting with flow cytomertry using fluorescence-labelled monoclonal antibodies.' W. Hbl, G. Wolfbauer, P.M. Bayer, Proceedings of the XVI. International Congress of Clinical Chemistry. 1996, 353-354 'Verification of impedance-based leukocyte counts using flow cytometry with fluorescence-labelled antibodies' G. Woflbauer, W. Hbl, P.M. Bayer, J. Lab. Med., 1996, 20, 727 'Toward a reference method for the five-part differential' G. Wolfbauer, W. Hbl, S. Andert, G. Thum, C. Hbner, J.Streicher, A. Lapin, P.M. Bayer, J. Lab. Med., 1996, 20, 728 'Determination of Reversible Redox Potentials of Highly Surface Active Polypyridine Ruthenium Complexes' G. Wolfbauer, A.M. Bond, J.A. Howitt, D.R. MacFarlane, Abstracts of RACI Inorganic Chemistry Division Conference, 1998, Wollongong, Australia, poster PT85 A Channel Flow Cell System Specifically Designed to Study the Efficiency of Redox Shuttles in Dye Sensitized Solar Cells Alan M. Bond, John C. Eklund, Doug MacFarlane, Georg Wolfbauer, Conference paper for the Twelfth International Conference on Photochemical Conversion and Storage of Solar Energy, 1998, Berlin, Germany, poster 1W36

2.

3.

4.

5.

6.

Вам также может понравиться