fusheng li 李福胜 -...

104
Akademisk avhandling som med tillstånd av Kungl Tekniska Högskolan i Stockholm framlägges till offentlig granskning för avläggande av doktorsexamen i kemi fredagen den 26 februari kl 10.00 i sal F3, KTH, Lindstedtsvägen 26, Stockholm. Avhandlingen försvaras på engelska. Opponent ä r Prof. Franc Meyer, Georg-August-Universitä t Göttingen, Germany. Design of Water Splitting Devices via Molecular Engineering Fusheng Li (李福胜) Doctoral Thesis Stockholm, 2016

Upload: domien

Post on 02-Apr-2018

215 views

Category:

Documents


1 download

TRANSCRIPT

Akademisk avhandling som med tillstånd av Kungl Tekniska Högskolan i Stockholm

framlägges till offentlig granskning för avläggande av doktorsexamen i kemi fredagen

den 26 februari kl 10.00 i sal F3, KTH, Lindstedtsvägen 26, Stockholm. Avhandlingen

försvaras på engelska. Opponent är Prof. Franc Meyer, Georg-August-Universität

Göttingen, Germany.

Design of Water Splitting Devices via

Molecular Engineering

Fusheng Li

(李福胜)

Doctoral Thesis

Stockholm, 2016

ISBN 978-91-7595-841-5

ISSN 1654-1081

TRITA-CHE Report 2016:6

© Fusheng Li, 2016

Universitetsservice US AB, Stockholm

To the people who ever helped me!

Fusheng Li, 2016: “Design of Water Splitting Devices via Molecular

Engineering”, School of Chemical Science and Engineering, KTH Royal

Institute of Technology, SE-100 44 Stockholm, Sweden.

Abstract

Converting solar energy to fuels such as hydrogen by the reaction of water

splitting is a promising solution for the future sustainable energy systems. The

theme of this thesis is to design water splitting devices via molecular

engineering; it concerns the studies of both electrochemical-driven and photo-

electrochemical driven molecular functional devices for water splitting.

The first chapter presents a general introduction about Solar Fuel Conversion.

It concerns molecular water splitting catalysts, light harvesting materials and

fabrication methods of water splitting devices.

The second chapter describes an electrode by immobilizing a molecular water

oxidation catalyst on carbon nanotubes through the hydrophobic interaction.

This fabrication method is corresponding to the question: “How to employ

catalysts in functional devices without affecting their performances?”

In the third chapter, molecular water oxidation catalysts were successfully

immobilized on glassy carbon electrode surface via electrochemical

polymerization method. The OO bond formation pathways of catalysts on

electrode surfaces were studied. This kinetic study is corresponding to the

question: “How to get kinetic information of RDS when a catalyst is

immobilized on the electrode surface?”

Chapter four explores molecular water oxidation catalysts immobilized on dye-

sensitized TiO2 electrode and Fe2O3 semiconductor electrode via different

fabrication methods. The reasons of photocurrent decay are discussed and two

potential solutions are provided. These studies are corresponding to the

question: “How to improve the stability of photo-electrodes?”

Finally, in the last chapter, two novel Pt-free Z-schemed molecular photo-

electrochemical cells with both photoactive cathode and photoactive anode for

visible light driven water splitting driven were demonstrated. These studies are

corresponding to the question: “How to utilize the concept of Z-scheme in

photosynthesis to fabricate Pt-free molecular based PEC cells?”

Keywords: electrochemical-driven water splitting, artificial photosynthesis,

molecular catalysts, photo-electrochemical cell, photo-electrodes,

electrochemistry, immobilization.

Sammanfattning på svenska

Omvandling av solens energi till bränsle, såsom väte genom sönderdelning av vatten,

utgör ett lovande alternativ för framtida hållbara energisystem. Huvudsyftet med denna

doktorsavhandling är att utforma elektrokemiska celler för sönderdelning av vatten, där

det huvudsakliga verktyget är molekylär design; både avseende funktionella

elektrokemiska och fotoelektrokemiska celler.

Det första kapitlet ger en allmän introduktion till forskningsområdet för omvandling till

solbränsle. Kapitlets innehåll fokuserar främst på molekylära katalysatorer för

sönderdelning av vatten, ljusabsorberande material och utformningen av

elektrokemiska och fotoelektrokemiska celler.

Det andra kapitlet beskriver hur en elektrod för vattenoxidation kan

konstrueras genom immobilisering av en molekylär katalysator på ytan av

kolnanotuber genom utnyttjande av deras hydrofoba interaktion.

Tillverkningsmetoden adresserar frågan: ”Hur kan katalysatorer användas i

funktionella enheter utan att deras funktion påverkas?”

I det tredje kapitlet beskrivs hur en molekylär katalysator för sönderdelning av vatten

kan immobiliseras på en elektrodyta av glasaktigt kol med hjälp av en elektrokemisk

polymerisationsmetod. Mekanismen för katalysen av O-O-bindningar vid elektrodytan

studerades i detalj. Denna kinetiska studie berör frågan: ”Hur kan man erhålla

kinetisk information om det hastighetsbestämmande steget i en katalytisk process

då katalysatorn är bunden till en elektrodyta?”

Kapitel fyra ägnas åt studier av katalysatorer för sönderdelning av vatten, vilka genom

olika tillverkningstekniker immobiliserats på färgämnessensiterade elektroder eller

halvledande elektroder bestående av Fe2O3. Orsaken till den observerade minskningen

av fotoströmmen diskuteras och två möjliga orsaker identifieras. Dessa delstudier är

relaterade till frågan: ”Hur kan stabiliteten av fotoelektroder förbättras?”

Slutligen, i det sista kapitlet beskrivs två nya molekylära fotoelektrokemiska

celler av tandemtyp utan användning av platina. I dessa celler används både en

fotoaktiv katod och fotoaktiv anod för att driva sönderdelningen av vatten med

hjälp av synligt ljus. Dessa celler hänför sig till frågan: ”Hur kan Z-konceptet

användas för att tillverka molekylära fotoelektrokemiska celler utan

användning av platina?”

Nyckelord: elektrokemisk sönderdelning av vatten, artificiell fotosyntes,

molekylära katalysatorer, fotoelektrokemisk cell, fotoelektroder, elektrokemi,

immobilisering.

Abbreviations

APT atomic proton transfer

bpy 2,2-bipyridine

CB conduction band

CeIV cerium (IV) ammonium nitrate, Ce(NH4)2(NO3)6

CV cyclic voltammetry

DFT density functional theory

DMSO dimethyl sulfoxide

DPV differential pulse voltammetry

E0 standard potential of a reaction, redox couple and electrode

E1/2 half-wave potential in voltammetry

GC glassy carbon

H2BDA 2,2-bipyridine-6,6-dicarboxylic acid

HGC hydrogen generation catalyst

H2PDA 1,10-phenanthroline-2,9-dicarboxylic acid

H2PDC 2,6-pryidinedicarboxylic acid

IPCE incident photon-to-electron conversion efficiency

I2M interaction of two M−O units

LSV linear scan voltammetry

MS mass spectrometry

NHE normal hydrogen electrode

NMR nuclear magnetic resonance

OEC oxygen-evolving-complex

PCET proton coupled electron transfer

PDC 2,6-pyridinedicarboxylic acid

pic 4-picoline

PS I photosystem I

PS II photosystem II

PV photovoltaic

py pyridine

tpy 2,2 - 6, 2-terpyridine

TOF turnover frequency

TON turnover number

RDS rate determining step

RC radical coupling

WNA water nucleophilic attack

WOC water oxidation catalyst

VB valence band

ƞ over potential of catalytic reaction

2-E two-electrode

3-E three-electrode

List of Publications

This thesis is based on the following papers, referred to in the text by their

Roman numerals I-VI:

I. Immobilization of a molecular catalyst on carbon nanotubes for

highly efficient electro-catalytic water oxidation

Fusheng Li, Lin Li, Lianpeng Tong, Quentin Daniel, Mats Göthelid, and

Licheng Sun

Chem. Commun. 2014, 50, 13948-13951.

II. Control the OO bond formation pathways by immobilizing

molecular catalysts on glassy carbon via electrochemical

polymerization

Fusheng Li, Lele Duan, Ke Fan, Lei Wang, Quentin Daniel, Hong Chen,

and Licheng Sun

Manuscript

III. Immobilizing Ru(BDA) catalyst on a photoanode via

electrochemical polymerization for light-driven water splitting

Fusheng Li, Ke Fan, Lei Wang, Quentin Daniel, Lele Duan, and

Licheng Sun

ACS Catal. 2015, 5, 3786−3790.

IV. Immobilization of a molecular ruthenium catalyst on hematite

nanorod arrays for water oxidation with stable photocurrent

Ke Fan,† Fusheng Li,

† Lei Wang, Quentin Daniel, Hong Chen, Erik

Gabrielsson, Junliang Sun, and Licheng Sun († authors contributed

equally to this work)

ChemSusChem 2015, 8, 3242-3247.

V. Pt-free tandem molecular photoelectrochemical cells for water

splitting driven by visible light

Ke Fan,† Fusheng Li,

† Lei Wang, Quentin Daniel, Erik Gabrielsson, and

Licheng Sun († authors contributed equally to this work)

Phys. Chem. Chem. Phys. 2014, 16, 25234-25240.

VI. Dye-Sensitized tandem photoelectrochemical cell for light driven

total water splitting

Fusheng Li, Ke Fan, Bo Xu, Erik Gabrielsson, Quentin Daniel, Lin Li,

and Licheng Sun

J. Am. Chem. Soc. 2015, 137, 9153-9159.

Papers not included in this thesis:

VII. Highly Efficient Bioinspired Molecular Ru Water Oxidation

Catalysts with Negatively Charged Backbone Ligands

Lele Duan, Lei Wang, Fusheng Li, Fei Li, and Licheng Sun

Acc. Chem. Res. 2015, 48, 2084−2096

VIII. Sensitizer-Catalyst Assemblies for Water Oxidation

Lei Wang, Mohammad Mirmohades, Allison Brown, Lele Duan,

Fusheng Li, Quentin Daniel, Reiner Lomoth, Licheng Sun and Leif

Hammarström

Inorg. Chem. 2015, 54, 2742-2751.

IX. Electrochemical Driven Water Oxidation by Molecular Catalysts in

situ Polymerized on the Surface of Graphite Carbon Electrode

Lei Wang, Ke Fan, Quentin Daniel, Lele Duan, Fusheng Li, Bertrand

Philippe, Håkan Rensmo, Hong Chen, Junliang Sun, and Licheng Sun

Chem. Commun. 2015, 51, 7883-7886.

X. Dipicolinic acid: a strong anchoring group with tunable redox and

spectral behavior for stable dye-sensitized solar

Erik Gabrielsson, Haining Tian, Susanna K Eriksson, Jiajia Gao, Hong

Chen, Fusheng Li, Johan Oscarsson, Junliang Sun, Håkan Rensmo, Lars

Kloo, Anders Hagfeldt, and Licheng Sun

Chem. Commun. 2015, 51, 3858-3861.

XI. Visible light-driven water oxidation using a covalently-linked

molecular catalyst–sensitizer dyad assembled on a TiO2 electrode

Masanori Yamamoto, Lei Wang, Fusheng Li, Takashi Fukushima, Koji

Tanaka, Licheng Sun, and Hiroshi Imahori

Chem. Sci., 2016, DOI: 10.1039/C5SC03669K

XII. Tailored design for RuBDA water oxidation catalyst

Quentin Daniel, Lei Wang, Lele Duan, Fusheng Li, and Licheng Sun

Manuscript.

XIII. Rearranging from 6- to 7- coordination initiates the catalytic

activity: an EPR study on a Ru-BDA water oxidation catalyst

Quentin Daniel, Ping Huang, Ting Fan, Lele Duan, Lei Wang, Fusheng

Li, Mårten Ahlquist, Fikret Mamedov, Stenbjörn Styring, Licheng Sun

Manuscript.

XIV. 3D-Structured Nickel-Vanadium Monolayer Double Hydroxide for

Efficient Electrochemical Water Oxidation

Ke Fan, Hong Chen, Yongfei Ji, Hui Huang, Per M. Claesson, Quentin

Daniel, Philippe Bertrand, Håkan Rensmo, Fusheng Li, Yi Luo, and

Licheng Sun

Manuscript.

5

Table of Contents Abstract Abbreviations List of publications

1. Introduction ..................................................................................... 1 1.1. Solar Fuels ‒ two different approaches ................................................ 2 1.2. Indirect approach: PV-electrolyzer ........................................................ 3 1.3. Direct approach: Photosynthesis .......................................................... 4

1.3.1. Natural Photosynthesis ............................................................................ 4 1.3.2. Artificial Photosynthesis - integrated assembly systems ........................... 5 1.3.3. Artificial Photosynthesis - photoelectrochemical cells ............................... 6

1.4. Catalysts for Solar Fuel Conversion ..................................................... 8 1.4.1. Water Oxidation Catalysts ........................................................................ 8 1.4.2. Hydrogen Generation Catalysts ..............................................................11

1.5. Light Harvesting Materials for Artificial Photosynthesis ...................... 11 1.5.1. Semiconductors ......................................................................................12 1.5.2. Molecular Photosensitizers .....................................................................13

1.6. Functional electrodes for water splitting .............................................. 14 1.6.1. Electrochemical-driven water splitting .....................................................15 1.6.2. Light-driven water splitting ......................................................................18

1.7. Challenges for Solar to Fuel Conversion ............................................ 23 1.8. The Aim of This Thesis ....................................................................... 24

2. Immobilization of a Molecular Catalyst on Carbon Nanotubes for

Highly Efficient Electrochemical-driven Water Oxidation ........................ 25 2.1. Introduction ......................................................................................... 27 2.2. Synthesis ............................................................................................ 27 2.3. Results and Discussions ..................................................................... 28

2.3.1. Electrochemical properties .......................................................................28 2.3.2. Preparation of Electrode ..........................................................................29 2.3.3. Electro-driven Water Oxidation ................................................................30

2.5. Summary ............................................................................................ 33 3. Control the O‒O bond formation pathway by immobilizing

molecular water oxidation catalysts on the electrode surface ................. 35 3.1. Introduction ......................................................................................... 37 3.2. Design, Synthesis and Preparation of Electrodes ............................... 38 3.3. Results and Discussions ..................................................................... 39

3.3.1. CH3CN ligand exchange ..........................................................................39 3.3.2. pH Dependent Electrochemistry ...............................................................40 3.3.3. Methods for Electrode Kinetics.................................................................41 3.3.4. The O‒O Bond Formation Pathways on Electrodes .................................43

3.4. Summary ............................................................................................ 49 4. Immobilizing Ru catalysts on photoanodes for light-driven water

splitting..................................................................................................... 51

4.1. Introduction ......................................................................................... 53 4.2. Synthesis and Preparation of Electrodes ............................................ 54 4.3. Results and Discussions ..................................................................... 55

4.3.1. Electrochemical Properties ......................................................................55 4.3.2. Photo-Electrochemistry ............................................................................56 4.3.3. Kinetics Study ..........................................................................................57 4.3.4. Replacement of the Molecular Photosensitizer by α-Fe2O3 ......................59

4.4. PDC as the Anchoring Group for Photoanode .................................... 60 4.5. Summary ............................................................................................ 62

5. Dye-sensitized tandem PEC cells for light driven total water

splitting by employing the concept of Z-scheme ..................................... 65 5.1. Introduction ......................................................................................... 67 5.2. Rational Design of Tandem Dye-sensitized PEC cells ....................... 68 5.3. Synthesis and Preparation of Electrodes ............................................ 68 5.3. Results and Discussions ..................................................................... 69

5.3.1. Electrochemical Properties ......................................................................69 5.3.2. Photo-Electrochemistry ............................................................................70

5.4. Redesign and Replacement of RuP.................................................... 72 5.5. Summary ............................................................................................ 79

6. Concluding Remarks .................................................................... 81 7. Future prospects ........................................................................... 83 Acknowledgements ................................................................................. 84 Appendix A .............................................................................................. 85 References .............................................................................................. 86

1

1. Introduction

Energy plays a crucial role in the development of modern human society.

Nowadays, our energy systems are mainly produced from fossil fuels such as

petroleum, natural gas and coal.1 Unfortunately, global energy demand and

consumption are rapidly increasing with both the expansion of population and

the improvement of living standard, leading to the extensive use of fossil fuels

and consequently giving rise to problems such as environmental pollution,

global warming crisis and sulfur dioxide emissions etc.. As a result of

continuous combustion of carbon-based fuels, the unacceptable changes of the

Earth‟s climate start to occur. All these issues point to the requirement of new

sustainable, carbon-neutral energy systems.2

With the advantage of renewable and carbon-neutral properties, hydrogen is

considered as a green energy carrier with a high energy density by weight.

Currently, hydrogen is mainly produced from fossil fuels by steam reforming,

partial oxidation of methane and coal gasification, which are of course not

sustainable and carbon-neutral approach. Producing hydrogen by splitting

water with renewable electricity or artificial photosynthesis is considered to be

an ideal way for developing sustainable energy systems.3,4

There are two half reactions consisted in the whole water splitting process: (I)

water oxidation, where two water molecules are oxidized to one molecular

oxygen and four protons are produced (eqn. 1); (II) proton reduction, where

protons are reduced to hydrogen (eqn. 2).

Sunlight is clean and highly abundant energy source on the Earth. Utilization

of solar energy to produce energy-rich hydrogen by the splitting of water (eqn.

3) is one of the most promising approaches.5-8

This carbon-neutral chemical

reaction just requires the free energy source (sunlight) and nearly unlimited

starting materials (H2O) and produces only one byproduct (O2).9 To reduce the

energy barrier and accelerate water splitting reactions, great efforts in this field

have been devoted to develop highly efficient water oxidation catalysts (WOCs)

and hydrogen generation catalysts (HGCs).10,11

However, for the future

application, several other challenges still need to be addressed. One of the

most challenging tasks is: how to employ these catalysts to build functional

devices? Fabrication of molecular-precise architectures confined at interfaces

through molecular engineering can provide efficient and versatile tools for

building functional electrodes.

2

1.1. Solar Fuels ‒ two different approaches

Converting solar energy to hydrogen and oxygen through water splitting

reaction has been considered as a promising idea for a long time. Jules Verne

wrote that: “water will be the coal of the future”in his fiction “The Mysterious

Island” in 1874. Since sunlight is the most abundant energy source on the

Earth, and there are two approaches that can be used to convert solar energy to

fuels such as hydrogen (as shown in Figure 1). One is indirect approach, in

which solar electricity is produced by photovoltaic (PV) devices and then used

to produce hydrogen by electrolysis of water. The other one is an direct

approach, where solar energy can be used to produce hydrogen by splitting

water.1

Figure 1. Renewable fuels generation ‒ two different approaches.

The photovoltaic cell is one of the most popular solar technologies and it

clearly has great potential for serving carbon-free energy. However, electricity

cannot drive everything! According to the electricity statistics of International

Energy Agency (IEA), only 22% of the total energy production will be used as

electricity in 2030, which means we still need a huge amount of energy in the

form of fuels. On the other hand, solar cell technologies can only convert

sunlight into electricity which is hard to be stored in big capacity for

transportation and nighttime use.

Instead of direct utilization, the electricity can be used in device such as

electrolyzer to produce hydrogen from water, which is an energy storage

process. This indirect approach can be easily designed by employing

appropriate catalysts, electrodes and electrolytes in the electrolyzer, and solar

energy is converted to electricity by solar cells to drive the water splitting

reaction.

The second approach is more straightforward. Similar to the natural

photosynthesis system, solar energy is directly collected and used to produce

hydrogen by the reaction of water splitting, without intermediate energy

carriers (such as solar electricity). This approach is called “Artificial

3

Photosynthesis”, which is an example of a direct pathway for storage of solar

energy.8

Both of the direct and indirect approaches are chemical processes for storing

energy from sunlight into chemical bonds of fuels (solar fuels). In this thesis,

two topics of electrochemical-driven water splitting (electrolyzer) and light-

driven water splitting (artificial photosynthesis) will be discussed.

1.2. Indirect approach: PV-electrolyzer

For the indirect approach, a PV-electrolyzer device is used to carry out solar-

to-fuels conversion (as shown in Figure 2). Efficient PVs and electrolyzers

with high faradaic efficiency are required.12,13

The solar-to-fuels conversion

efficiency (SFE) for a PV-electrolyzer device can be calculated by eqn.4,13-16

where Jop is the point of intersection and gives the operation current density, V

is the water splitting potential required (1.23 V), Plight is incident light power

and ηF is faradaic efficiency.

0.0 0.5 1.0 1.5 2.0S

FE

%

Cu

rre

nt

de

nsity

Voltage (V) Figure 2. (left)Schematic diagrams of the PV-electrolyzer device. (right) the

relationship between solar to fuel efficiency (SFE), the performances of PV and

electrolyzer, red lines are the J-V curves of PV, solid line means PV with higher

efficiency. Blue lines are the linear sweep voltammetry ( LSV) curves of electrolyzer.

Solid line means electrolyzer with higher activity and lower over potential.

From the crossing points of PV curves and LSVs, we can find that by

combining a high efficiency PV with a high active electrolyzer, a high Jop can

be obtained, correspondingly, resulting in a highly solar-to-fuels conversion

efficiency. Much effort has been devoted to the development of high efficiency

PVs. Unfortunately, the cost per watt of solar photovoltaic electricity is still

presently too high for a practical water splitting device.

2

water splitting

-2

( [ ] ( 1.23V) ). 4

[mWcm ]

op F

light

J mAcm VSFE eqn

P

4

On the other side, due to the presence of over-potentials for both half reactions

(oxygen-evolution reaction-OER and hydrogen-evolution reaction-HER) and

the resistance losses in the electrolyzer (such as contact resistance, solution

resistance and the resistance of electrodes), the actual voltage (Vactual) required

is always higher than the ideal thermodynamic potential of 1.23V. The Vactual

can be described by eqn. 5.

1.23V .5OER HERactualV iR eqn

For the principle of PV-electrolyzer devices design, electrodes (cathode and

anode) with both low over-potentials and high activities for water splitting are

urgently needed to minimize the value of Vactual, which highly depends on the

properties of catalysts and electrode fabrication methods.

1.3. Direct approach: Photosynthesis

1.3.1. Natural Photosynthesis

In nature, solar energy can be collected and stored as carbohydrates and other

biomass. However, natural photosynthesis is not very efficient (yields of

biomass efficiencies is usually less than 1%).17

But natural photosynthesis

system provides an ideal blue print of solar energy conversion.18

In nature,

water oxidation is catalyzed by an oxygen-evolving-complex (OEC, a

Mn4CaO5 cluster) in the enzyme of photosystem II (PS II) driven by sunlight.

Through an electron transportation chain,i the electrons extracted from water

are delivered to photosystem I (PS I) where they are pumped to a higher

energy level for the conversion of NADP+ to NADPH. At the same time,

energy also can be used to convert ADP to ATP during the electron transfer

process.19

Figure 3. Simplified Z-scheme of light-induced reaction for natural photosynthesis.

i Electron transport chain in Z-scheme. ET = electron transport. i) OEC→Tyr→P680. ii)

P680*→Pheo→Qa→Qb→Cyt b6f→PC→P700. iii) P700*→A0→A1→Fx/a/b→FD→FNR. Tyr =

tyrosine; Pheo= pheophytin; Qa and Qb =Quinones A and B; Cyt b6f = an Fe-S Rieske center complex; PC = plastoquinone binding sites; A0 = chlorophy II; A1 = phylloquinone; Fx/a/b = three

separated Fe-S protein; FD = ferredoxin; FNR = ferredoxin-nicotinamide-adenine dinucleotide

phosphate reductase; P680 and P700 = pigments with absorption maximum at 680nm or 700nm.

5

In this whole process, known as the Z-scheme (as shown in Figure 3), P680 in

PS II and P700 in PS I are the light-harvesting antennas, which create light-

induced charge separations when they are excited (collecting solar energy).

Meanwhile, water serves as electron and proton donor to produce the energy-

carrying agents (ATP and NADPH) for the “dark” reaction where CO2 is

converted to carbohydrates (storing solar energy).

Collecting and storing solar energy into chemical bonds, as nature

accomplishes through the fantastic photosynthesis, is a highly desirable

approach to solve the energy challenges.6 This Z-scheme provides scientists a

general direction for how to create artificial photosynthesis systems with man-

made materials.20

1.3.2. Artificial Photosynthesis - integrated assembly systems

There is no doubt that artificially duplicating any functional components of

natural photosynthesis is a tremendous challenge. For artificial photosynthesis

system, two different strategies are currently accepted, one is an integrated

assembly (IA) system, and the other one is a photo-electrochemical cell (PEC

cell). Usually the IA systems are molecular or particle devices (as shown in

Figure 4), in which electron-hole pairs are generated by the light-induced

charge-separation from chromophores (dye molecules or light-absorbing

semiconductors), then electrons are transferred to the HGC for hydrogen

generation. Simultaneously, holes are transferred to the WOC for the oxidation

of water.

Figure 4. Schematic diagrams of integrated assembly systems, a) molecular dye system.

b) Single-step semiconductor particle with attached HGC and WOC as co-catalysts. c)

Mixture of semiconductor particles with attached HGC or WOC as co-catalysts and

redox electrolyte couples.

As shown in Figure 5, several principal requirements must be met for the

design of the integrated assembly systems. Thermodynamically, the LUMO

level of chromophores must be more negative than the onset potential of HGC,

and the HOMO level of chromophores must be more positive than the onset

potential of WOCs. Kinetically, the electron transfer between the excited state

of chromophores and catalysts must be fast enough to avoid the quenching of

6

chromophores* by other processes (charge recombinations etc.). As these

integrated assembly systems are homogeneous or semi-homogeneous reactions,

the hydrogen and oxygen gases generated during the water splitting process are

present as a mixture in the same reaction chamber.7

Figure 5. Schematic diagram of the simplified electron flows and related potentials for

integrated assembly water splitting systems.

1.3.3. Artificial Photosynthesis - photoelectrochemical cells

The so-called PEC cell is based on the integration of functional materials with

electrodes, in which the water oxidation and hydrogen generation sites are

separated on two different electrodes.

Figure 6. Schematic diagrams, PEC cells with one photo-active electrode. Left, a

photoanode works as photo-active center. Right, a photocathode works as photo-active

center

7

The working principles of PEC cells are shown in Figure 6. When a

photoanode combined with a counter electrode (Figure 6 left), the electrons

are extracted from water by light-induced charge-separation generated from

the chromophore, and pumped to the counter electrode where the hydrogen

generation reaction occurs. In this case, the photoanode is the photo-active

center. A photocathode can also be the photo-active center with the similar

device structure, as shown in Figure 6 right. Due to the thermodynamic

requirements, for example, when a photoanode combined with Pt (as HGC),

the LUMO level (or the conduction band corresponding to the semiconductor)

of chromophore must be more negative than the onset potential of Pt for

hydrogen generation, and the HOMO level (or valence band corresponding to

the semiconductor) must be more positive than the onset potential of WOC for

water oxidation. Nonetheless, chromophores usually cannot meet these two

requirements at the same time. As a result, the properties of photoanodes or

photocathodes are more extensively studied instead of whole PEC devices, in

which three-electrode (3-E) setups are employed, bias are usually needed to

balance the charge transfer between working and counter electrodes, and

counter electrodes obtain extra energy from potentiostat.

Figure 7. Schematic diagram, simplified electron flows of a Z-schemed PEC cell with

two photo-active electrodes.

Another type of PEC devices is a Z-schemed PEC cell. Figure 7 shows the

simplified electron flows via the external conductive wire from photoanode,

where the water oxidation reaction occurs, to the photocathode, where the

hydrogen generation reaction happens. In principle, the tandem PEC cell has to

meet the following criteria to work properly: the HOMO level of

chromophore 1 is more positive than the onset potential of the WOC; the

8

LUMO level of chromophore 2 is more negative than the onset potential of

the HGC; and the LUMO level of chromophore 1 is more negative than the

HOMO level of chromophore 2. Two chromophores with different spectra

can expand the solar light absorption of PEC cells. Comparing to single photo-

active electrode PEC devices, Z-schemed PEC cells have two light harvesting

units: one handles the water oxidation reaction and the other manages the

hydrogen generation reaction. In theory, extra bias potentials can be avoided.

Meanwhile, both electricity and solar fuels can be generated from the Z-

schemed PEC cell at the same time. Solar energy conversion by a Z-schemed

PEC cell is very similar to the natural photosynthesis. They all share the same

functional components (water as the electron/proton donor, two light

harvesting centers as well as two charge pumping processes) and similar

energy storage processes (hydrogen corresponding to NADPH, electricity

corresponding to ATP).

To date, due to the difficulty in matching the energy levels of both photo-

active electrodes, molecular Z-schemed PEC cells are rarely reported. In this

thesis, our efforts on the design of molecular Z-schemed PEC cells will be

demonstrated.

1.4. Catalysts for Solar Fuel Conversion

No matter which device configuration is employed for solar fuel generation,

the kinetics of the water splitting reaction is a key element. Therefore to

accelerate the water splitting reaction, incorporating the electrode surface with

efficient WOCs and HGCs is highly required.

1.4.1. Water Oxidation Catalysts

Water oxidation requires simultaneous transfer of four protons and four

electrons from two water molecules and therefore is the bottleneck of the full

water splitting process.10,11

Water oxidation catalysts can be divided into inorganic materials and

molecular catalysts. For inorganic materials, typical WOCs include metal

oxides, such as Ir-, Ru-, Fe-, Co-, Ni-, and Mn-based transition-metal oxides.21-

26 In this thesis, we focus on the molecular catalysts; metal oxide based WOCs

are beyond the scope of this thesis and will not be discussed.

Molecular WOCs, especially, Ruthenium based complexes have drawn much

attention because they can be easily modified structurally and electronically,

and their ability of mediating multiple-electron transfer reactions can provide

opportunities for comprehensive mechanistic studies.20,27-41

9

A few well-known Ru-based WOCs are representatively shown in Figure 8. In

1982 the first Ru-based molecular WOC so called “blue dimer” 1 was reported

by Meyer et al. with a turnover number (TON) of 13 and turnover frequency

(TOF) of 0.004 s-1

.36

From then on, bimolecular Ru WOCs have been rapidly

developed through tailored ligand design, Thummel‟s complex 2 exhibits

larger TON and TOF toward CeIV

-driven water oxidation.39

Recently Franc

Meyer et al. reported a novel binuclear Ru-WOC 3 based on a specific ligand

tailoring design and the performance of this catalyst has been systematically

improved. The study on complex 3 also demonstrated how subtle ligand

modifications cause a change in the O−O bond formation mechanism, which is

very instructive.34

Figure 8. Structures of bimolecular and monomolecular Ru WOCs and their catalytic

performances are given in the form of (TON, TOF) based on Ce(NH4)2(NO3)6 as the

oxidant at pH = 1, the TOF of Mn4CaO5 cluster (OEC of PS II) is also given.ii

ii TON and TOF are two major parameters to exam the catalytic performances of catalysts. Water

soluble single-electron chemical oxidant [Ce(NH4)2(NO3)6] (CeIV) is usually used to evaluate

catalytic activity of molecular WOCs in aqueous medium. CeIV has a very high oxidation potential

around 1.7 V at pH 0. The corresponding electron-accepting half reaction and total reaction are described as follows:

10

In 2005, Thummel et al. published the first monomolecular Ru WOC 5 with a

TON of 260 and TOF of 0.014 s-1

.39

Since then, the catalytic performances of

monomolecular Ru WOCs have been significantly improved. Among them, the

star molecules [Ru(BDA)L2] (BDA2

= 2,2-bipyridine-6,6-dicarboxylate; L =

N-cyclic aromatic ligands) can achieve the highest TOF of 1000 s-1

(8) and the

highest TON of 100 000 (9).42

It takes only 32 years that the artificial WOCs

have already obtained a much better activity than the oxygen-evolving-

complex from Photosystem II!

Studies on Ru-based WOCs, as model complexes, can not only provide the

information about the structure-performance correlations for the design of non-

noble metal WOCs, but also give chemists a deeper understanding of the

mechanism for the O-O bond formation.

The O-O bond formation is the most important step for water oxidation, there

are two widely accepted pathways for Ru WOCs as described in eqn. 6 and

eqn .7.31

One is water nucleophilic attack (WNA) pathway (eqn. 6), in which a

water molecule nucleophilically attacks the oxo group of RuV=O, resulting in a

two-electron reduction of RuV=O and the O-O bond formation. Most Ru-based

WOCs catalyze water oxidation through the WNA pathway.

Another pathway is the radical coupling (RC) pathway as described in eqn. 7,

in which two mono-radical RuIV

-O∙ units couple with each other, forming a

peroxo intermediate. [Ru(BDA)L2] catalyzes water oxidation through the RC

pathway.

Recently, several molecular WOCs based on first-row transition-metal such as

Mn, Fe, Co, Ni and Cu have been developed.43-53

Although the catalytic

mechanism and structure-activity correlation of these WOCs are still not

known as much as that of Ru-based WOCs, they are very attractive for opening

up the way to develop WOCs based on earth-abundant elements.

In this thesis, complexes 6 (WNA catalyst) and 7 (RC catalyst) are selected as

model complexes for electrodes modification via different fabrication methods.

With the rapid development of WOCs, we hope our approaches and methods

will inspire the research on non-noble metal WOCs-based water splitting

devices in the future.

11

1.4.2. Hydrogen Generation Catalysts

Nature also provides exquisite HGCs in the form of hydrogenases, which are

based on earth-abundant metals like iron and interconvert protons to hydrogen

with high TOFs up to 100-10000 per second.54-56

Artificial or man-made

hydrogen generation catalysts can also be divided into inorganic materials and

molecular catalysts. Platinum and other precious metals are considered as

highly-active and robust catalysts, but suffering from high cost and low

reserves.57-60

Recently, materials based on earth-abundant elements, such as

metal alloys,61

and metal sulfide,62-64

have been developed. Molecular

approaches for hydrogen production also have been well developed, and these

important topics have been the subjects of several recent reviews.43,60,65,66

Figure 9. Structures of molecular hydrogen generation catalysts based on Fe, Co and

Ni as metal center.

Figure 9 shows some well-studied HGCs based on Fe, Co and Ni. The

majority of Fe-based HGCs, such as complex 10, are targeted to mimic the

function of nature Fe–Fe hydrogenase, and most of them operate in organic

media.54,67

Complexes 11 and 12 can be operated in organic / aqueous mixed

media.68,69

In the family of HGCs, complexes 13 and 14 are two important

members which have been developed during the past decade by chemists for

chemical, electrochemical and photochemical proton reduction,70-74

since

cobalt complexes with diglyoxime ligands (cobaloximes) as catalysts for water

reduction were first reported in 1983.75

The most important part for

cobaloximes is that they can be operated in aqueous media with low over

potentials. Due to this important property, complexes 13 and 14 have been

selected as model complexes for the water reduction studies in this thesis.

1.5. Light Harvesting Materials for Artificial Photosynthesis

Light harvesting material is one of the most challenging obstacles for solar

energy conversion. As introduced above, light harvesting materials are

required for both indirect (PV-Electrolyzer) and direct (Artificial

Photosynthesis) approaches of solar fuel generation. Photovoltaic is beyond

the scope of the thesis and will not be discussed. Man-made light harvesting

12

materials for artificial photosynthesis especially for PEC cells can be divided

into two types: (1) semiconductors; (2) molecular photosensitizers.

1.5.1. Semiconductors

Semiconductors can be used as photo-active electrode materials, because they

can produce charge separation by absorbing solar light. Oxygen evolution was

observed as early as 1968 by illuminating a TiO2 photo-electrode,76

and then

Fujishima and Honda first pointed out the concept of using an n-type

semiconductor as part of the PEC cell.77

For the utilization of semiconductors

in PEC devices several principal requirements should be met:

1. The absorption of the semiconductor should overlap with the solar

spectrum.

2. Suitable energy band positions to promote the water splitting reaction.

3. Efficient electron-transport and hole-transport properties.

4. Large specific surface area for loading catalysts.

5. Good stability.

To achieve overall water splitting, the bottom of the CB of the semiconductors

must be located at a more negative potential than the proton reduction potential

(0 V vs. NHE at pH 0). Furthermore, to facilitate water oxidation, the VB of

the semiconductors must exceed the oxidation potential of water (+1.23 V vs.

NHE at pH 0). Thus, a theoretical 1.23 eV is required to drive the overall

water-splitting reaction. This energy is equivalent to the energy of a photon

with a wavelength around 1010 nm,iii and hence ca. 70% of all solar photons

are theoretically available for water splitting as shown in table 1.78

Table 1. Energy distribution in the terrestrial solar spectrum (AM1.5).

Spectral region Wavelength

[nm]

Energy

[eV]

Contribution to

total spectrum [%]

near-UV 315-400 3.93-3.09 2.9

blue 400-510 3.09-2.42 14.6

green-yellow 510-610 2.42-2.03 16.0

Red 610-700 2.03-1.77 13.8

Near-IR 700-920 1.77-1.34 23.5

Infrared 920- >1400 1.34- <0.88 29.4

The key requirement for semiconductors is to absorb solar energy efficiently.

Visible light from 400 nm to 800 nm is desired to be the best range of solar

spectrum for solar energy conversion, which is corresponding to the band gap

of semiconductor from 3.12 eV to 1.56 eV. As mentioned previously in section

iii

Band gap can be calculate by the equation: Band gap (eV) = 1240/λ (nm).

Band positions usually shift with pH, in accordance with ‒ 0.059×pH (V).

13

1.3.3, band position is another major requirement for semiconductors besides

the band gap. Band gaps and positions of commonly used semiconductors are

summarized in Figure 10. Among these semiconductors, a number of

semiconductors with suitable band positions are good candidates for water

oxidation or water reduction respectively. For some semiconductors like NiO,

TiO2 and ZnO, their band positions are suitable for the total water splitting

reaction but their band gaps are bigger than 3.12 eV, leading to UV absorption

only rather than visible light.

Current research in this field is focusing on two parts: one is enhancing the

visible light absorption with smaller band gap semiconductors such as BiVO4,

Fe2O3, WO3 etc. and organic small molecular semiconductors; the other is

enhancing the visible light absorption with the concept of dye-sensitized

semiconductor.

-10

-9

-8

-7

-6

-5

-4

-3

-7.4

6

-7.4

1

-6.3

7

-6.3

8

-5.2

-5.2

-5.9

-7.3

7

-7.6

3-7.0

8

-7.3

7

-7.5

-6.9

-4.1

-4.2

1

-4.2

6

-3.9

8 -3.6

7

-3.8

-4.2

-4.8

3

-3.4

7

-4.5

-4.8

8

-4.6

7

-4.5

3.23

.2

2.7

2.4

1.4

1.7

1.12

.4

2.82.23.52.7

Zn

O

TiO

2

C3

N4

Cd

S

Cd

Te

Cd

Se

Si

Bi 2

O3

Cu

2O

Fe

2O

3

NiO

WO

3

E v

s. V

ac

uu

m

BiV

O4

2.4

5

4

3

2

1

0

-1

E v

s. N

HE

(p

H=

0)

H2O/H2

O2/H

2O

Figure 10. Band positions of semiconductors in the pH 0 aqueous electrolyte compared

with the energy potential for water splitting reaction. red = CB; blue = VB.

1.5.2. Molecular Photosensitizers

In the concept of dye-sensitized semiconductor PEC cells, dyes, pigments and

quantum dots etc. are used as photosensitizers. Two key issues should be

considered as shown in Figure 11:

1. For a photoanode, the excited state of the dye can inject an electron into

the CB of a semiconductor from its LUMO. While for a photocathode,

the excited state of the dye can inject a hole into the VB of a

semiconductor from its HOMO.

14

2. For a photoanode, the oxidation potential Eox of the dye should be more

positive than the onset potential of WOCs. For a photocathode the

reduction potential Ered of the dye should be more negative than the

onset potential of HGCs.

There are a large amount of dyes and pigments documented already, and

therefore we have to choose suitable dyes and combine them with wide band

gap semiconductors. However, it is challenging in practice to find a „„perfect‟‟

dye exhibiting all desirable properties with synthetic tractability, favorable

oxidation/reduction potentials, high extinction co-efficient, broad spectral

absorption and high photochemical stability. Derivatives of Ru(bpy)32+

[Tris(2,2'-bipyridine)Ruthenium(II)] have been widely studied for both water

oxidation and reduction (section 1.6.2), respectively. Dyes like Porphyrin

(TPP),79

and Perylene Bisimide (PDI) have also drawn a lot attention

recently.80

81

Figure 11. Schematic diagram, simplified electron flows of dye-sensitized

semiconductors. (left) for water oxidation, (right) for water reduction.

1.6. Functional electrodes for water splitting

As introduced above, sophisticated catalysts, light harvesting materials and

semiconductors have been developed for water splitting. The remaining

challenging task is to employ these man-made materials to build functional

devices. The study and fabrication of molecular-precise architectures on

interfaces has rapidly emerged as a scientific approach towards molecular

engineering. Tailor-designed molecules, supramolecular scaffolding and nano-

objects with conducting or semiconducting properties are efficient and

versatile tools for functional electrodes fabrication.82-84

Based on different

approaches, water splitting functional electrodes can be divided into two types:

electrochemical-driven water splitting electrodes and light-driven water

15

splitting electrodes which are respectively corresponding to “electrolyzer” and

“Artificial Photosynthesis”.

1.6.1. Electrochemical-driven water splitting

Recently, a great deal of novel relevant methods for immobilizing the

molecular catalysts on functional electrodes to construct efficient and stable

devices for electrochemical-driven water splitting has risen. So far, different

strategies concerning carbon surfaces (such as glassy carbon, graphite and

carbon nanotubes) and conductive metal oxides (FTO and ITO) have been

applied and reported in the literature (see the following corresponding

sections). These strategies mainly contain two kinds of fabrication methods:

covalent attachment and non-covalent attachment.

Covalent attachment

Covalent attachment mainly contains (1) the direct C−C bond linkage between

catalysts and carbon surfaces,85-88

(2) ligate molecular catalysts on the surface

of conductive metal oxides through carboxylic, phosphoric and siloxane ester

band linkage.89-91

C−C bond linkage attachment

(1) For the direct C−C bond linkage attachment, examples of immobilization

of diazoniumsalts for WOCs and HGCs by electrochemical-reduction have

been recently reported by our group and Fontecave et al., respectively.85,86

Figure 12. Examples of C−C bond linkage covalent modification method, (left) anode

from ref.[85], (right) cathode from ref.[86]

In these cases, diazoniumsalts derivatives were first electrodeposited onto a

GC electrode, further treatment with catalytic units yielded catalysts modified

electrodes. Structures of the electrodes are shown in Figure 12. Route for

modification of the glassy carbon electrode is shown in Figure 13.

16

Figure 13. The route of glassy carbon electrode C-C bond modification. R are the

linkage groups for catalysts.

Ester bond linkage attachment

(2) For ester bond linkage attachment, ITO (indium tin oxide) nano-particles

are usually used as the substrate for adsorbing catalysts. The hydroxyl groups

on the ITO form ester bonds with the carboxyl or phosphoric acid groups of

catalysts. As ITO is transparent, modifying catalysts on ITO surface is

beneficial for mechanistic studies with optical spectroscopy such as UV-vis,

spectro-electrochemistry etc. Figure 14 shows two examples of this type of

method. The left one is reported by Meyer et al. in which a WOC was

immobilized on conducting oxide surfaces. By optical spectroscopy study, the

authors claimed that the WOC retains the same mechanism as in homogenous

system.89

The right one is reported by Erwin Reisner et al., in which a cobalt

HGC with phosphonic acid groups as strong anchors to metal oxide surfaces

results in a functionalized electrode. This electrode shows high current density

for electrochemical proton reduction in pH 7 aqueous solution, and molecular

nature of the catalyst was studied by spectro-electrochemistry.92

Figure 14. Examples of ester linkage covalent modify catalysts on metal oxide, (left)

anode from ref.[89], (right) cathode from ref.[92].

Non-covalent attachment

Non-covalent attachment strategy is usually used for the molecular catalysts

immobilization on carbon based electrodes, which mainly contains: (1) π−π

stacking between carbon nanotubes and pyrene units of catalyst. (2)

Entrapment of the catalyst within polymer materials (polymer materials).

17

π−π stacking

(1) π−π stacking. Among the available functionalization approaches for CNTs,

non-covalent π−π stacking of conjugated molecules such as pyrene derivatives

on the carbon nanotube surface is quite attractive. As the extraordinary

electronic, thermal, optical, and mechanical properties of CNTs, this

methodology preserves desired properties of CNTs and allows tuning of the

catalyst loading by increasing the amount of supporting CNTs. Therefore, it is

believed that this approach is quite useful for the design of efficient electro-

catalytic-materials.93

This modification strategy usually need three steps: (i)

deposition of CNTs onto the electrode-supports; (ii) modification of the CNTs

with the pyrene contained catalysts and (iii) post-functionalization treatment

such as adding a protective layer like Nafion.

Recently, WOC and HGC functionalized CNTs through π−π stacking have

been demonstrated by our group and Vincent Artero‟s groups.94,95

The

structures of these electrodes are shown in Figure 15. For the anode,

[Ru(BDA)] catalyst was immobilized on CNTs via π−π stacking strategy, and a

sustained current density as high as 220 mA cm-2

was obtained at 1.4 V.(vs.

NHE) in pH 7 solution. For the cathode, [Ni(PPh

2NAr

2)2] catalyst was

immobilized on CNTs, and the resulting cathode could give a high current

density of 250 mA cm2

mgcatalyst-1

.

Figure 15. Examples of π−π stacking entrapment catalysts on CNTs, (left) anode from

ref.[95], (right) cathode from ref.[94].

Polymer materials

(2) Polymer materials. Through in-situ electro-polymerization, conductive

polymers can also be easily coated on carbon-based electrodes.96

For this

modification strategy, different processes have been used. For example,

electro-polymerized of pyrrole, thiophene etc. followed by functionalization of

the resulting polymer with catalysts is the one method, which is a two-step

process. The other is to directly electro-polymerize the catalyst which contains

groups such as pyrrole or thiophene. Immobilizations of WOCs and HGCs via

in-situ electro-polymerization have been recently demonstrated by our group

18

and C. J. Pickett‟s group, respectively.97,98

As shown in Figure 16, through

direct approach, [RuBDA] catalyst with a pyrrole unit was electro-polymerized

on glassy carbon electrode, and a high TOF of 10.47 s-1

at η = 0.77 V for water

oxidation was obtained on the electrode surface. For mimicking the active site

of [Fe-Fe] hydrogenases, C. J. Pickett et al. used the two-step process method

to prepare the cathode; a hydrogenase model compound was immobilized on

the glassy carbon electrode and the electro-driven proton reduction was

achieved

Among these strategies, C−C bond linkage attachment is desired benefit for

the stability; ligate molecular catalysts on the surface of conductive metal

oxides through ester bond is beneficial for mechanistic studies; CNTs based

electrodes through π−π stacking strategy can obtain high current densities; and

entrapment of the catalyst within polymer materials is quite straight forward.

There is no clear conclusion about which strategy is better. We believe that it is

important choosing a suitable strategy or strategies according to the

experimental motivation.

Figure 16. Examples for non-covalent entrapment of the catalyst within polymer

materials, (left) anode from ref.[97], (right) cathode from ref.[98].

1.6.2. Light-driven water splitting

Compared with electro-driven water splitting, there are no such varied

modification methods for the fabrication of light-driven water splitting

electrodes. According to the configurations and components, the strategies of

molecular-based photo-electrodes preparation can be divided into four methods:

(1) simple entrapment (catalyst/sensitizer), (2) co-adsorption

(catalyst+sensitizer), (3) link of catalyst and sensitizer (catalyst-sensitizer), (4)

hybrid catalyst and small band gap semiconductor together

(catalyst@semiconductor).

19

Simple entrapment

(1) Simple entrapment (catalyst / sensitizer). The first relevance for

construction of photoanodes was reported by Spiccia and co-workers,99

in

which a manganese Mn cluster catalyst was imbedded in a Nafion membrane

and coated on a sensitized TiO2 photoanode. Under visible light irradiation,

this photoanode can oxidize water in a pH 6.5 aqueous solution with a

photocurrent density of 31 μA cm−2

. In parallel, our group developed a dye-

sensitized photoanode by imbedding a [RuBDA] WOC into a pH-modified

Nafion membrane, which was also coated on a sensitized TiO2 photoanode.

Under the light illumination, molecular oxygen was confirmed by using this

photoanode in a PEC device.100

In 2012, our group conducted the first example

of dye-sensitized photocathode for visible light-driven proton reduction, in

which an organic dye sensitized NiO film was coupled with a cobaloxime

HGC by the simple entrapment strategy. This photocathode produced an initial

photocurrent density of ca. 20 μA cm−2

under visible light illumination, and

photo-produced hydrogen was confirmed.101

Recently, Yang‟s group

demonstrated a photocathode, in which a CdS quantum dot sensitized p-type

NiO film was used as the photocathode. By integrating this photocathode with

a cobaloxime catalyst in a photoelectrochemical cell light-driven proton

reduction with an initial photocurrent density of ca. 17 μA cm−2

was

achieved.102

Figure 17. Examples of simple entrapment strategy, photoanodes from ref.[99] and

[100], photocathodes from ref.[101] and [102].

20

The main drawback of simple entrapment strategy is the poor stability. The

catalysts can easily loose contact with the electrode surface resulting in fast

photocurrent decay. Co-adsorption strategy is considered as a solution to

overcome this problem.

Co-adsorption

(2) Co-adsorption. Recently, our group has published a series of photoanodes

based on co-adsorption strategy, which made a great contribution to the area of

dye-sensitized photoanode. First, a silane group was introduced into [Ru(BDA)]

catalysts as the anchoring group for the attachment of [Ru(BDA)] catalysts on

Ru dye-sensitized TiO2 films via chemical bonding.103

Figure 18. Examples of co-adsorption strategy, photoanodes from ref.[103-105],

photocathodes from ref.[106] and [107].

The resulting photoanode exhibited a significant high photocurrent density for

water splitting. An initial photocurrent density of 3.1 mA cm−2

was obtained at

pH 7. Then, the linkage length of catalyst was also investigated, and the

photoanodes with a longer flexible linkage showed a significant higher

photocurrent density than the shorter one.104

Considering the [Ru(BDA)]

WOCs catalyze water oxidation via a bimolecular reaction mechanism,

accordingly, a dimer molecular [Ru(BDA)] catalyst was synthesized and co-

adsorbed on the TiO2 film together with Ru sensitizer, resulting in an enhanced

21

performance.105

The relevance of the co-adsorption strategy for the hydrogen

generation dye-sensitized photocathodes construction has been firstly reported

by Hammarström and Ott.106

Although this photo-electrode has not been

assessed for hydrogen production, the electron transfer from light-driven

reduced dye to the catalyst was observed. Another related work was

demonstrated by our group recently, in which CdSe quantum dots and

cobaloxime catalyst were co-grafted on a NiO film as a photocathode, H2

evolution was achieved with a stable photocurrent.107

As electron transfer kinetics play an important role in light-driven water

splitting, linking catalyst and sensitizer together to improve the electron

transfer between them is considered as a promising strategy.108,109

Link of catalyst and sensitizer

(3) Linking catalyst and sensitizer together. Meyer and co-workers have

carried out several attempts to immobilize supramolecular assemblies, which

composed of Ru sensitizers and Ru-based WOCs onto the surface of nano-

porous semiconductor oxide.110-112

On the basis of this strategy, recently, a

catalyst-sensitizer dyad has been synthesized and anchored onto an ITO film.

Then, a thin TiO2 layer, which formed by atomic layer deposition (ALD), were

coated to protect the molecules from the desorption.113

This device achieved

visible light-driven water splitting into hydrogen and oxygen with stabilized

photocurrent density of 100 μA cm−2

. For photocathode, Wu and coworkers

constructed a sensitizer–catalyst dyad assembly super-molecular photo-catalyst

by binding of the classical cobaloxime catalyst on a pyridine-substituted Ru

sensitizer anchored onto alumina coated NiO. The resulting photo-electrode

exhibits a stable photocurrent of 10 μA cm−2

.114

Figure 19. Examples of strategy link catalyst and sensitizer together, photoanode from

ref.[113], photocathode from ref.[114]

As it is challenging to synthesize supramolecular catalysts with complex

structures, there is a novel and straightforward strategy to construct photo-

22

electrodes, which is immobilizing catalysts on visible light-absorbing

semiconductors to form hybrid materials.

Catalyst@semiconductor

(4) Hybrid catalyst and small band gap semiconductor. In 2012, Zhuang et al.

first reported photoanodes by combining hematite films and Ru WOCs

together and achieved light-driven water oxidation. This work opened a new

avenue for the solar-to-fuel conversion.115

Later, Fujita et al. reported a

photoanode based on the same strategy, in which a Ru WOC was adsorbed on

a WO3 film. Under the illumination of 1 sun AM 1.5 simulated sunlight, the

photoanode exhibits a cathodic shift of 60 mV and the photocurrent was

slightly improved.116

In 2014, Bartlett and coworkers successfully

demonstrated a photoanode by anchoring a molecular iron WOC onto WO3

film, and the rate of photo-electrochemical water oxidation was dramatically

increased with a high faradaic efficiency.117

Figure 20. Examples of the strategy Hybridizing catalyst and small band gap

semiconductor together, photoanodes from ref. [115] and [117], photocathodes from

ref.[118] and [119].

Very recently, Bernard Dam and coworkers combined a Ru based molecular

catalyst with a BiVO4 film, in which photocurrent at constant potential is

23

slowly decreasing in time with a photocurrent 1.3 mA cm-2

at 1.23 V (RHE)

bias potential.120

For photocathodes, Rose and coworkers constructed a

photocathode with a nickelbis(diphosphine) complex onto a methyl protected

p-Si, and the onset-potential for light-driven proton reduction is positively

shifted by 200 mV and proton reduction is confirmed.118

Moore and coworkers

used a similar strategy to immobilize the classical molecular HGC cobaloxime

on a poly-vinyl-pyridine (PVP) film coating GaP substrate. A stable activity

was obtained in aqueous media and hydrogen evolution was detected by gas

chromatography.119

Later, the performances of the cobaloxime-GaP

photocathode were further improved by changing the light intensity, pH value

of the electrolyte and different catalysts.121,122

Solar-to-fuel conversion technologies have been rapidly developed while more

research effort particularly in durability and performance is absolutely needed.

1.7. Challenges for Solar to Fuel Conversion

As mentioned above, the strategies, components and construction methods for

fabricating solar fuel conversion devices have been introduced. However, in

my opinion, one challenging aspect has to be solved before the construction of

a practical water splitting device; that is the “cost”.

Benefit is the main driving force for the transformation from new technology

to industry. Reducing the cost of devices and further decreasing the price of

solar fuels will not only give a bright future for this technology itself, but also

provide a feasible and sustainable energy system for our society.

Fundamental scientific research could provide the following points to reduce

the cost of solar fuels:

(1). Catalysts with high activity and devices with high efficiency.

(2). Catalysts with high stability and devices with long term durability.

(3). Components in low-cost and low-cost fabrication processes.

Mechanistic studies could provide the correlations between structures,

activities and stabilities for the design of highly active and stable catalysts.

Electrode kinetic studies could provide the information of how to improve

electron transfer and obtain highly effective devices.

Tailored construction methods may provide low-cost fabrication processes and

firm immobilization of catalysts onto electrodes for long term durability.

Employing low-cost components as much as possible can further reduce the

cost.

24

All the related research works are important for the development of solar fuel

technologies.

1.8. The Aim of This Thesis

The aim of this thesis has been to:

1) Immobilize WOCs onto the surface of conductive material towards

the functional electrode for electrochemical-driven water oxidation

with high efficiency.

2) Design functional electrodes for water oxidation, and exam their

corresponding kinetics of water oxidation on electrode surface to

obtain mechanistic information.

3) Immobilize WOCs on the surface of semiconductors towards

photoanodes with high stability for water oxidation.

4) Fabricate tandem PEC cells with two photo-active electrodes by using

the concept of Z-scheme via molecular engineering to replace the

high cost counter electrode material Pt.

25

2. Immobilization of a Molecular Catalyst on

Carbon Nanotubes for Highly Efficient Electrochemical-driven Water Oxidation

(Paper I)

26

27

2.1. Introduction

For implementation of any catalysts in electro-driven or light-driven devices

for water splitting, the catalysts have to be immobilized on the surface of

electrodes.123,124

However, challenges in modifying the substrate catalysts and

fabricating into devices hinder the further application of molecular devices,

and therefore, widely applicable approaches and well-designed modification

methods are urgently needed. At the same time, an essential question that

needs to be answered is “How to employ catalysts in functional devices

without affecting their performances?” Previously, our group has reported

two electrodes for electro-driven water oxidation. One is a [Ru(PDC)(pic)3]

modified glassy carbon electrode via click chemistry (as shown in Figure 12).

Only 0.06 mA cm-2

current density was obtained at 700 mV over-potential

while a high TOF of 1.6 s-1

was achieved, which is much higher than the TOF

of 0.23 s–1

evaluated from CeIV

-driven water oxidation.85

Another one is

[Ru(BDA)(Py-pyrene)2] modified multi-walled carbon nanotube (MWCNT)

based electrode through electrostatic π–π stacking interactions (as shown in

Figure 15).125

As [Ru(BDA)] catalyze water oxidation through a bimolecular

radical coupling pathway, immobilization of mononuclear [Ru(BDA)] WOCs

on electrode surface disfavors the bimolecular pathway, resulting in a low TOF

of 0.3 s-1

. However, a high current density of 220 mA cm-2

can be achieved

due to the introduction of MWCNTs which largely increase the surface area of

the electrode. It is interesting to find a immobilization method which can keep

activity of the catalysts and obtain a high performance of electrode at the same

time. Thereby, we designed a mononuclear WOC 15 based on [Ru(PDC)]

which can be immobilized onto carbon nanotubes via hydrophobic interaction,

the prepared electrode [15-MWCNTsCOOH@GDL] was expected to retain

the high activity of catalyst and receive a high efficiency of electrode for water

oxidation.

2.2. Synthesis

The synthetic route of the target complex 15 is shown in Figure 21. First, a

dodecyloxy group was introduced to diethyl 4-hydroxypyridine-2,6-

dicarboxylate by a simple alkylation reaction; after hydrolysis of the ester

group, 4-dodecyloxypyridine-2,6-dicarboxylic acid was obtained. It was then

reacted with cis-[Ru(dmso)4Cl2] in the presence of triethylaimine; 4‒picoline

was then introduced to the solution, giving the target complex 15.

28

Figure 21. Synthesis of complex 15 and the structure of the reference complex 6

2.3. Results and Discussions

2.3.1. Electrochemical properties

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.8 0.9 1.0 1.1 1.2 1.3

0.005

0.010

15

6

BLANK

I (m

A)

Potential V vs. NHE Figure 22. Cyclic voltammograms (CV) of 1 mM 15 (red line) and 6 (blue line) in 0.1M

potassium phosphate buffer (pH 7.0, 10% acetonitrile), with glassy carbon as the

working electrode, Ag/AgCl as the reference electrode, and Pt as the counter electrode.

Scan rate is 100 mV s-1.

To measure the electro-catalytic activity of 15 for water oxidation, cyclic

voltammetry (CV) measurements were conducted. Figure 22 shows CV

curves of 15 and 6 in pH 7.0 phosphate buffer solutions. Complex 15 exhibits

one irreversible RuIII/II

redox peak at E= 0.82 V vs. NHE, indicating the ligand

exchange process of 4‒picoline by water molecule during the redox reaction as

discovered earlier in our group with the reference catalyst 6. An onset potential

of water oxidation was observed around at 1.2 V for 15, which value is very

similar to 6. However, the catalytic current of 15 is higher than that of 6 at

29

potentials beyond 1.4 V, which can be ascribed to: (1) The electron donating

effect of dodecyloxy substituent in 15 is beneficial to accelerate the RDS in the

catalytic cycle. (2) The dodecyloxy chain of 15 is attractive to the surface of

glassy carbon electrode, leading to higher local concentration of 15 than that of

6 at electrode surface.

2.3.2. Preparation of Electrode

Figure 23. The solution of 15 (0.37 mg, 0.0005 mmol) in 10mL water containing 20%

trifluoroethanol (left). MWCNTs-COOH (0.15mg) dispersed in 10ml water containing

20% trifluoroethanol (right), and a mixed solution of 15 (0.37 mg, 0.0005 mmol) and

MWCNTs-COOH (0.15 mg) in 10mL water containing 20% trifluoroethanol (middle).

And SEM image of the surface of [15-MWCNTsCOOH@GDL] electrode.

The catalyst 15 was then immobilized on carboxyl functionalized multi-

walled carbon nanotubes (MWCNTsCOOH). After mixing

MWCNTsCOOH with 15, a colorless solution with precipitation was

observed as shown in Figure 23. This precipitation phenomenon can be

interpreted as follows: (1) because of the hydrophilic carboxyl groups,

MWCNTsCOOH can be dispersed in water; (2) when 15 is immobilized

on the surface of MWCNTsCOOH, the surface of CNTs becomes more

hydrophobic, leading to formation of precipitates. Notably, these

precipitates are the desired electro-catalytic nano-material. They can be

easily collected and applied on conductive substrates for further device

fabrication. In this work, the 15-MWCNTsCOOH precipitates were

collected by filtration through a conductive carbon substrate, so-called

gas diffusion layer (GDL), and a carbon based molecular modified

electrode [15-MWCNTsCOOH@GDL] was then obtained in such a

straight-forward way. The SEM image (Figure 23 right) shows that the

nano-structure of CNTs was well retained and presented a large surface

area after material 15-MWCNTsCOOH was collected by GDL. For

comparison, the pristine [MWCNTsCOOH@GDL] electrode as well as

the [6-MWCNTsCOOH@GDL] electrode was prepared by the same

procedure.

30

2.3.3. Electro-driven Water Oxidation

The surface of [15-MWCNTsCOOH@GDL] is highly hydrophobic, a rotating

disk electrode was employed to get rid of generated oxygen bubbles on the

surface and ensure an effective contact between electrode and electrolyte

during the measurements.

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8

0.0

2.5

5.0

7.5

10.0

12.5

Cu

rre

nt D

en

sity (

mA

cm

-2)

15-MWCNTsCOOH@GDL

6-MWCNTsCOOH@GDL

MWCNTsCOOH@GDL

GDL

Potential (V vs. NHE) Figure 24. CVs of [15-MWCNTsCOOH@GDL] electrode (blue), [6-

MWCNTsCOOH@GDL] electrode (red), without catalyst [MWCNTsCOOH@GDL]

(light blue), only GDL (black). All CV curves taken by using rotating disk electrode

(1000 rpm, 0.2 cm2) as working electrode, Ag/AgCl as the reference electrode, and Pt

as the counter electrode, in aqueous solution (phosphate buffer, pH 7.0, ionic

strength=0.1), with scan rate100 mVs-1.

To determine the stability of the physical attachments between these

components 15, CNTs and GDL, the [15-MWCNTsCOOH@GDL] electrode

was rotated at 1000 rpm for 2 h without applying any potential. CVs before

and after rotation were identical, which suggests a steady and stable

combination among the components of [15-MWCNTsCOOH@GDL].

To measure the electro-catalytic water oxidation activities of theses electrodes,

CV measurements were conducted. The results are shown in Figure 24. When

[MWCNTsCOOH@GDL] was used, no significant catalytic current could be

observed. [6-MWCNTsCOOH@GDL] electrode shows only a slight

enhancement of the catalytic current in comparison to the blank electrode

[MWCNTsCOOH@GDL]. In contrast, a significant enhancement of catalytic

current was observed when [15-MWCNTsCOOH@GDL] was used as working

electrode, which is because of the long hydrophobic aliphatic chain assisted

immobilization of 15 on MWCNTsCOOH. The [15-MWCNTsCOOH@GDL]

electrode exhibits an identical redox waves for the RuIII/II

following an uprising

current at around 1.15 V. The catalytic current density of [15-

MWCNTsCOOH@GDL] is up to 5.6 mA cm-2

at 1.4 V vs. NHE.

31

500 1000 1500 2000 2500 3000 3500

0

1

2

3

4

5

C

urr

en

t D

en

sity (

mA

cm

-2)

time (s)

0 2 4 6 8 10 12 14

0

1

2

3

4

Cu

rre

nt

De

nsity (

mA

cm

-2)

time (h)

Figure 25. Electrolysis of [15-MWCNTsCOOH@GDL] electrode at 1.3V (vs. NHE)

applied potential in 50mM phosphate buffer, pH = 7.0 electrolyte for 1 h (upper) and

15h (lower).

Bulk electrolysis was carried out by applying 1.3 V potential on [15-

MWCNTsCOOH@GDL] for time periods of 1 hour and 15 hours in 50 mM

potassium phosphate buffer (pH = 7.0). The corresponding amount of charge

was recorded. The values of charge were divided by 4F, yielding the

theoretical amounts of O2, which were 4.9 and 31.2 μmol for 1 hour and 15

hours electrolysis, respectively. The real amount of oxygen was calibrated by

GC, resulting 4.7 μmol O2 after 1 hour and 29.0 μmol O2 after 15 hours,

corresponding to Faraday efficiencies of 96% and 93%, respectively. After 15

h electrolysis at a relatively low potential of 1.3 V (vs. NHE), 0.84 mA cm-2

current density was retained. The decay of current density during the bulk

electrolysis suggests the decomposition of the catalyst.

The effective catalyst loading of WOCs (Γ0, mol cm-2

) on the [15-

MWCNTsCOOH@GDL] surface was estimated according to an established

method.126

An average catalyst loading of 8.5 ± 0.4 (10-10

mol cm–2

) was

obtained. This calculated average catalyst loading is far less than the total

amount of 15 used for preparing [15-MWCNTsCOOH@GDL] (0.24 mg cm–

32

2), which is due to that only the outside layer of the hybrid material can contact

with the electrolyte and respond to the CV measurements.

0

.8Q

TOF eqnt F

1 1=4

The electro-catalytic activity of [15-MWCNTsCOOH@GDL] was further

evaluated by chronoamperometric experiments. In typical experiments, the

current was recorded continuously under a sequence of applied potential steps

in the range from 1.12 to 1.32 V (Figure 27 upper). With the assumption of

100% Faraday efficiency, TOFs were calculated according to eqn. 8, where F

is the Faraday constant, Q the integrated charge through the electrode, Γ0 the

effective catalyst loading 8.5 ± 0.4 (10-10

mol cm–2

), t the integration time (30

s), and the number 4 is the equivalent constant because the generation of one

O2 molecule requires the transfer of four electrons.

0 50 100 150 200 250 300

-2

0

2

4

Curr

ent D

ensity (

mA

/cm

2)

time (s)

1.12V1.17V

1.22V

1.27V

1.32V

300 350 400 450 500

0.50

0.75

1.00

1.25

Overpotential (mV)

logT

OF

(s-1

)

3.3s-1

5.1s-1

7.5s-1

10.0s-1

13.7s-1

Figure 26. (upper) Chronoamperometric current densities measured in phosphate

buffer (pH 7.0, IS = 0.1M) under application of a sequential potential steps. (lower)

TOF plot of [15-MWCNTsCOOH@GDL] electrode as a function of overpotential η.

The logarithm of yielded TOFs varies linearly on the applied over-potentials

from 300 to 500 mV(Figure 27 lower). A TOF of 5.1 s–1

was achieved at η =

33

350 mV, and it reaches as high as 13.7 s–1

at η = 500 mV. The calculated initial

TOF is 12.4 s-1

, under the bulk electrolysis conditions at 1.3 V applied

potential. As the resistance losses are included in the measurement, all the

values are underestimated, and the actual TOF values should be bigger than the

given ones.

With the data of Γ0, the average value of TOFs based on the overall TON can

be calculated. Average TOFs of 7.6 s-1

for 1 hour electrolysis, and 3.5 s-1

for

15 hours electrolysis are achieved. The performance of [15-

MWCNTsCOOH@GDL] electrode and different anode electrodes for electro-

catalytic water oxidation from literature are listed in Table 2. [15-

MWCNTsCOOH@GDL] is clearly far superior than other systems in the table

in terms of applied potential, current density, TONs and TOFs.

Table 2. The performance of [15-MWCNTsCOOH@GDL] electrode and different

anode electrodes for electro-catalytic water oxidation from literature reported ones.

Applied

potential

[V vs.NHE]

pH

Current

density

[mAcm-2]

Faraday

efficiency TON TOFa) TOFb)

This work 1h

electrolysis 1.3 7.0 2.2 96% 27647 7.6

12.4

This work 15h

electrolysis 1.3 7.0 0.84 93% 170588 3.5

ITO\MWCNT\

[Ru(BDA)(Py-

pyrene)2] 20h125

1.4 7.0 0.22 96% 11000 0.3

ITO\MWCNT\

polyoxometalate

cluster 127

1.42 7.0 ≈0.2 - - - 0.085

ITO\ [Ru

Mebimpy]

13h89,91,128

1.8 1.0 - >95 % 28000 0.6

ITO\ [Ru

tpy]13h89,91,128 1.8 1.0 0.0067 >95 % 8900 0.3

a). Average TOFs in electrolysis experiments for 1h and 15h; b) the TOF in

chronoamperometric experiments is the initial TOF.

2.5. Summary

In summary, a molecular WOC 15 with dodecyloxy group was successfully

designed and synthesized. Complex 15 was immobilized on MWCNTs

through the hydrophobic interaction to form a hybrid nano-material. This

material was deposited onto GDL by filtration and formed a [15-

34

MWCNTsCOOH@GDL] electrode. This fabricated electrode not only

preserves the water oxidation activity of the molecular catalyst with an

essentially high initial TOF of 12.4 s-1

, but also achieved a surprisingly high

initial catalytic current density of 5.6 mA cm-2

(at 1.3 V applied potential).

This methodology provides a simple and reliable approach of transformation

from a homogeneous catalyst to a functional electrochemical electrode with

retained high activity for water oxidation.

35

3. Control the O‒O bond formation pathway by

immobilizing molecular water oxidation catalysts on the electrode surface

(Paper II)

36

37

3.1. Introduction

In the last chapter, we demonstrated a novel methodology which can keep the

activity of the catalysts and meanwhile obtain a functionalized electrode with a

good performance. Another interesting and important question is “Will the

mechanistic pathway of catalysts change when they are immobilized on the

electrode surface?” To answer this question, we need a methodology

corresponding to the question: “How to get kinetic information of RDS on the

electrode surface?”

Our group has recently developed two series of WOCs [Ru(PDA)L2] and

[Ru(BDA)L2] (PDA2

= 1,10-phenanthroline-2,9-dicarboxylate; BDA2

= 2,2-

bipyridine-6,6-dicarboxylate; L = N-cyclic aromatic ligands),129-134

and these

catalysts show high activities for water oxidation. For [Ru(PDA)] catalysts, O2

is produced via the water nucleophilic attack (WNA) pathway as shown in eqn.

6. In contrast, [Ru(BDA)] WOCs catalyze the oxidation of water through a

radical coupling (RC) pathway as described in eqn. 7. As shown in these two

pathways, for both [Ru(PDA)] and [Ru(BDA)], the OO bond formation are

triggered by RuV=O species. The WNA pathway is a first order reaction for

catalyst while the RC pathway is a second order reaction for catalyst.

Therefore two obvious questions can be raised: (1) is it possible that

[Ru(BDA)] catalysts catalyze water oxidation through the WNA pathway

without sacrificing the high activity of water oxidation if the catalytic centers

are far away from each other? (2) On the contrary, can the [Ru(PDA)] catalysts

go through the RC pathway if the two catalytic centers are close to each other?

The answers to these two questions will not only help us to understand the

mechanisms of molecular catalysts deeply, but also may provide new concepts

for the design of more efficient molecular WOCs. In this chapter, [Ru(PDA)]

and [Ru(BDA)] were chosen as model catalysts, and the electrochemical

polymerization method was used to control the distance of different catalytic

centers by using/without using co-polymer units, with a view to govern the

mechanisms of [Ru(BDA)] and [Ru(PDA)] on the electrode surface.

38

3.2. Design, Synthesis and Preparation of Electrodes

We have designed functional electrodes based on the following analysis:

(1) To force the [Ru(BDA)] catalysts to go through the WNA pathway, one can

prevent two RuV=O units from coupling to each other by immobilizing the

catalyst on electrode materials and meanwhile increasing the distance of the

[Ru(BDA)] units.

(2) To force the [Ru(PDA)] catalysts to go through the RC pathway, linking

[Ru(PDA)] catalysts together via proper organic linkers is a potential method.

In order to control the distance of different catalytic centers, the

electrochemical polymerization method was employed.135,136

Herein, 4-vinyl

pyridine was chosen as the polymerization functional group, and the target

complexes 16 and 17 were synthesized via microwave reactions by heating a

methanol solution containing the corresponding tetra-dentate ligand H2BDA or

H2PDA, cis-Ru(DMSO)4Cl2 and 4-vinyl pyridine in the presence of

triethylamine at 100℃ for 30 min. They were fully characterized by NMR,

elemental analysis and high-resolution MS.

Figure 27. Structures of complexes 16, 17 and schematic diagram of electrodes

prepared via electrochemical polymerization.

By using the method described by Meyer et. al.,137

the glassy carbon working

electrodes were cycled in the solution of 16 or 17 [0.5 mM in complex, 0.1 M

tetra-n-butylammonium hexafluorophosphate (TBAPF6) / acetonitrile] from 0

to -2.4 V (vs. Ag/AgNO3) by cyclic voltammetry. For poly-16+PSt@GC

electrode, the glassy carbon working electrode was cycled in a solution

containing 0.5 mM complex 16, 12.5 mM styrene and 0.1 M TBAPF6 /

acetonitrile from 0 to -2.4 V (vs. Ag/AgNO3) in successive scans. Poly-

39

16@GC and poly-17@GC electrodes were respectively obtained with

dimerization favored configuration. Poly-16+PSt@GC electrode is in

dimerization disfavored configuration, as shown in Figure 27.

3.3. Results and Discussions

3.3.1. CH3CN ligand exchange

Figure 28. 1H-NMR (500 MHz) spectra of (upper) 16 in CD3OD, (lower) 16 in

CD3OD/CD3CN (3:1).

Due to the C2V symmetry of complex 16, the BDA unit renders two doublets

(8.07 and 8.64 ppm) and one triplet (7.93 ppm) in d4-methanol. These three

peaks split into four peaks as shown in Figure 28, when d3-acetonitrile/d4-

methanol are employed as the solvent, which indicates that CH3CN can

coordinate to the Ru center at the same time one carboxylic acid of the BDA

ligand de-coordinates from the Ru center. A similar phenomenon was also

observed for complex 17, which are in agreement with our previous studies on

[Ru(BDA)] and [Ru(PDA)] catalysts with 4-picoline as the axial

ligand.133,134,138

When the electrochemical polymerization experiments are conducted in

acetonitrile containing solutions, acetonitrile-coordinating electrodes (poly-16-

NCCH3@GC) can be obtained. Thereby, we have prepared the polymerized

materials of 16 in two different solvents, acetonitrile and chloroform, resulting

in the corresponding poly-16-NCCH3@GC and poly-16@GC electrodes,

respectively. Their differential pulse voltammograms (DPVs) in pH 7.0

phosphate buffer solutions are depicted in Figure 29. The poly-16-

NCCH3@GC electrode displayed a higher potential of RuIII/II

(0.68 V vs. NHE)

than that of the poly-16@GC electrode (0.64 V vs. NHE). This result is in

agreement with the coordination of CH3CN to the [Ru(BDA)] center for the

poly-16-NCCH3@GC electrode.133,134,138

After the poly-16-NCCH3@GC

electrode was scanned from 0 and 1.6 V in the pH 7.0 phosphate buffer where

water oxidation happened, the original peak of RuIII/II

at 0.68 V disappeared

and a new peak emerged at 0.64 V; this phenomenon indicates that CH3CN can

40

be removed from poly-16-NCCH3@GC after an anodic scan. As chloroform

itself can easily generate radicals which may influence the distance control, in

this work, all the CV and PDV data were obtained by using the electrodes

prepared from acetonitrile via one anodic scan.

0.5 0.6 0.7 0.8 0.9 1.0 1.1Potential V vs. NHE

poly-16@GC obtained from CHCl3

poly-16@GC obtained from CH3CN (first scan)

poly-16@GC obtained from CH3CN (second scan)

Cu

rre

nt

1A

Figure 29. DPV curves of poly-16@GC electrodes prepared from different solvents

(acetonitrile and chloroform). Conditions: 50 mM pH 7.0 phosphate buffer solution;

scan rate 50 mV s−1.

3.3.2. pH Dependent Electrochemistry

The DPVs of poly-16@GC and poly-16+PSt@GC at various pH were

recorded, and the pourbaix diagrams were then constructed as displayed in

Figure 30. These two electrodes show similar pH-dependent redox properties.

The RuIII/II

couple is pH independent below pH 6 and becomes a proton-

coupled electron transfer (PCET) process in the region of pH 611; the latter is

ascribed as RuIII

-OH/RuII-OH2. Thereby, the pKa of Ru

III-OH2 is 6.

For RuIV/III

, a one-electron-one-proton PCET process can be observed over the

whole pH range from 2 to 11, corresponding to either RuIV

-OH/RuIII

-OH2 or

RuIV

=O/RuIII

-OH.

The oxidation of RuIV

=O only involves electron transfer at pH > 6, while

oxidation of RuIV

-OH is accompanied by a proton transfer below pH < 6.

The Pourbaix diagrams of poly-16@GC and poly-16+PSt@GC reveal that the

three oxidation peaks at pH 7 correspond to RuIII

-OH/RuII-OH2, Ru

IV=O/Ru

III-

OH and RuV=O/Ru

IV=O, respectively, and the OO bond formation are

triggered by RuV=O species for both electrodes, which is the foundation of this

work as we mentioned in the introduction of this chapter.

41

2 3 4 5 6 7 8 9 10 11

0.4

0.6

0.8

1.0

1.2

1.4

RuV-O or Ru

IV-O

RuIV

-OH

RuIV

-O

RuIII

-OH

RuII-OH2

Po

ten

tia

l V

vs

. N

HE

pH

RuIII

-OH2

2 3 4 5 6 7 8 9 10 11

0.4

0.6

0.8

1.0

1.2

1.4

pH

Po

ten

tia

l V

vs

. N

HE

RuII-OH2

RuV-O or Ru

IV-O

RuIV

-OH

RuIII

-OH2

RuIII

-OH

RuIV

-O

Figure 30. Pourbaix diagrams of poly-16@GC (upper) and poly-16+PSt@GC (lower).

3.3.3. Methods for Electrode Kinetics

The electrode kinetic study is a challenging task and very different with

homogeneous kinetic studies; the current method for electrode mechanistic

study mainly depends on time-resolved spectroscopy (electrochemistry, IR,

Raman, XPS). Very few reports can provide a clear evidence of the mechanism

about OO bond formation on electrode surface, and there is no report about

molecular WOCs which can go through RC pathway on electrode surface. In

this work, we focus on the difference of the OO bond formation between

WNA and RC pathways and provide a method for the catalytic OO bond

formation study on the functionalized electrode surface.

From eqn. 6 and 7 we can find two key differences between WNA and RC

pathways:

1) Reaction order for the O-O bond formation. The O-O bond formation is first

order in catalyst for the WNA pathway and becomes second order in catalyst

for the RC pathway.

42

2) Proton transfer. The proton transfer process is involved for the WNA step,

but not the case for the RC step, which can be reflected by kinetic isotope

effects (KIEs). At the same time, base as a proton acceptor can decrease the

barrier in the O‒O bond formation step for the WNA pathway which means

base is involved for the WNA pathway, but not the case for RC pathway.

According to the above analysis, to study the mechanistic pathways of poly-

16@GC, poly-16+PSt@GC and poly-17@GC electrodes, reaction orders and

KIEs should be examined. The reaction order is an empirical factor widely

used in chemical kinetic studies, which can give an insight into the events at

the molecular level for kinetics of complex reactions.139

log

log. 9

ii

k i

j

ceq

cn

In electrode kinetics, the reaction order (ρi) of species at concentration ci can

be determined from electro-kinetic measurements when the concentration of

all other species (ck≠i) are held constant, and is given by eqn. 9,139,140

where j is

a partial anodic current corresponding in our case to the catalytic current (mA),

and 𝑐𝑖 is the concentration of species (mol L-1

).

Consider that Γ (catalyst coverage mol cm-2

) can be accounted as the

concentration of catalyst on the electrode. And Γ has the linear relationship

with the peak current ip (here is the redox peak of RuII/III

) given by eqn. 10,126

where np is number of electrons (here for RuIII/II

np is 1), ν scan rate (V s-1

), A

surface area (cm2), F Faraday‟s constant (96485 C mol

−1), R ideal gas constant

(8.314 JK–1

mol–1

) and T temperature (298 K).

/

2 2

( ) .104

III IIp

p Ru eqnn F A

RTi

By combining eqn.9 and 10, the reaction order of catalyst coverage ρΓ can be

given by eqn. 11. To obtain ρΓ, only catalytic currents j and peak current ip of

RuII/III

are needed, which can be easily obtained by CV measurements.

log

log. 11

p

j

ieqn

The KIE values reflect the kinetic information of water oxidation reactions and

help us to interpret the rate determining step (RDS) of the catalytic

processes.141-144

The presence of KIE is considered as evidence that proton

transfer is involved in the RDS (or, at least, in one of the steps affecting the

reaction rate). At the same time, absence of KIE is widely accepted as a line of

evidence for the RDS not involving the proton transfer.145

The KIE for water

oxidation catalytic reactions can be divided into two kinds on the basis of

different reaction mechanisms and RDSs:

43

(1) When the RDS is the water nucleophilic attack process (eqn. 6), the OH

bond cleavage is involved, the reaction shows a primary isotope effect, and the

value of KIEH/D is usually more than 2.141-144

(2) When the RDS is the dimerization of two RuV=O units (eqn. 7), no such

OH bond cleavage is then involved, and the reaction will show a secondary

isotope effect in which the value of KIEH/D is usually between 0.7 and

1.5.27,138,146

By combining the reaction order and kinetic isotope effect, the kinetic

information and mechanistic pathways should be realistically reflected.

3.3.4. The O‒O Bond Formation Pathways on Electrodes

Firstly, the poly-16@GC electrode was examined. By controlling the time of

polymerization or scan numbers of CV experiments, poly-16@GC electrodes

with various Γ were obtained. As shown in Figure 31A, the anodic current

increased with the increase of Γ under the same conditions. According to eqn.

11, the reaction orders ρΓ of poly-16@GC at various Γ are calculated to be

around 2 at different applied potentials. This value indicates that water

oxidation catalyzed by poly-16@GC is a second order reaction in catalyst

[Ru(BDA)] and reveals a RC pathway for the O-O bond formation.

The reaction order of phosphate was also examined. As the analysis for the

WNA pathway before, atom–proton transfer (APT) with base (B) as a proton

acceptor can decrease the barrier in the O‒O bond formation step as shown in

eqn. 12, which means that base such as phosphate is usually involved in the

WNA process.147,148

By keeping the catalyst coverage Γ and other parameters

constant and only changing the concentration of phosphate buffer (Na2SO4 salt

is added to maintain the ionic strength to be 0.1 M), the reaction orders of

phosphate (ρpi) are calculated (Figure 31B) according to eqn. 9, giving a ρpi= 0

for poly-16@GC which means that phosphate is not involved in the RDS. This

is also in a good agreement with that poly-16@GC catalyzes water oxidation

through the RC pathway.

KIEs can reflect the kinetic information of the RDS. Here, we controlled the

catalyst coverage Γ of poly-16@GC and other parameters, changed the

phosphate buffer to 50 mM Na2SO4 H2O or D2O solutions. CV curves of poly-

16@GC in H2O and D2O were compared (Figure 31C). A secondary isotope

effect was observed as evidenced by the KIE(H/D) values less than 1.5 in the

range of 1.1 V to 1.4 V, suggesting that the O-H bond cleavage is not involved

44

in the RDS by using poly-16@GC electrode. According to ρΓ, ρpi and KIE data,

we can conclude that poly-16@GC goes through the RC pathway.

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.3

0.6

0.9

1.2

1.2V vs. NHE

1.3V vs. NHE

1.4V vs. NHE

1.5V vs. NHE

-3.60 -3.55 -3.50 -3.45 -3.40

-2.4

-2.0

-1.6

-1.2

log(ip of RuII/Ru

III mA)

log

j (m

A )

y = a + b*x Value R^2

Slope 1.2V 1.86 0.986

Slope 1.3V 1.97 0.951

Slope 1.4V 1.95 0.977

Slope 1.5V 1.91 0.961

Potential V vs. NHECu

rre

nt

De

ns

ity

(m

A c

m-2

)

A

loa

din

g in

cre

as

e

50mM

100mM

200mM

300mM

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.6

1.2

1.8

2.4

3.0

-1.25 -1.00 -0.75 -0.50-3

-2

-1

0

log [Pi] (m ol L-1)

log

j (

mA

)

1.2V vs. NHE

1.3V vs. NHE

1.4V vs. NHE

y = a + b*x

Slope 1.4V 0.011

Slope 1.3V -0.043

Slope 1.2V 0.107

Potential V vs. NHE

Cu

rre

nt

De

ns

ity

(m

A c

m-2

)

B

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.1

0.2

0.3

0.4

1.1 1.2 1.3 1.4 1.50.0

0.5

1.0

1.5

2.0

Potential V vs. NHE

KIE

( j

H2

O /

j D

2O

)

Potential V vs. NHE

Cu

rren

t D

en

sit

y (

mA

cm

-2)

poly-16 in H2O

poly-16 in D2O

C

Figure 31. (A) CVs of poly-16@GC at various catalyst coverage Γ. Inset is the plot of

log (j) versus log (iRuIII/II); the slopes of the linear fitting curves reflect the reaction

orders in catalyst coverage. (B) CVs of poly-16@GC at various phosphate buffer

concentration [Pi]. Inset is the plot of log (i) versus log ([Pi]); the slopes of the linear

fitting curves reflect the reaction orders in [Pi]. (C) CVs of poly-16@GC in D2O and

H2O (electrolyte = Na2SO4; 50 mM). Inset shows the KIE values versus E.

45

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.3

0.6

0.9

1.2

1.5

1.3V vs. NHE

1.4V vs. NHE

1.5V vs. NHE

1.6V vs. NHE

-4.2 -4.1 -4.0 -3.9 -3.8 -3.7 -3.6

-2.4

-2.0

-1.6

-1.2

log

j (

mA

)

logip of RuII

/RuIII

(mA)

y = a + b*x Value R^2

slope 1.3V 1.12 0.976

slope 1.4V 1.18 0.922

slope 1.5V 1.31 0.984

slope 1.6V 1.29 0.991

Potential V vs. NHE

Cu

rre

nt

de

ns

ity

(m

A c

m-2

)

load

ing

incre

ase

A

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.2

0.4

0.6

0.8

-1.4 -1.2 -1.0 -0.8 -0.6 -0.4

-2.0

-1.8

-1.6

-1.4

-1.2

lo

g j

(m

A)

1.6V vs.NHE

1.5V vs.NHE

1.4V vs.NHE

y = a + b*x

Slope 1.3V 0.174

Slope 1.4V 0.196Slope 1.5V 0.212

Slope 1.6V 0.222

1.3V vs.NHE

log [Pi] mol L-1

Potential V vs. NHE

50mM

100mM

200mM

300mM

Cu

rre

nt

de

ns

ity

(m

A c

m-2

)

B

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.1

0.2

1.1 1 .2 1 .3 1 .4 1 .50 .0

0 .5

1 .0

1 .5

2 .0

2 .5

KIE

(

j H2

O /

j D2

O )

P o ten tia l V vs. N H E

poly 16+PSt in H2O

poly 16+PSt in D2O

Potential V vs. NHE

Cu

rre

nt

De

ns

ity

(m

A c

m-2

)

C

Figure 32. (A) CVs of poly-16+PSt@GC at various catalyst coverage Γ. Inset is the

plot of log (j) versus log (iRuIII/II); the slopes of the linear fitting curves reflect the reaction

orders in catalyst coverage. (B) CVs of poly-16+PSt @GC at various phosphate buffer

concentration [Pi]. Inset is the plot of log (i) versus log ([Pi]); the slopes of the linear

fitting curves reflect the reaction orders in [Pi]. (C) CVs of poly-16+PSt @GC in D2O

and H2O (electrolyte = Na2SO4; 50 mM). Inset shows the KIE values versus E.

Next, the poly-16+PSt@GC electrodes were prepared by co-polymerizing 1-

equivalent catalyst 16 and 25-equivalent styrene on the glassy carbon

46

electrode. As shown in Figure 27, the catalytic centers of poly-16+PSt@GC

are considered to be far away to each other because poly-styrene (PSt)

blocking units were added to separate the catalyst units.

The kinetics of poly-16+PSt@GC was examined by using the same method as

above. As shown in Figure 32, ρΓ of poly-16+PSt@GC was calculated to be

1.2, ρpi is 0.2 and KIEs(H/D) is 2 at 1.4 V vs. NHE. The value of ρΓ (1<ρΓ<2)

indicates that poly-16+PSt@GC oxidizes water through both WNA and RC

pathways with the former WNA pathway as the dominating one. This is

because that a small amount of catalytic centers in a short distance in still have

chance to meet each other the polymer, resulting in mixed pathways for poly-

16+PSt@GC.

Since the WNA pathway is dominating, the value of ρpi was observed to be

0.2. In comparison with the ρpi of poly-16@GC (ρpi = 0), the value of ρpi = 0.2

clearly revealed that phosphate is involved in the RDS for poly-16+PSt@GC,

indicating that the atom–proton transfer process occurs. At the same time,

KIE(H/D) values of poly-16+PSt@GC show a primary isotope effect, and

thereby the OH bond cleavage is involved in the RDS. In consideration of ρΓ,

ρpi and KIE, poly-16+PSt@GC mainly goes through the WNA pathway.

Clearly, the WNA pathway is favored for water oxidation when catalytic

centers are far away to each other.

1.2 1.3 1.4 1.5 1.6

0

2

4

6

8

10

12

14

poly-16@GC

poly-16+PSt@GC

TO

F (

s-1)

Potential (V vs. NHE)

Figure 33. Plot of TOFs versus potential.(●) poly-16@GC; (▲) poly-16+PSt@GC.

TOFs were calculated according to eqn.8 in chapter 1.

Turnover frequencies (TOFs) of poly-16@GC and poly-16+PSt@GC were

compared (Figure 33). At similar catalyst coverage (1.2×10-11

mol cm–2

for

poly-16@GC and 1.8×10-11

mol cm–2

for poly-16+PSt@GC), poly-16@GC

shows much higher TOF than that of poly-16+PSt@GC at the same applied

potential. For example, poly-16@GC gives a TOF of 11.2 s-1

at 1.5 V vs. NHE

while poly-16+PSt@GC just shows a TOF of 3.9 s-1

. Accordingly, the

[Ru(BDA)] catalyst on the electrode surface catalyzes water oxidation more

effectively via the RC pathway than the WNA pathway

47

10 20 30 40 50

0.001

0.002

0.003

0.004

0.005

0.006

0.007

Cu

rre

nt

De

nsity (

mA

cm

-2)

Time (s)

Figure 34. Chronocoulometry of poly-17@GC with different catalyst coverage.

In contrast with [Ru(BDA)] catalysts, we have earlier shown that the

[Ru(PDA)] water oxidation catalysts catalyze CeIV

-driven water oxidation via

the WNA path in homogeneous system.133

The OO bond formation energy

barriers for a [Ru(PDA)] catalyst were calculated to be 13.3 kcal mol-1

via the

WNA pathway and 16.4 kcal mol-1

via the RC pathway. The energetic

difference between these two processes is only 3.1 kcal mol-1

. Will the

[Ru(PDA)] catalyst change the O-O bond formation mechanism from the

WNA pathway to the RC pathway if the [Ru(PDA)] units are enforced close to

each other? Can we really make the energy barrier for the RC pathway less

than that of the WNA pathway through the electro-polymerization method?

To address these questions, catalyst 17 was polymerized on glassy carbon by

the same method as for poly-16@GC, resulting in the poly-17@GC electrode.

As the RuII/III

peak of catalyst 17 is not reversible, eqn. 10 is not suitable for

poly-17@GC electrodes. While the catalyst coverage Γ can also be calculated

by eqn. 13,149

where Q is the integrated current under the RuIII/II

of poly-17, F

Faraday‟s constant (96485 C mol−1

), n the number of electrons transferred (n =

1), and A the area of the electrode (~ 0.07 cm2).

.13Q

nFAeqn

Then the corresponding ρΓ can be calculated by eqn. 14.

log

log.14

jeqn

To calculate the catalyst coverage Γ, chronocoulometry of poly-17@GC was

operated in 50 mM potassium phosphate buffer pH 7.0. To exclude the effect

of capacitance, first, 0.6V vs. NHE potential were added to working electrodes

for 10s, then 0.85V vs. NHE potential were added to the working electrodes

for 50s, and the current densities were recorded as shown in Figure 34.

48

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.00

0.25

0.50

0.75

-9.35 -9.30 -9.25 -9.20-2.0

-1.8

-1.6

-1.4

1.6V vs. NHE

1.5V vs. NHE

log mol cm-2

log j

(m

A)

y = a + b*x value R^2

Slope 1.45V 1.98 0.95

Slope 1.5V 1.98 0.94

Slope 1.6V 1.92 0.91

1.45V vs. NHE

Potential (V vs. NHE)

Cu

rre

nt

de

ns

ity

(m

A c

m-2

)

load

ing

incre

as

e

A

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.5

1.0

1.5

2.0

-1.4 -1.2 -1.0 -0.8 -0.6 -0.4

-5.0

-4.5

-4.0

-3.5

-3.0

l

og

j (m

A)

log [Pi] mol L-1

1.3V vs. NHE

1.4V vs. NHE

1.5V vs. NHE

y = a + b*x

1.3V Slope -3.4364E-4

1.4V Slope 0.02456

1.5V Slope 0.0342

1.6V Slope 0.06551.6V vs. NHE

50mM

100mM

200mM

300mM

Potential V vs. NHE

Cu

rre

nt

de

ns

ity

(m

A c

m-2

)

B

0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.1

0.2

0.3

0.4

0.5

1.0 1.2 1.4 1.60.0

0.5

1.0

1.5

2.0

Potential V vs. NHE

KIE

(j

H2

O /

j D

2O

)

Potential V vs. NHE

poly-17 in H2O

poly-17 in D2O

Cu

rren

t d

en

sit

y (

mA

cm

-2)

C

Figure 35 (A) CVs of poly-17@GC at various catalyst coverage Γ. Inset is the plot of

log (j) versus log (Γ); the slopes of the linear fitting curves reflect the reaction orders in

catalyst coverage. (B) CVs of poly-17@GC at various phosphate buffer concentration

[Pi]. Inset is the plot of log (i) versus log ([Pi]); the slopes of the linear fitting curves

reflect the reaction orders in [Pi]. (C) CVs of poly-17@GC in D2O and H2O

(electrolyte = Na2SO4; 50 mM). Inset shows the KIE values versus E.

Poly-17@GC electrodes were examined by the same methods. As shown in

Figure 35, ρΓ of poly-17@GC is calculated to be around 1.9, ρpi is 0 and

KIE(H/D) is around 1.1. These three values suggest that poly-17@GC goes

through the RC pathway. The RC pathway is favored when catalytic centers

are close to each other, like the case of polymerization of [Ru(PDA)] on

49

electrode surface, although the [Ru(PDA)] catalyst itself favors the WNA

pathway in the homogeneous system. This is the first example clearly showing

that we can switch the mechanisms of a catalyst from the WNA to the RC

through catalyst engineering.

3.4. Summary

In summary, molecular WOCs 16 and 17 with 4-vinyl pyridine as axial ligands

were successfully polymerized on glassy carbon electrodes, and the resulting

poly-16@GC and poly-17@GC electrodes were obtained. By co-polymerizing

catalyst 16 with styrene, poly-16+PSt@GC electrode was also prepared. The

reaction orders in catalyst and phosphate as well as the kinetic isotope effects

for electrochemical water oxidation were examined. Both poly-16@GC and

poly-17@GC electrodes catalyze the electrochemical water oxidation reaction

through the radical coupling (RC) pathway, even water nucleophilic attack

pathway is dominated for catalyst 17 in the homogenous system. On the

contrary, the poly-16+PSt@GC electrode with styrene as blocking units favors

the WNA pathway. Comparing the TOFs of poly-16@GC with poly-

16+PSt@GC, the RC pathway shows a faster reaction rate than that of the

WNA pathway. These results provide guidance for controlling the reaction

mechanism of molecular water oxidation catalysts on the electrode surface.

Our work provides solid evidences for understanding the reaction mechanisms

of water oxidation catalysts on the electrode surface.

50

51

4. Immobilizing Ru catalysts on photoanodes for light-driven water splitting

(Paper III and IV)

52

53

4.1. Introduction

In the last two chapters, we have discussed anodes for electro-driven water

oxidation, which are related to the indirect approach for solar fuel conversion.

In the following two chapters, we will discuss the PEC cells for direct solar

fuel conversion.

Recently, our group has published a series of photoanodes based on [Ru(BDA)]

catalysts. We had constructed a photoanode in which the [Ru(BDA)] catatlyst

was encapsulated in the Nafion film covered onto a Ru dye-sensitized TiO2

film and achieved light-driven water splitting.150

Later on, phosphonic acid and

silane groups, were introduced on [Ru(BDA)] catalysts as anchoring groups for

the attachment of [Ru(BDA)] catalysts on the Ru dye -sensitized TiO2 films via

chemical bonds.104,151

The resulting photoanode gave significant high

photocurrent density for water splitting. Since the [Ru(BDA)] WOCs catalyze

water oxidation via a bimolecular reaction mechanism and the rate determining

step of the catalytic cycle is the dimerization of two RuV=O species,

131

immobilization of mononuclear [Ru(BDA)] WOCs on the electrode surface

disfavors the bimolecular pathway. Accordingly, we synthesized a dimer

molecular [Ru(BDA)] catalyst with an anchoring group and co-adsorbed it on

the TiO2 film together with Ru dye, achieving an enhanced photoanode

performance.105

In the last chapter, we have studied the O-O bond formation mechanistic

pathways on electrodes prepared via electrochemical polymerization method.

As the RC pathway shows a faster reaction rate than that of the WNA pathway,

we would like to use poly-16 on photoanodes for light-driven water oxidation

in this chapter. We will try to find the reason(s) for the fast photo-current decay

through kinetic studies. Further, we will give a solution about “How to

improve the stability of photoanodes?”

54

4.2. Synthesis and Preparation of Electrodes

Figure 36. The chemical structures of complexes 16, 18 and RuP.

The catalyst 16 was prepared by the method as described in the last

chapter. For comparison, catalyst 18 was also prepared by a simple

reaction as shown in Figure 37. To prepare the photoanode, TiO2 films

were prepared according to our previous report.151

Figure 37. Synthesis route of complex 18

Briefly, 18 nm anatase TiO2 paste was screen-printed on FTO glass, then

annealed at 500 ºC for 30 min. The thickness of obtained bare TiO2 film

is ca. 8 μm. The active areas of the photoelectrodes were 1 cm2.

18+RuP@TiO2 electrodes were prepared according to our previous

work.151

Poly-16+RuP@TiO2 electrodes were prepared as follows:

complex 16 was polymerized on RuP sensitized TiO2 films by

electrolysis of a complex 16 solution (1 mM in 100 mM

TBAPF6/acetonitrile solution) at 2.0 V (vs Ag/AgNO3) for 200 seconds.

All of the tests were operated under 300 mW cm-2

of light density.

55

4.3. Results and Discussions

4.3.1. Electrochemical Properties

After the polymerization of catalyst 16 on RuP@TiO2, the prepared poly-

16+RuP@TiO2 electrode showed an orange color (Figure 38 inset), while the

RuP@TiO2 electrode showed a light yellow color. The difference in colors

indicates the formation of the catalyst polymer in the case of poly-

16+RuP@TiO2 electrode.

0.4 0.6 0.8 1.0 1.2 1.4 1.6

0.0

0.2

0.4

0.6

0.8

0.4 0.6 0.8 1.0 1.2

0

5

10

15

Potential(vs. NHE)

Cu

rre

nt

A

poly-16@TiO2

Potential V vs. NHE

Curr

ent

(mA

)

poly-16+RuP@TiO2

RuP@TiO2

Figure 38. Upper, CV curves of poly-16+RuP@TiO2(red line), 18+RuP@TiO2(blue

line) and RuP@TiO2(black line). Lower, DPV curves of poly-16+RuP@TiO2(red line)

and RuP@TiO2(blue line) and inset poly-16@TiO2.

From cyclic voltammogram (Figure 38 upper), 18+RuP@TiO2 electrode

displays a redox peak of RuII/Ru

III at 0.65 V vs. NHE followed by a catalytic

wave with an onset potential Eonset = 1.1 V vs. NHE. For that of poly-

16+RuP@TiO2 a clearly redox peak of RuII/Ru

III can be observed at 0.70 V vs.

NHE, followed by a catalytic wave with an Eonset = 1.15 V. Besides the

RuII/Ru

III peak at 0.65 V, the PDV (Figure 38 lower) of the poly-

16+RuP@TiO2 electrode exhibits RuII/Ru

III peak at 0.70 V and Ru

III/Ru

IV peak

56

at 0.89 V, following a RuII/Ru

III peak of RuP overlapping with Ru

IV/Ru

V(1.1V)

peak of [Ru(BDA)]. The DPV of 18+RuP@TiO2 electrode is similar with our

previous work.151

These electrochemical results indicate that

thermodynamically, a photo-generated RuIII

P can oxidize [Ru(BDA)] from

RuII to Ru

III–H2O, Ru

IV–OH and even till Ru

V=O species.

4.3.2. Photo-Electrochemistry

A 3-E PEC cell was set up with the photoanode poly-16+RuP@TiO2 as

working electrode, Ag/AgCl as reference electrode and Pt net as counter

electrode. First, LSV experiments were carried out as shown in Figure 39. For

poly-16+RuP@TiO2 under light illumination, the photocurrent rapidly

increased with the rise of applied potential from -0.25 V to 0.20 V (vs. NHE).

The photocurrent reached a plateau at E > 0.20 V with a photocurrent density

of 1.4 mA cm-2

, which is in agreement with our earlier reported results.104,151

The LSV of 18+RuP@TiO2 electrode is similar with our previous work.151

-0.25 0.00 0.25 0.50 0.75 1.00 1.25-0.5

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Potential V vs. NHE

Curr

ent

Density (

mA

cm

-2)

poly-16+RuP@TiO2 Light

poly-16+RuP@TiO2 Dark

Figure 39. LSV measurements of the WEs under light illumination through a 400 nm

long-pass filter (light intensity 300 mW cm-2), scan rate= 50 mV s-1

The photocurrent of the photoanode under a bias potential of 0.2 V vs. NHE

was measured. A remarkable initial photocurrent density of about 3.0 mA·cm−2

was obtained for the poly-16+RuP@TiO2 photoanode, while for the

RuP@TiO2 photoanode, no significant photocurrent could be observed.

However, the photocurrent of the poly-16+RuP@TiO2 photoanode is unstable

with time. As shown in Figure 40, the photocurrent density decreased from 3

to 0.3 mA cm-2

after 200 seconds. For 18+RuP@TiO2, the photocurrent

density decreased even faster. This instability problem is a general problem for

all developed photo-active electrodes at the moment.104,105,151

Improvement of

the long time durability of photo-electrodes is urgently needed for the further

applications of this type device.

57

0 100 200 300 400 500

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

poly-16+RuP@TiO2

18+RuP@TiO2

RuP@TiO2

time (s)

Cu

rre

nt

de

nsity (

mA

cm

-2)

Figure 40. The full time photocurrent of illumination on photoanodes under an applied

potential of 0.2 V vs. NHE in the PEC, operated in a 50 mM pH 7.0 phosphate buffer

solution.

4.3.3. Kinetics Study

Kinetic information may help us to find the reason of the fast photo-current

decay. Kinetic isotope effects (KIEs) of poly-16+RuP@TiO2 and

18+RuP@TiO2 were studied to reveal the RDS of our photoanodes during the

catalytic processes. KIE reflects the kinetic information of water oxidation

reactions as we have introduced in the last chapter.

Firstly, the electro-catalytic KIEs of poly-16+RuP@TiO2 and 18+RuP@TiO2

electrodes were compared under dark conditions in order to exclude the effect

of photo-induced electron transfer. Chronoamperometric experiments were

conducted in both H2O and D2O solutions with 100 mM Na2SO4 as supporting

electrolyte. Upon applying potentials on poly-16+RuP@TiO2, the current

densities showed negligible difference between H2O and D2O solutions. The

KIEH/D value was calculated as ca. 1.1 (Figure 41 upper).

Under the same conditions, the 18+RuP@TiO2 electrode shows a KIEH/D of

ca.1.9, which is significantly different from that of poly-16+RuP@TiO2.

Accordingly, poly-16+RuP@TiO2 has a secondary isotope effect and the

18+RuP@TiO2 electrode has a primary isotope effect. Therefore poly-16 was

proposed to catalyze water oxidation through the RC pathway, and in contrast,

complex 18 was proposed to catalyze water oxidation via the WNA pathway or

mixed pathways. The results obtained here are in agreement with the

conclusion from last chapter. The distance between catalytic centers will affect

the mechanism of water oxidation.

58

0 50 100 150 200 250

0.0

0.5

1.0

1.5

2.0

2.5

1.5V

time (s)

Cu

rre

nt

den

sit

y (

mA

cm

-2)

poly-16+RuP@TiO2 in H2O

poly-16+RuP@TiO2 in D2O

1.2V

1.3V

1.1V

1.4V

a

0 50 100 150 200 250

0.0

0.5

1.0

1.5

time (s)

Cu

rre

nt

de

ns

ity

(m

A c

m-2)

18+RuP@TiO2 in H

2O

18+RuP@TiO2 in D

2O

1.1V

1.2V

1.3V 1.4V

1.5V

30 40 50 60 70

0.0

0.3

0.6

0.9

Cu

rre

nt

den

sit

y (

mA

cm

-2)

time (s)

poly-16+RuP@TiO2 in H2O

poly-16+RuP@TiO2 in D2O

80 100 120

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

Cu

rre

nt

De

ns

ity

(m

A c

m-2)

Time (s)

18+RuP@TiO2 in H2O

18+RuP@TiO2 in D2O

Figure 41. (upper). Chronoamperometric current densities measured under application

of a sequential potential steps. (Lower).The transient current responses to on˗off cycles

of illumination on photoanodes under an applied potential of 0.2 V vs. NHE operated in

a 100 mM Na2SO4 H2O and D2O solution.

To study the photocurrent density decay, the KIEs measurements were also

conducted under light illumination as shown in Figure 41 lower. The

photocurrents of either poly-16+RuP@TiO2 or 18+RuP@TiO2 show no clear

difference in H2O and D2O solutions under illumination, which means that the

RDSs for these two electrodes are not related to the O-H bond cleavage for

light-driven water oxidation. This result for 18+RuP@TiO2 seems

contradictory to the KIE = 1.9 under electro-driven conditions. We believe the

water oxidation mechanism of catalyst 18 for light-driven water oxidation still

keeps the same as that under electro-driven conditions, i.e. WNA, but it is not

the rate limiting step for the whole PEC process. In another word, the KIEs of

light-driven water oxidation reflect the RDS of the whole PEC process instead

of the catalytic cycle of catalysts. Since the electron injection from dye into

TiO2 is ultrafast,152

and the catalytic process of catalyst is not the RDS, it

means that the photo-induced electron transfer from catalyst to the oxidized

RuP is not efficient enough as shown in Figure 42. A reasonable explanation

is that the regeneration of RuP becomes the RDS for the whole PEC process

for 18+RuP@TiO2 and the same may be also applicable for poly-

16+RuP@TiO2.

59

Figure 42. Simplified electron flows of catalyst-dye-semiconductor.

In light-driven reactions, the RDS is not the oxidation of water by the catalyst,

which is also observed from other photosystems,153

and the slow regeneration

of RuP can lead to the decomposition/desorption of RuP.154

Therefore,

photocurrent decay under long time light illumination is probably caused by

the decomposition/desorption of dye due to the slow electron transfer from

catalyst to dye+.

4.3.4. Replacement of the Molecular Photosensitizer by α-Fe2O3

Based on the above analysis, there are several strategies which can be used to

improve stability of photo-electrodes during long time illumination. Obviously,

linking catalyst and dye together to improve the electron transport between

them can be a reliable strategy. Related work has been demonstrated by our

group,155

where a [Ru(BDA)] catalyst and a Ru-based dye were linked by Zr4+

ions bridge and a relatively stable photocurrent was obtained. Recently, Meyer

et al. published that the same catalyst 16, propylene carbonate and a Ru-based

dye were copolymerized on TiO2. A low photocurrent of 40 μA cm-2

was

achieved, but a relatively stable photocurrent can be obtained.137

These two

works support that enhancing the electron transport between catalyst and

photosensitizer will increase the stability of this type of PEC device.

Combining with our kinetic study in this work, we have proved that the fast

electron transfer is beneficial for the stability of the device.

Another more straightforward strategy can be also proposed intuitively, which

is replacing RuP by a more stable photosensitizer. Here we would like to

demonstrate our efforts on replacing RuP@TiO2 by an α-Fe2O3 nano-rod array

in order to get a stable photocurrent.

The α-Fe2O3 nano-rod array serves as the photosensitizer and semiconductor.

Again, catalyst 16 was electrochemically polymerized on the Fe2O3 nano-rod

electrodes. An experiment of transient current responses to on˗off cycles of

light on poly-16@Fe2O3 electrode under an external bias of 0.6 V vs. NHE was

then performed as shown in Figure 43. An initial photocurrent density of 600

μA cm−2

was obtained by using the poly-16@Fe2O3 photoanode. In

comparison, the bare α-Fe2O3 nano-rod electrode yielded an initial

photocurrent density of 300 µA·cm−2

. For long time light illumination, a

60

remarkable stable photocurrent of ca. 400 μA cm-2

can be retained for poly-

16@Fe2O3 photoanode. Although poly-16@Fe2O3 photoanode does not

produce a photocurrent as high as poly-16+RuP@TiO2, photosensitizer

decomposition can be substantially suppressed by using α-Fe2O3 comparing to

RuP@TiO2, and a stable photocurrent can be obtained.

0 100 200 300 400 500 600

0.0

0.2

0.4

0.6

0.8

Time s

Cu

rre

nt

De

nsity (

mA

cm

-2)

poly-16@Fe2O3

Fe2O3

Figure 43. The transient current responses to on˗off cycles of illumination on photo-

anodes under an applied potential of 0.6 V vs. NHE in the 3-E system PEC with Pt as

counter electrode, operated in a 50 mM pH 7.0 phosphate buffer solution.

4.4. PDC as the Anchoring Group for Photoanode

As above, we have found the main reason of photocurrent decay during long

time illumination for the co-adsorption strategy, and two potential solutions

have been demonstrated. However, more extensive and universal methods to

improve the stability of photoanodes are required.

For the molecular PEC cells, in the normally used Nafion film, carboxylic or

phosphate acid as anchoring group of molecular catalyst can detach from the

electrode surface in aqueous solution easily, yielding significant photocurrent

decrease in a very short time, which is another reason of photocurrent

decay.99,156

A good anchoring group which can tolerate broad pH range will be

very helpful for the improvement of device stability.

Very recently, we developed an organic dye with 2,6-pyridinedicarboxylic acid

(PDC) as the anchoring group, which showed an impressive stability under a

long-term test.157

Even in a basic desorption experiment for a long time, this

novel anchoring group still exhibited a strong attachment to the metal oxide

film. Therefore it is expected that a stable photocurrent can be obtained by

using PDC as the anchoring group in the catalyst@Fe2O3 photoanode.

As introduced in chapter one, most of molecular Ru WOCs catalyze water

oxidation through the WNA pathway. Here, to demonstrate a more

61

representative model, complex 6 was chosen as the model WNA catalyst. By

introducing PDC as the anchoring group, complex 19 (Figure 44) was prepared.

Figure 44. Synthesis route of complex 19.

We immobilized the molecular catalyst 19 by soaking the Fe2O3 films in

solution of 19 overnight, resulting in the 19@ Fe2O3 photoanodes. It is well

known that alkaline solution is more preferable for water oxidation because

high pH value can lower the onset potential and accelerate the rate of water

oxidation.158

However, ester bond linkage can be easily hydrolyzed at high pH

values, therefore PEC cells involving molecular ester bond linkage are rarely

tested under alkaline conditions.

We examined the performances of the 19@ Fe2O3 photoanode in 1 M KOH

solution. From the LSV as shown in Figure 45, 19@Fe2O3 displays rapid

photocurrent increase under light illumination when the bias is above 0.1 V (vs.

NHE). At 0.6 V bias potential, a high photocurrent density ca. 3 mA cm-2

can

be achieved. For bare Fe2O3 electrode, only 2 mA cm-2

can be obtained, which

indicates the essential catalytic role of the catalyst 19 in the system.

-0.2 0.0 0.2 0.4 0.6 0.8

0

2

4

6

Voltage (V vs. NHE)

Fe2O3 dark

Fe2O3 light

19@ Fe2O3 dark

19@ Fe2O3 light

Cu

rre

nt

de

nsity (

mA

cm

-2)

Figure 45. LSV curves of the bare Fe2O3 electrode and catalyst modified electrode

19@Fe2O3 in 1 M KOH solution under visible light illumination and dark conditions.

62

0 100 200 300 400 500 600 700

0

1

2

3

4

5

Fe2O3

19@Fe2O3

Cu

rre

nt d

en

sity (

mA

cm

-2)

Time (s)

0 2000 4000 6000 8000 100000

1

2

3

4

Cu

rre

nt d

en

sity (

mA

cm

-2)

Time (s)

Fe2O3

19@Fe2O3

Figure 46. The photocurrent decay curves, (upper) and transient photocurrents, (lower)

continuously illumination on potoanodes in 1 M KOH solution, 0.6 V vs. NHE bias was

applied.

As shown in Figure 46, both of 19@Fe2O3 and Fe2O3 electrodes showed

photocurrent with excellent stability for over 10 000 seconds. The dark current

dropped to nearly zero when the light was switched off, confirming that the

obtained current under light illumination was indeed photocurrent. Such stable

photocurrent indicates the high stability of our molecular-based photoanode

under alkaline conditions. So far, the stability of molecular catalysts on

photoanodes for water splitting was relatively poor.99,151,156,159,160

To our delight,

in our case, the molecular catalyst based electrode 19@ Fe2O3 exhibits a stable

photocurrent in 1 M KOH, which is an impressive result.

4.5. Summary

In summary, molecular [Ru(BDA)] catalysts 16 and 18 are successfully

immobilized on RuP sensitized TiO2 films, and poly-16+RuP@TiO2 and

18+RuP@TiO2 electrodes are obtained. The KIE studies on poly-

63

16+RuP@TiO2 and 18+RuP@TiO2 suggest that poly-16 undergoes water

oxidation via the radical coupling mechanism and 18 on the electrode surface

catalyzes water oxidation via water nucleophilic attack pathway. For the

photoelectrochemical reaction, the RDS is proposed to be the regeneration of

RuP, and the decomposition of photosensitizer RuP limits the stability of the

whole PEC devices.

By replacing dye-sensitized semiconductor RuP@TiO2 with α-Fe2O3

semiconductor, a stable photocurrent for poly-16+RuP@Fe2O3 electrode can

be obtained. Furthermore, we employed a new anchoring group 2,6-

pyridinedicarboxylic acid and successfully prepared complex 19. The

molecular catalyst based electrode 19@Fe2O3 exhibits impressive stability

under harsh reaction conditions (1 M KOH).

These results can inspire researchers for the design of more stable PEC devices

for water splitting.

64

65

5. Dye-sensitized tandem PEC cells for light driven total water splitting by

employing the concept of Z-scheme

(Paper V and VI)

66

67

5.1. Introduction

The conversion of solar energy into solar fuels by water splitting is one of the

most promising strategies for sustainable energy systems. Much attention has

been paid to the construction of either anode or cathode, where typically

oxygen and hydrogen can be generated respectively. However, some

challenges still remain in the construction of a full PEC cell.

One challenge is the stability, which is caused by the decomposition and

desorption of catalyst and photosensitizer from the photo-electrode surface.99

In the last chapter, we have introduced a new anchoring group 2,6-

pyridinedicarboxylic acid (PDC) for WOC 19, and replaced the dye-sensitized

semiconductor RuP@TiO2 with an α-Fe2O3 semiconductor. The molecular

catalyst based electrode 19@Fe2O3 exhibits impressive stability under harsh

conditions (1 M KOH). The PDC anchoring group can provide a super strong

binding between the catalysts and the metal oxide surfaces.

Another challenge is cost. The Pt electrode is often used as the counter

electrode for oxygen or hydrogen generation. However, the scarcity and high

price of Pt make it impractical for global-scale applications. Therefore it is

very important and valuable to construct PEC devices without Pt. In this

chapter, we will demonstrate “How to utilize the concept of Z-scheme to

fabricate Pt-free molecular PEC cells”.

The Z-scheme from Natural Photosynthesis is a sophisticated design for water

splitting. In Artificial Photosynthesis, the Z-scheme has been achieved by

some inorganic material based devices.161-164

However, there is no report on

molecular catalyst based devices employing Z-scheme to fabricate Pt-free PEC

cells with both photoanode and photocathode for water splitting. Applying

molecular catalysts instead of Pt in PEC devices can reduce the cost of solar

fuel, which is beneficial for the widely practical application of solar driven

water splitting.

68

5.2. Rational Design of Tandem Dye-sensitized PEC cells

To make rational design of a Z-schemed tandem dye-sensitized-PEC (DS-PEC)

cell, several key issues should be considered:

(i) The energy levels of the photoanode and the photocathode should match

with each other. Here we use TiO2 as the photoanode material and NiO as the

photocathode material, which is similar to the design of tandem pn-DSSCs.165

(ii) The n-type dye can inject an electron into the CB of TiO2 from its LUMO,

meanwhile, the p-type dye can inject a hole into the VB of the NiO from its

HOMO.

(iii) For the photoanode the dye‟s oxidation potential Eox should be more

positive than the onset potential of the WOC, while for the photocathode the

dye‟s reduction potential Ered should be more negative than the onset potential

of the HGC.

(iv) According to our previous study,104

the distance between the dye and the

surface of semiconductor should be shorter than the distance between the

catalyst and the surface of semiconductor to facilitate the desired electron/hole

injection and subsequent electron transfer.

5.3. Synthesis and Preparation of Electrodes

Following the above considerations, RuP was first chosen as the

photosensitizer, complex 19 was chosen as the WOC, and a HGC complex 20

with PDC as the anchoring group was synthesized as shown in Figure 47.

Figure 47. Synthesis route of complex 20

69

TiO2 and NiO films were prepared according to our previous

reports.151,166

The thicknesses of the obtained bare TiO2 and NiO film are

ca. 8 μm and 1 μm, respectively. The bare TiO2 and NiO films were

dipped into 0.3 mM RuP water solution for 1 h. After rinsed by water

and ethanol, the photosensitized TiO2 and NiO films were immersed in

saturated solution (CH2Cl2) of catalyst 19 and solution of catalyst 20

(methanol) overnight respectively.

5.3. Results and Discussions

5.3.1. Electrochemical Properties

RuP and catalysts are essential in the electron/hole-transfer processes

occurring in the completed tandem PEC cell as shown in Figure 48. Briefly,

the photosensitizer RuP is excited by visible light, followed by electron or hole

injection to respective photo-electrodes.

Figure 48. Schematic energy diagram for 19, 20, RuP, FTO, TiO2 and NiO (all data

shown at pH 7).

For the photoanode, the excited state of RuP transfers an electron to the CB of

TiO2 and forms an oxidation state RuP+. The catalyst 19 transfers one electron

to the RuP+, after four times electron-transfer processes, water oxidation

occurs on the photoanode side. For the photocathode, the reduced

photosensitizer RuP─

is generated after hole injection from the excited state of

70

RuP to the VB of NiO. By extracting holes from the catalyst 20, RuP is

regenerated and hydrogen generation occurs on the photocathode side.

5.3.2. Photo-Electrochemistry

Transient photocurrent curve of 19+RuP@TiO2/Pt cell is shown in Figure 49

upper. Under 0.2 V bias vs. NHE, an average photocurrent density of ca. 100

μA cm-2

was produced when 19+RuP@TiO2 was irradiated by visible light.

Figure 49 lower exhibits the transient photocurrent curve of 20+RuP@NiO/Pt

PEC cell under the bias of -0.2 V vs. NHE. An obvious cathodic photocurrent

of ca. -13 μA cm-2

can be obtained when the light was switched on. This

cathodic photocurrent indicates that the photocathode electrode with molecular

catalyst 20 is indeed photo-active.

60 80 100 120

-100

0

100

200

300

light off

Curr

ent

density (A

cm

-2)

Time (s)

light on

110 120 130 140 150 160

-15

-10

-5

light on

light off

Curr

ent densi

ty (A

cm

-2)

Time (s)

Figure 49. (upper) Transient photocurrents of individual 19+RuP@TiO2 electrode

under 0.2 bias and (lower) 20+RuP@NiO electrode under -0.2 bias vs. NHE. Pt wire

was used as the counter electrode and Ag/AgCl electrode was used as the reference

electrode.

A PEC cell with 2-E configuration was constructed by connecting the

19+RuP@TiO2 photoanode and the 20+RuP@NiO photocathode. In this

designed PEC device, the bias was applied between the photoanode and

71

photocathode, which can be irradiated by visible light from both photo-

electrode sides simultaneously.

100 110 120 130 140 150-20

0

20

40

60

80

light on

light offC

urr

ent

density (A

cm

-2)

Time (s)

0 20 40 60 80 100 120 140 160 180 200

10

20

30

40

50

60

Cu

rre

nt d

en

sity (A

cm

-2)

Time (s) Figure 50. Transient photocurrent (upper) and photocurrent decay (lower) of

19+RuP@TiO2/20+RuP@NiO tandem PEC cell, with 0.4 V bias applied between the

two photoelectrodes, in pH 7 phosphate buffer solution.

Figure 50 upper demonstrates the transient photocurrent generated upon

the on-off cycles of light illumination for the tandem PEC cell

19+RuP@TiO2/20+RuP@NiO. Under 0.4 V bias between the

photoanode and photocathode, a significant photocurrent was observed.

The photocurrent dropped to the baseline when light was switched off.

The rise-fall of the photocurrents corresponded well to the switching on-

off of the illumination. Such rapid and reversible photocurrent responses

by the light chop indicates the high sensitivity of the photo-response of

the tandem PEC cell and efficient separation of the electron-hole pairs

generated by visible light. In comparison to the 20+RuP@NiO/Pt cell

which has a photocurrent density of 13 μA cm-2

, the average

photocurrent density of the tandem PEC cell is ca 40 μA cm-2

. This

implies that the designed PEC system based on both photoanode and

photocathode is more effective than only photocathode alone for water

splitting, and tandem PEC cell has great potential for the fabrication of Pt

free devices.

72

Molecular catalysts 19 and 20 with strong anchoring group PDC were used

respectively for WOC and HGC in photoanode and photocathode, and a

designed Pt-free tandem molecular PEC cell was fabricated for the first time.

5.4. Redesign and Replacement of RuP

In order to avoid the use of noble metal-Pt, a tandem molecular PEC cell for

water splitting has been demonstrated, which indicates that the Pt-free tandem

molecular PEC cell can achieve total water splitting driven by visible light.

However, the high cost of Ru-based photosensitizer is another limitation for

the future large-scale application. Due to the high efficiency, easy modification,

tunable electron transfer process and low cost, metal-free organic dyes have

been widely used in dye-sensitized solar cells. Thereby, organic dyes can be

considered as the alternatives for Ru-based photosensitizers in PEC devices.167

Recently, encouraging works have been demonstrated, such as Finke et al. used

perylene diimide as an n-type semiconductor to drive CoOX for water oxidation

and a photocurrent of 150 μA cm-2

was obtained,81

but a very high bias was

required (1.0 V vs. Ag/AgCl) for this photoanode. Mallouk et al. employed

porphyrin dyes to drive IrOX for water oxidation, but the photocurrent (less 50

μA cm-2

) and stability were not satisfactory.79

So far, no organic molecule can

provide a better performance than that of Ru-based dye for light-driving water

splitting in PEC device.

To replace RuP, according to the rules of rational design for tandem dye-

sensitized PEC cells we have introduced before, a tandem DS-PEC cell was

redesigned, in which the photoanode is co-sensitized by a simple organic dye

L0 and catalyst 19 on meso-TiO2, and the photocathode employs an organic

dye P1 and catalyst 21 co-sensitized on NiO, as shown in Figure 51.

Figure 51. The representation of the photoanode with organic dye L0 and catalyst 19

on TiO2, and the photocathode with organic dye P1 and catalyst 21 on NiO

73

The synthesis of HGC 21 was shown in Figure 52. And electrodes were

prepared by the similar procedure of 19+RuP@TiO2 and 20+RuP@NiO

electrodes, resulting in 19+L0@TiO2 photoanode and 21+P1@NiO

photocathodes.

Figure 52. Synthesis of complex 21

According to the electrochemical properties of the dyes, catalysts and

semiconductors (as shown in Table 3), a schematic energy diagram was

illustrated in Figure 53. When the photoanode was irradiated by light, dye L0

can inject an electron into the CB of TiO2 and the photo-generated L0+ can

oxidize the catalyst 19, leading to water oxidation after multi-electron transfer

processes. Meanwhile, at the photocathode, dye P1 can inject hole into the VB

of NiO and the formed P1─ which can reduce the catalyst 21 for eventual

hydrogen generation.

First, as shown in Figure 54 (upper), the LSV of 19+L0@TiO2 in 3-E setup

under illumination demonstrates that the photocurrent rapidly increased with

the applied potential from -0.25 V to -0.15 V (vs. NHE), and reached a plateau

at E > -0.1 V with a photocurrent density of 0.42 mA cm-2

. This value is

significantly higher than the current density under dark conditions, which

Table 3. Optical and Electrochemical Properties of the Dyes and Catalysts

Dyes and

Catalysts

λabs (ε/M-1 cm1)

/ nm

λem

/

nm

Eox(HOMO)

/ V vs. NHE

E0-0

/ eV

LUMO

/ V vs.

NHE

P1 348(34720);481(57900) 618 1.32 2.25 – 0.93

L0 373 387(36000) 509 1.37 2.90 -1.53

19 19 onset potential for water oxidation on TiO2 is around 1.20 V

21 21 onset potential for hydrogen generation on NiO is around -0.54 V

74

indicated that the working electrode is indeed photoactive. Here 0.2 V vs. NHE

as the applied bias was selected for the photocurrent measurements.

Figure 53. Schematic energy diagram for L0, P1, 19, 21, FTO, TiO2 and NiO (all data

shown at pH 7).

Transient current responses to on˗off cycles and full time photocurrent under

illumination were studied and corresponding results are shown in Figure 54

(lower). For 19+L0@TiO2 photoanode, a remarkable average photocurrent of

ca. 300 μA cm-2

was produced. For the L0@TiO2 photoanode without catalyst

19, only ca. 30 μA cm-2

photocurrent density was produced under the same

conditions. The significant increase of photocurrent confirms the highly

catalytic activity of 19 for water oxidation, and the electron transfer from the

catalyst to the oxidized dye should occur as anticipated. In comparison to the

electrode 19+RuP@TiO2 which only gave a photocurrent density of ca. 100

µA cm-2

, the higher photocurrent density obtained from the photoanode

19+L0@TiO2 indicates that this simple organic dye L0 based device can

provide much better device performance than that of the expensive

[Ru(bpy)3]2+

photosensitizer based device. Additionally, compared with

photoanodes reported in the literatures,79,81

our photoanode 19+L0@TiO2

shows advances of low applied bias potential and high current density.

75

-0.25 0.00 0.25 0.50 0.75 1.00

0.0

0.2

0.4

0.6

0.8

Potential V vs. NHE

Cu

rre

nt

De

ns

ity

(m

A c

m-2

)

19+L0@TiO2 ligth

19+L0@TiO2 dark

Scanning direction

0 100 200 300 400 500 600

0.0

0.2

0.4

0.6

0.8

19+L0@TiO2

L0@TiO2

time (s)

Cu

rre

nt

De

ns

ity

(m

A c

m-2

)

on

off

Figure 54. (upper) LSV measurements of the WEs under light illumination (light

intensity 100 mW cm-2), scan rate = 50 mV s-1. (lower)The transient current responses

to on˗off cycles and full time photocurrent of illumination (light intensity 100 mW cm-2)

on photoanodes under an applied bias potential of 0 V vs. Ag/AgCl in a 3-E PEC cell

with Pt as counter electrode, operated in a 50 mM pH 7.0 phosphate buffer solution.

In a long term illumination experiment, the photocurrent of 19+L0@TiO2 showed a relatively slow decay (Figure 54 lower). In our case, the stronger

PDC in comparison to the commonly used carboxylic acid and phosphonic

acid anchoring groups was used for catalyst 19.157,168

The organic dye L0 is

insoluble in water, which also can hinder the desorption of L0 from the surface

of TiO2. Due to these reasons, our photoanode 19+L0@TiO2 shows a

reasonable good stability.

76

-0.6 -0.4 -0.2 0.0 0.2 0.4

-60

-40

-20

0

Potential V vs. NHE

Cu

rre

nt

De

ns

ity

(A

cm

-2)

21+P1@NiO dark

21+P1@NiO light

Scanning direction

100 200 300 400 500 600

-80

-60

-40

-20

0

time ( s )

Cu

rre

nt

De

ns

ity

(A

cm

-2)

P1@NiO

21+P1@NiOon

off

Figure 55. (upper) LSV measurements of the WEs under light illumination (light

intensity 100 mW cm-2), scan rate = 50 mV s-1. (lower)The transient current responses

to on˗off cycles and full time photocurrent of illumination (light intensity 100 mW cm-2)

on photocathodes under an applied potential of -0.2 V vs. Ag/AgCl in a 3-E PEC cell

with Pt as counter electrode, operated in a 50 mM pH 7.0 phosphate buffer solution.

The LSV of the assembled 21+P1@NiO/Pt in a 3-E setup showed rapid

photocurrent increase with the applied potential from 0.4 V to -0.1 V (vs. NHE),

and the photocurrent reached a plateau at E < -0.1 V with a photocurrent

density of ca. -45 μA cm-2

(Figure 55 upper). The transient current responses

to on˗off cycles showed that the 21+P1@NiO photocathode could produce an

average photocurrent of ca. -35 μA cm-2

(at -0.2 V vs. Ag/AgCl applied

potential), while the reference photocathode P1@NiO without the catalyst 21

produced only -4 μA cm-2

current under the same conditions. Compared to our

previous work where P1 and [Co(dmgBF2)2(H2O)2] were encapsulated together

on the NiO film as shown in Figure 17,101

the photo-stability of the present

device was significantly improved, indicating that the phosphonic acid

anchoring group used in the catalyst 21 benefits for the photo-stability of this

PEC device.

77

-1.00 -0.75 -0.50 -0.25 0.00 0.25 0.50 0.75

-100

-80

-60

-40

-20

0

20

40

Potential V

21+P1@NiO/19+L0@TiO2 dark

21+P1@NiO/19+L0@TiO2 light

Scanning directionCu

rre

nt

de

ns

ity

(A

cm

-2)

100 200 300 400 500 600

-300

-250

-200

-150

-100

-50

0

21+P1@NiO/Pt

21+P1@NiO/19+L0@TiO2

time (s)

Cu

rre

nt

De

ns

ity

(A

cm

-2)

on

off

Figure 56.(upper) LSV measurements of the tandem cell under light illumination (light

intensity 100 mW cm-2), scan rate= 50 mV s-1, in a 2-E PEC cell. (lower) The transient

current responses to on˗off cycles and the full time photocurrent of illumination on

photoelectrode in 2-E setups without any bias. Operated in a 50 mM pH 7.0 phosphate

buffer solution (light intensity 100 mW cm-2).

With both photoanode and photocathode in hands, we prepared a tandem PEC

cell with 2-E configuration. The working electrode (WE) is the 21+P1@NiO

photocathode, and the counter electrode (CE) is the 19+L0@TiO2 photoanode.

From the LSV of this tandem PEC cell 21+P1@NiO/19+L0@TiO2 as shown

in Figure 56 (upper, it was found that the photocurrent rapidly increased with

the changing of applied potential from 0.8 V to 0.6 V, and reached a plateau at

E < 0.6 V with a photocurrent density of ca. -70 μA cm-2

). It clearly shows that

this tandem PEC cell can work without any applied bias (0 V bias, -70 μA cm-

2). Transient current responses to on˗off cycles and full time photocurrent

under illumination without bias then were studied as shown in Figure 56 lower.

Firstly, the photocurrent density of 21+P1@NiO/Pt in the 2-E setup was much

lower (5 μA cm-2

, Figure 56 blue) than the 3-E setup one (35 μA cm-2

, Figure

55 red). This behavior can be explained as follows: in the 2-E setup, the

valence band of NiO locates around 0.5 V vs. NHE, which is not high enough

78

to drive water oxidation, that‟s the possible reason of 21+P1@NiO/Pt PEC cell

almost does not work in the 2-E setup. While, in 3-E setup, the Pt counter

electrode can get extra potential from electrochemical workstation for water

oxidation. Secondly, the photocurrent density of the tandem cell

(21+P1@NiO/19+L0@TiO2) shows a significant enhancement (ca 70 μA cm-2

)

compared to 21+P1@NiO/Pt. This enhancement is due to the replacement of

Pt by 19+L0@TiO2 photoanode, in this case, the photo-generated electrons and

holes can flow in this PEC cell as shown in Figure 53. The difference between

21+P1@NiO/Pt and 21+P1@NiO/19+L0@TiO2 PEC cells indicates that a

good photoanode can not only provide protons for the hydrogen generation

half-reaction, but also can assist the charge flow between two electrodes. It is

in agreement with 19+RuP@TiO2/20+RuP@NiO tandem cell we have

demonstrated before.

In a full time photocurrent measurement, photo-generated hydrogen was

collected and measured by GC from the tandem PEC cell

(P1+Co1@NiO/L0+Ru1@TiO2) with the 2-E setup without external

bias for 100 min illumination. In total, 0.33 μmol H2 was produced with

0.117 C of charges passing through the electrodes, which corresponds to

55% Faraday efficiency. 2

-2

( [ ] (1.23V ) )e .15

[mWcm ]

op bias F

STH

light

J mAcm Vqn

P

With the Faraday efficiency, the performance of the tandem PEC cell can be

assessed by the corresponding solar-to-hydrogen conversion efficiency ηSTH

with eqn. 15,14,15,169

where Jop is the effective operating current density

measured during device operation (70 μA cm-2

), Vbias is the bias voltage that

can be added in series with the two electrodes (0 V), Plight is the incident light

power (100 mW cm-2

), ηF is Faraday efficiency (55%). Accordingly, ηSTH of

this tandem device is calculated to be 0.05%.

A monochromatic incident photon-to-electron conversion efficiency (IPCE)

measurement was performed. As shown in Figure 57, the tandem PEC cell

(21+P1@NiO/19+L0@TiO2) shows an IPCE of 25.2% at 380 nm (maxabs of

L0), 3.9% at 480 nm (maxabs of P1). This IPCE indicates both L0 and P1

contribute for the photon-to-electron conversion in this tandem PEC cell. This

is an advantage of using the concept of Z-scheme with different dyes to

achieve a broader absorption of visible light.

This is the first case that organic dyes are employed as photosensitizers in both photoanode and photocathode of a tandem molecular PEC device for light-driven total water splitting.

79

400 450 500 550 600 650

0

10

20

30

IPC

E %

Wavelength (nm)

Figure 57. The IPEC spectra of the tandem cell in 50 mM pH 7 phosphate buffer, in 2-

E setup without any bias voltage (data points are from average of three independent

experiments).

5.5. Summary

Molecular catalysts 19 and 20 with a strong anchoring group pyridine-2,6-

dicarboxylic acid (PDC) were used respectively for oxygen and hydrogen

generation in the photoanode and photocathode of a PEC cell, and a well-

designed Pt-free tandem molecular PEC cell was fabricated for the first time.

The tandem molecular PEC cell 19+RuP@TiO2/20+RuP@NiO can split water

with a reasonably high steady photocurrent density under visible light

illumination under neutral pH condition.

To replace high cost dye RuP, organic dyes L0 and P1 were used in the

photoanode and photocathode, respectively. Visible light driven water

oxidation by using photoanode 19+L0@TiO2 was successfully achieved.

19+L0@TiO2 can produce a remarkable average photocurrent of ~ 300 μA cm-

2 for water oxidation under neutral conditions. For photocathode 21+P1@NiO,

both photo-activity and photo-stability were improved compared to our

previous reported system. A tandem DS-PEC cell was designed and prepared

by connecting the 19+L0@TiO2 photoanode and 21+P1@NiO photocathode.

For the first time, a metal free organic dyes sensitized Z-schemed PEC cell that

can split water by visible light under neutral pH conditions without bias was

obtained.

Our results provide a new guidance for the design of low-cost Pt-free

molecular PEC devices for artificial photosynthesis in the future.

80

81

6. Concluding Remarks

The first parts (Chapters 2 and 3) of this thesis describe our efforts towards

developing molecular based anodes for electrochemical-driven water oxidation.

Two aspects of this issue are concerned.

One is the immobilization of dodecyloxy group modified WOC 15 onto the

surface of MWCNTs through the hydrophobic interaction, forming a hybrid

electro-catalytic nano-material. This material was further deposited onto GDL

by filtration, giving the [15-MWCNTsCOOH@GDL] electrode. This

fabricated electrode not only preserves the water oxidation activity of the

molecular catalyst with an essentially high initial TOF of 12.4 s-1

, but also

gives a surprisingly high initial catalytic current density of 5.6 mA cm-2

. This

work provides a methodology about: “How to employ catalysts in functional

devices without affecting their performances?”.

The other is the kinetic study of the O-O bond formation mechanistic pathways

on the electrode surface, and reaction orders and kinetic isotope effects were

examined. This work provides a methodology about: “How to get kinetic

information of RDS when a catalyst is immobilized on the electrode

surface?” Solid evidences of reaction mechanisms for poly-16@GC, poly-

16+PSt@GC and poly-17@GC electrodes were demonstrated. Not only the

ligand design can improve the activity of a catalyst, but also can the catalytic

environment such as spatial distance between catalytic centers also plays an

important rule for the catalyst‟s performance.

The second parts (Chapter 4 and 5) of this thesis describe our attempts of

applying molecular catalysts aforementioned in artificial photosynthesis

devices. Two aspects of this issue are concerned.

One is the design of photoanodes for light-driven water oxidation. First, poly-

16 and complex 18 were immobilized on RuP sensitized TiO2 films. Then,

KIE study of poly-16+RuP@TiO2 and 18+RuP@TiO2 suggests that (1) the

RDS for the photo-electrochemical reaction is the regeneration of dye; (2) the

decomposition of photosensitizer limits the stability of the whole PEC devices.

By replacing RuP@TiO2 with α-Fe2O3 as semiconductor, a stable photocurrent

was obtained. Furthermore, we employed a strong anchoring group PDC for

complex 19, and the corresponding electrode 19@Fe2O3 exhibits impressive

stability under harsh conditions (1 M KOH). This two works are corresponding

to the question: “How to improve the stability of photoanodes?”

The other is the dye-sensitized tandem PEC cells for light driven total water

splitting. In which, for the first time, a tandem molecular PEC cell

19+RuP@TiO2/20+RuP@NiO for water splitting driven by visible light was

82

demonstrated. Further by replacing the high cost Ru-dye with the organic dyes

L0 and P1, for the first time, a metal free organic dyes sensitized tandem PEC

cell that can split water by visible light under neutral pH conditions without

bias was obtained. Our results provide guidance about: “How to fabricate low-

cost Pt-free molecular PEC devices by employing the concept of Z-scheme?”

My research works have been focusing on the design and fabrication of water

splitting devices. Novel devices have been demonstrated with the view to

achieve the criteria of good long term stability, high efficiency as well as low

cost. Although none of my devices can be used in the large scale application

until now, my results broaden the understanding about molecular water

oxidation catalysts on electrode surface and provide guidance in the

construction of water splitting devices.

83

7. Future prospects

It only takes 32 years for artificial WOCs to reach a much better activity than

the OEC in PSII which was developed over millions years! Artificial HGCs

also can compete with nature hydrogenases now. But all these achievements

are just a start.

I strongly believe Solar Fuel Conversion by the reaction of water splitting has a

bright future, and for sure, this technology will provide a feasible and

sustainable energy system for human society one day.

Just like steam engine technology, electricity revolution, automobile industry

and smartphone, water splitting will change our life!

I'm glad I was in this fascinating transition, and I will continue.

Challenges are still there, and they all are related to cutting down the cost.

84

Acknowledgements

The work included in this thesis was accomplished primarily in the past four

and half years. During this period of time, I got a lot supports and

encouragements from many people around me. Without their help, none of the

pieces for this thesis would have been achieved. I would like to express my

sincere gratitude to:

First and foremost, my supervisor, Prof. Licheng Sun, for giving me the

opportunity to join the group and letting me work on this challenging but

interesting project, and also for your patience, encouragement, suggestion and

inspiration.

Assistant Prof. Lele Duan for your patient guidance, sharing your knowledge

and inspiring discussions.

Dr. Ke Fan for the help on experiments and fruitful collaborations.

All past and present members at Licheng‟s group.

Dr. Lin Li, Dr. Lianpeng Tong, Assistant Prof. Haining Tian, Dr. Erik

Gabrielsson and Dr. Martin Karlsson for giving me a fantastic start of my

PhD study.

Quentin Daniel, Dr. Lei Wang, Dr. Bo Xu, Dr. Ram Babruvan Ambre, Dr.

Hong Chen, Dr. Ming Cheng, Dr. Yong Hua, Dr. Biaobiao Zhang, Dr. Peili

Zhang, Linqing Wang and Lizhou Fan for the countless nice discussions,

collaborations.

Prof. Lars Kloo for teaching me the Swedish.

Prof. Torbjörn Norin and the whole Licheng‟s group for poof-reading my

thesis. I appreciate your time and effort!

Henry, Ulla and Zoltan for the pleasant working environment.

All past and present members at the department of Organic Chemistry.

The China Scholarship Council (CSC), the Swedish Research Council, the

Swedish Energy Agency, the Kunt and Alice Wallenberg Foundation and Aulin

Erdtmann foundation for financial supports.

My wife Wei Liu and my son Siyuan Li for your support, encouragement,

understanding and love.

85

Appendix A

The following is a description of my contribution to Publications I to VI, as

requested by KTH.

Paper I-VI: My contributions are the initiation of these projects, most of the

synthesis, measurements, kinetic studies as well as the writing of major part

for the manuscripts.

86

References

(1) Styring, S. Faraday Discuss. 2012, 155, 357.

(2) Osterloh, F. E.; Parkinson, B. A. MRS Bull. 2011, 36, 17.

(3) Blankenship, R. E. Molecular mechanisms of photosynthesis; John Wiley & Sons, 2013.

(4) Bard, A. J.; Fox, M. A. Acc. Chem. Res. 1995, 28, 141.

(5) Gratzel, M. Nature 2001, 414, 338.

(6) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q.; Santori, E. A.; Lewis,

N. S. Chem. Rev. 2010, 110, 6446.

(7) Tachibana, Y.; Vayssieres, L.; Durrant, J. R. Nat. Photon. 2012, 6, 511.

(8) Hammarström, L.; Hammes-Schiffer, S. Acc. Chem. Res. 2009, 42, 1859.

(9) Sun, L.; Hammarstrom, L.; Akermark, B.; Styring, S. Chem. Soc. Rev. 2001, 30, 36.

(10) Pinaud, B. A.; Benck, J. D.; Seitz, L. C.; Forman, A. J.; Chen, Z.; Deutsch, T. G.; James, B.

D.; Baum, K. N.; Baum, G. N.; Ardo, S.; Wang, H.; Miller, E.; Jaramillo, T. F. Energ Environ Sci

2013, 6, 1983.

(11) Kimoto, A.; Yamauchi, K.; Yoshida, M.; Masaoka, S.; Sakai, K. Chem Commun 2012, 48,

239.

(12) Luo, J.; Im, J.-H.; Mayer, M. T.; Schreier, M.; Nazeeruddin, M. K.; Park, N.-G.; Tilley, S. D.;

Fan, H. J.; Grätzel, M. Science 2014, 345, 1593.

(13) Cox, C. R.; Lee, J. Z.; Nocera, D. G.; Buonassisi, T. Proc. Natl. Acad. Sci. U.S.A. 2014, 111,

14057.

(14) Van de Krol, R.; Schoonman, J. In Sustainable Energy Technologies; Hanjalić, K., Van de

Krol, R., Lekić, A., Eds.; Springer Netherlands: 2008.

(15) van de Krol, R.; Liang, Y.; Schoonman, J. J. Mater. Chem. 2008, 18, 2311.

(16) Hisatomi, T.; Kubota, J.; Domen, K. Chem. Soc. Rev. 2014.

(17) Barber, J. Chem. Soc. Rev. 2009, 38, 185.

(18) Dau, H.; Zaharieva, I. Acc. Chem. Res. 2009, 42, 1861.

(19) Suga, M.; Akita, F.; Hirata, K.; Ueno, G.; Murakami, H.; Nakajima, Y.; Shimizu, T.;

Yamashita, K.; Yamamoto, M.; Ago, H.; Shen, J.-R. Nature 2015, 517, 99.

(20) Kärkäs, M. D.; Verho, O.; Johnston, E. V.; Åkermark, B. Chem. Rev. 2014, 114, 11863.

(21) Kanan, M. W.; Surendranath, Y.; Nocera, D. G. Chem. Soc. Rev. 2009, 38, 109.

(22) Trasatti, S. J.Electr.Chem.Interfa.Electrochem. 1980, 111, 125.

(23) Man, I. C.; Su, H.-Y.; Calle-Vallejo, F.; Hansen, H. A.; Martínez, J. I.; Inoglu, N. G.; Kitchin,

J.; Jaramillo, T. F.; Nørskov, J. K.; Rossmeisl, J. ChemCatChem 2011, 3, 1159.

(24) Trotochaud, L.; Ranney, J. K.; Williams, K. N.; Boettcher, S. W. J. Am. Chem. Soc. 2012,

134, 17253.

(25) Song, F.; Hu, X. Nat. Commun. 2014, 5, 4477.

(26) Dau, H.; Limberg, C.; Reier, T.; Risch, M.; Roggan, S.; Strasser, P. ChemCatChem 2010, 2,

724.

(27) Hirahara, M.; Nagai, S.; Takahashi, K.; Saito, K.; Yui, T.; Yagi, M. Inorg. Chem. 2015, 54,

7627.

(28) Wada, T.; Muckerman, J. T.; Fujita, E.; Tanaka, K. Dalton Trans. 2011, 40, 2225.

87

(29) Isobe, H.; Tanaka, K.; Shen, J.-R.; Yamaguchi, K. Inorg. Chem. 2014, 53, 3973.

(30) Clark, A. E.; Hurst, J. K. In Prog. Inorg. Chem.; John Wiley & Sons, Inc.: 2011.

(31) Romain, S.; Vigara, L.; Llobet, A. Acc. Chem. Res. 2009, 42, 1944.

(32) Concepcion, J. J.; Jurss, J. W.; Brennaman, M. K.; Hoertz, P. G.; Patrocinio, A. O. T.;

Murakami Iha, N. Y.; Templeton, J. L.; Meyer, T. J. Acc. Chem. Res. 2009, 42, 1954.

(33) Wada, T.; Tsuge, K.; Tanaka, K. Angew. Chem. Int. Ed. 2000, 39, 1479.

(34) Neudeck, S.; Maji, S.; López, I.; Meyer, S.; Meyer, F.; Llobet, A. J. Am. Chem. Soc. 2014,

136, 24.

(35) Zeng, Q.; Lewis, F. W.; Harwood, L. M.; Hartl, F. Coord. Chem. Rev. 2015, 304–305, 88.

(36) Gersten, S. W.; Samuels, G. J.; Meyer, T. J. J. Am. Chem. Soc. 1982, 104, 4029.

(37) Concepcion, J. J.; Jurss, J. W.; Templeton, J. L.; Meyer, T. J. J. Am. Chem. Soc. 2008, 130,

16462.

(38) Tseng, H.-W.; Zong, R.; Muckerman, J. T.; Thummel, R. Inorg. Chem. 2008, 47, 11763.

(39) Zong, R.; Thummel, R. P. J. Am. Chem. Soc. 2005, 127, 12802.

(40) Duan, L.; Xu, Y.; Gorlov, M.; Tong, L.; Andersson, S.; Sun, L. Chem. Eur. J. 2010, 16, 4659.

(41) Blakemore, J. D.; Crabtree, R. H.; Brudvig, G. W. Chem. Rev. 2015.

(42) Wang, L.; Duan, L.; Wang, Y.; Ahlquist, M. S. G.; Sun, L. Chem. Commun. 2014.

(43) Du, P.; Eisenberg, R. Energy Environ. Sci. 2012, 5, 6012.

(44) Poulsen, A. K.; Rompel, A.; McKenzie, C. J. Angew. Chem., Int. Ed. 2005, 44, 6916.

(45) Dismukes, G. C.; Brimblecombe, R.; Felton, G. A. N.; Pryadun, R. S.; Sheats, J. E.; Spiccia,

L.; Swiegers, G. F. Acc. Chem. Res. 2009, 42, 1935.

(46) Brimblecombe, R.; Swiegers, G. F.; Dismukes, G. C.; Spiccia, L. Angew. Chem., Int. Ed.

2008, 47, 7335.

(47) Ma, L.; Wang, Q.; Man, W.-L.; Kwong, H.-K.; Ko, C.-C.; Lau, T.-C. Angew. Chem., Int. Ed.

2015, 54, 5246.

(48) Lee, W.-T.; Muñoz, S. B.; Dickie, D. A.; Smith, J. M. Angew. Chem., Int. Ed. 2014, 53, 9856.

(49) Wasylenko, D. J.; Ganesamoorthy, C.; Borau-Garcia, J.; Berlinguette, C. P. Chem. Commun.

2011, 47, 4249.

(50) Dogutan, D. K.; McGuire, R.; Nocera, D. G. J. Am. Chem. Soc. 2011, 133, 9178.

(51) Fillol, J. L.; Codola, Z.; Garcia-Bosch, I.; Gomez, L.; Pla, J. J.; Costas, M. Nat. Chem. 2011,

3, 807.

(52) Ellis, W. C.; McDaniel, N. D.; Bernhard, S.; Collins, T. J. J. Am. Chem. Soc. 2010, 132,

10990.

(53) Zhang, M.-T.; Chen, Z.; Kang, P.; Meyer, T. J. J. Am. Chem. Soc. 2013, 135, 2048.

(54) Evans, D. J.; Pickett, C. J. Chem. Soc. Rev. 2003, 32, 268.

(55) Frey, M. ChemBioChem 2002, 3, 153.

(56) Armstrong, F. A. Curr. Opin. Chem. Biol. 2004, 8, 133.

(57) Steele, B. C. H.; Heinzel, A. Nature 2001, 414, 345.

(58) Debe, M. K. Nature 2012, 486, 43.

(59) Bockris, J. O. M.; Conway, B. E. J. Chem. Soc. Faraday Trans. 1949, 45, 989.

(60) Thoi, V. S.; Sun, Y.; Long, J. R.; Chang, C. J. Chem. Soc. Rev. 2013, 42, 2388.

(61) Cook, T. R.; Dogutan, D. K.; Reece, S. Y.; Surendranath, Y.; Teets, T. S.; Nocera, D. G.

Chem. Rev. 2010, 110, 6474.

88

(62) Chen, Y.; Tran, P. D.; Boix, P.; Ren, Y.; Chiam, S. Y.; Li, Z.; Fu, K.; Wong, L. H.; Barber, J.

ACS Nano 2015, 9, 3829.

(63) Merki, D.; Vrubel, H.; Rovelli, L.; Fierro, S.; Hu, X. Chem. Sci. 2012, 3, 2515.

(64) Tran, P. D.; Nguyen, M.; Pramana, S. S.; Bhattacharjee, A.; Chiam, S. Y.; Fize, J.; Field, M.

J.; Artero, V.; Wong, L. H.; Loo, J.; Barber, J. Energy Environ. Sci. 2012, 5, 8912.

(65) Artero, V.; Chavarot-Kerlidou, M.; Fontecave, M. Angew. Chem. Int. Ed. 2011, 50, 7238.

(66) Wang, M.; Chen, L.; Sun, L. Energy Environ. Sci. 2012, 5, 6763.

(67) Gloaguen, F.; Rauchfuss, T. B. Chem. Soc. Rev. 2009, 38, 100.

(68) Kilgore, U. J.; Roberts, J. A. S.; Pool, D. H.; Appel, A. M.; Stewart, M. P.; DuBois, M. R.;

Dougherty, W. G.; Kassel, W. S.; Bullock, R. M.; DuBois, D. L. J. Am. Chem. Soc. 2011, 133,

5861.

(69) Pool, D. H.; Stewart, M. P.; O‟Hagan, M.; Shaw, W. J.; Roberts, J. A. S.; Bullock, R. M.;

DuBois, D. L. Proceedings of the National Academy of Sciences 2012, 109, 15634.

(70) Baffert, C.; Artero, V.; Fontecave, M. Inorg. Chem. 2007, 46, 1817.

(71) Razavet, M.; Artero, V.; Fontecave, M. Inorg. Chem. 2005, 44, 4786.

(72) Hu, X.; Cossairt, B. M.; Brunschwig, B. S.; Lewis, N. S.; Peters, J. C. Chem. Commun. 2005,

4723.

(73) Hu, X.; Brunschwig, B. S.; Peters, J. C. J. Am. Chem. Soc. 2007, 129, 8988.

(74) Dempsey, J. L.; Brunschwig, B. S.; Winkler, J. R.; Gray, H. B. Acc. Chem. Res. 2009, 42,

1995.

(75) Hawecker, J.; Lehn, J. M.; Ziessel, R. New J. Chem. 1983, 7, 271.

(76) Boddy, P. J. J. Electrochem. Soc. 1968, 115, 199.

(77) Fujishima, A.; Honda, K. Nature 1972, 238, 37.

(78) Navarro Yerga, R. M.; Álvarez Galván, M. C.; Del Valle, F.; Villoria de la Mano, J. A.;

Fierro, J. L. G. ChemSusChem 2009, 2, 471.

(79) Swierk, J. R.; Méndez-Hernández, D. D.; McCool, N. S.; Liddell, P.; Terazono, Y.; Pahk, I.;

Tomlin, J. J.; Oster, N. V.; Moore, T. A.; Moore, A. L.; Gust, D.; Mallouk, T. E. Proceedings of

the National Academy of Sciences 2015, 112, 1681.

(80) Lindquist, R. J.; Phelan, B. T.; Reynal, A.; Margulies, E. A.; Shoer, L. E.; Durrant, J. R.;

Wasielewski, M. R. J. Mater. Chem. A 2016.

(81) Kirner, J. T.; Stracke, J. J.; Gregg, B. A.; Finke, R. G. ACS Appl. Mater. Interfaces 2014, 6,

13367.

(82) Palma, C.-A.; Samori, P. Nat Chem 2011, 3, 431.

(83) Barth, J. V.; Costantini, G.; Kern, K. Nature 2005, 437, 671.

(84) Ciesielski, A.; Palma, C.-A.; Bonini, M.; Samorì, P. Adv. Mater. 2010, 22, 3506.

(85) Tong, L.; Gothelid, M.; Sun, L. Chem. Commun. 2012, 48, 10025.

(86) Elgrishi, N.; Griveau, S.; Chambers, M. B.; Bedioui, F.; Fontecave, M. Chem. Commun. 2015,

51, 2995.

(87) Pinson, J.; Podvorica, F. Chem. Soc. Rev. 2005, 34, 429.

(88) Le Goff, A.; Artero, V.; Jousselme, B.; Tran, P. D.; Guillet, N.; Métayé, R.; Fihri, A.; Palacin,

S.; Fontecave, M. Science 2009, 326, 1384.

(89) Chen, Z.; Concepcion, J. J.; Jurss, J. W.; Meyer, T. J. J. Am. Chem. Soc. 2009, 131, 15580.

89

(90) Chen, Z.; Concepcion, J. J.; Luo, H.; Hull, J. F.; Paul, A.; Meyer, T. J. J. Am. Chem. Soc.

2010, 132, 17670.

(91) Liu, F.; Cardolaccia, T.; Hornstein, B. J.; Schoonover, J. R.; Meyer, T. J. J. Am. Chem. Soc.

2007, 129, 2446.

(92) Muresan, N. M.; Willkomm, J.; Mersch, D.; Vaynzof, Y.; Reisner, E. Angew. Chem. Int. Ed.

2012, 51, 12749.

(93) Queyriaux, N.; Kaeffer, N.; Morozan, A.; Chavarot-Kerlidou, M.; Artero, V. J. Photochem.

Photobiol. C: Photochem. Rev. 2015, 25, 90.

(94) Tran, P. D.; Le Goff, A.; Heidkamp, J.; Jousselme, B.; Guillet, N.; Palacin, S.; Dau, H.;

Fontecave, M.; Artero, V. Angew. Chem. Int. Ed. 2011, 50, 1371.

(95) Li, F.; Zhang, B.; Li, X.; Jiang, Y.; Chen, L.; Li, Y.; Sun, L. Angew. Chem. Int. Ed. Engl.

2011, 50, 12276.

(96) Shi, Y.; Peng, L.; Ding, Y.; Zhao, Y.; Yu, G. Chem. Soc. Rev. 2015, 44, 6684.

(97) Wang, L.; Fan, K.; Daniel, Q.; Duan, L.; Li, F.; Philippe, B.; Rensmo, H.; Chen, H.; Sun, J.;

Sun, L. Chem. Commun. 2015, 51, 7883.

(98) Ibrahim, S. K.; Liu, X.; Tard, C.; Pickett, C. J. Chem. Commun. 2007, 1535.

(99) Brimblecombe, R.; Koo, A.; Dismukes, G. C.; Swiegers, G. F.; Spiccia, L. J. Am. Chem. Soc.

2010, 132, 2892.

(100) Li, L.; Duan, L.; Xu, Y.; Gorlov, M.; Hagfeldt, A.; Sun, L. Chem. Commun. 2010, 46, 7307.

(101) Li, L.; Duan, L.; Wen, F.; Li, C.; Wang, M.; Hagfeldt, A.; Sun, L. Chem. Commun. 2012, 48,

988.

(102) Na, Y.; Hu, B.; Yang, Q.-L.; Liu, J.; Zhou, L.; Fan, R.-Q.; Yang, Y.-L. Chin. Chem. Lett.

2015, 26, 141.

(103) Gao, Y.; Ding, X.; Liu, J.; Wang, L.; Lu, Z.; Li, L.; Sun, L. J. Am. Chem. Soc. 2013, 135,

4219.

(104) Gao, Y.; Zhang, L.; Ding, X.; Sun, L. Phys. Chem. Chem. Phys. 2014, 16, 12008.

(105) Zhang, L.; Gao, Y.; Ding, X.; Yu, Z.; Sun, L. ChemSusChem 2014, 7, 2801.

(106) Gardner, J. M.; Beyler, M.; Karnahl, M.; Tschierlei, S.; Ott, S.; Hammarström, L. J. Am.

Chem. Soc. 2012, 134, 19322.

(107) Meng, P.; Wang, M.; Yang, Y.; Zhang, S.; Sun, L. J. Mater. Chem. A 2015, 3, 18852.

(108) Frischmann, P. D.; Mahata, K.; Wurthner, F. Chem. Soc. Rev. 2013, 42, 1847.

(109) Kaveevivitchai, N.; Chitta, R.; Zong, R.; El Ojaimi, M.; Thummel, R. P. J. Am. Chem. Soc.

2012, 134, 10721.

(110) Song, W.; Glasson, C. R. K.; Luo, H.; Hanson, K.; Brennaman, M. K.; Concepcion, J. J.;

Meyer, T. J. J. Phys. Chem. Lett. 2011, 2, 1808.

(111) Wang, L.; Ashford, D. L.; Thompson, D. W.; Meyer, T. J.; Papanikolas, J. M. J. Phys. Chem.

C 2013, 117, 24250.

(112) Bettis, S. E.; Ryan, D. M.; Gish, M. K.; Alibabaei, L.; Meyer, T. J.; Waters, M. L.;

Papanikolas, J. M. J. Phys. Chem. C 2014, 118, 6029.

(113) Alibabaei, L.; Brennaman, M. K.; Norris, M. R.; Kalanyan, B.; Song, W.; Losego, M. D.;

Concepcion, J. J.; Binstead, R. A.; Parsons, G. N.; Meyer, T. J. Proc. Natl. Acad. Sci. U.S.A. 2013,

110, 20008.

(114) Ji, Z.; He, M.; Huang, Z.; Ozkan, U.; Wu, Y. J. Am. Chem. Soc. 2013, 135, 11696.

90

(115) Chen, X.; Ren, X.; Liu, Z.; Zhuang, L.; Lu, J. Electrochem. Commun. 2013, 27, 148.

(116) Zhong, D. K.; Zhao, S.; Polyansky, D. E.; Fujita, E. J. Catal. 2013, 307, 140.

(117) Klepser, B. M.; Bartlett, B. M. J. Am. Chem. Soc. 2014, 136, 1694.

(118) Seo, J.; Pekarek, R. T.; Rose, M. J. Chem. Commun. 2015, 51, 13264.

(119) Krawicz, A.; Yang, J.; Anzenberg, E.; Yano, J.; Sharp, I. D.; Moore, G. F. J. Am. Chem. Soc.

2013, 135, 11861.

(120) de Respinis, M.; Joya, K. S.; De Groot, H. J. M.; D‟Souza, F.; Smith, W. A.; van de Krol, R.;

Dam, B. J. Phys. Chem. C 2015, 119, 7275.

(121) Cedeno, D.; Krawicz, A.; Doak, P.; Yu, M.; Neaton, J. B.; Moore, G. F. J. Phys. Chem. Lett.

2014, 5, 3222.

(122) Krawicz, A.; Cedeno, D.; Moore, G. F. Phys. Chem. Chem. Phys. 2014, 16, 15818.

(123) Duan, L.; Tong, L.; Xu, Y.; Sun, L. Energy Environ. Sci. 2011, 4, 3296.

(124) Tran, P. D.; Artero, V.; Fontecave, M. Energy Environ. Sci. 2010, 3, 727.

(125) Li, F.; Zhang, B.; Li, X.; Jiang, Y.; Chen, L.; Li, Y.; Sun, L. Angew. Chem. 2011, 123,

12484.

(126) Bard, A. J.; Faulkner, L. R. Electrochemical methods: fundamentals and applications; Wiley

New York, 1980; Vol. 2.

(127) Toma, F. M.; Sartorel, A.; Iurlo, M.; Carraro, M.; Parisse, P.; Maccato, C.; Rapino, S.;

Gonzalez, B. R.; Amenitsch, H.; Da Ros, T.; Casalis, L.; Goldoni, A.; Marcaccio, M.; Scorrano,

G.; Scoles, G.; Paolucci, F.; Prato, M.; Bonchio, M. Nat. Chem. 2010, 2, 826.

(128) Concepcion, J. J.; Jurss, J. W.; Hoertz, P. G.; Meyer, T. J. Angew. Chem. Int. Ed. 2009, 48,

9473.

(129) Duan, L.; Fischer, A.; Xu, Y.; Sun, L. J. Am. Chem. Soc. 2009, 131, 10397.

(130) Duan, L.; Araujo, C. M.; Ahlquist, M. S. G.; Sun, L. Proc. Natl. Acad. Sci. U.S.A. 2012, 109,

15584.

(131) Duan, L.; Bozoglian, F.; Mandal, S.; Stewart, B.; Privalov, T.; Llobet, A.; Sun, L. Nat.

Chem. 2012, 4, 418.

(132) Jiang, Y.; Li, F.; Zhang, B.; Li, X.; Wang, X.; Huang, F.; Sun, L. Angew. Chem. Int. Ed.

2013, 52, 3398.

(133) Tong, L.; Duan, L.; Xu, Y.; Privalov, T.; Sun , L. Angew. Chem. Int. Ed. 2011, 50, 445.

(134) Wang, L.; Duan, L.; Wang, Y.; Ahlquist, M. S. G.; Sun, L. Chem. Commun. 2014, 50,

12947.

(135) Lapides, A. M.; Ashford, D. L.; Hanson, K.; Torelli, D. A.; Templeton, J. L.; Meyer, T. J. J.

Am. Chem. Soc. 2013, 135, 15450.

(136) Fang, Z.; Keinan, S.; Alibabaei, L.; Luo, H.; Ito, A.; Meyer, T. J. Angew. Chem. 2014, 126,

4972.

(137) Ashford, D. L.; Sherman, B. D.; Binstead, R. A.; Templeton, J. L.; Meyer, T. J. Angew.

Chem. Int. Ed. 2015, 54, 4778.

(138) Li, F.; Fan, K.; Wang, L.; Daniel, Q.; Duan, L.; Sun, L. ACS Catal. 2015, 3786.

(139) Bamford, C. H.; Tipper, C. F. H.; Compton, R. G. Electrode Kinetics: Principles and

Methodology: Principles and Methodology; Elsevier, 1986.

(140) Huynh, M.; Bediako, D. K.; Nocera, D. G. J. Am. Chem. Soc. 2014, 136, 6002.

(141) Moonshiram, D.; Purohit, V.; Concepcion, J.; Meyer, T.; Pushkar, Y. Mater. 2013, 6, 392.

91

(142) Yamada, H.; Siems, W. F.; Koike, T.; Hurst, J. K. J. Am. Chem. Soc. 2004, 126, 9786.

(143) Liu, F.; Concepcion, J. J.; Jurss, J. W.; Cardolaccia, T.; Templeton, J. L.; Meyer, T. J. Inorg.

Chem. 2008, 47, 1727.

(144) Chen, Z.; Concepcion, J. J.; Hu, X.; Yang, W.; Hoertz, P. G.; Meyer, T. J. Proc. Natl. Acad.

Sci. U.S.A. 2010, 107, 7225.

(145) Krishtalik, L. I. Biochimica et Biophysica Acta (BBA) - Bioenergetics 2000, 1458, 6.

(146) Carey, F. A.; Sundberg, R. J. Advanced Organic Chemistry: Part A: Structure and

Mechanisms; Springer, 2007.

(147) Chen, Z.; Concepcion, J. J.; Hu, X.; Yang, W.; Hoertz, P. G.; Meyer, T. J. Proc Natl Acad

Sci U S A 2010, 107, 7225.

(148) Chen, Z.; Vannucci, A. K.; Concepcion, J. J.; Jurss, J. W.; Meyer, T. J. Proc. Natl. Acad. Sci.

U.S.A. 2011, 108, E1461.

(149) Matheu, R.; Francàs, L.; Chernev, P.; Ertem, M. Z.; Batista, V.; Haumann, M.; Sala, X.;

Llobet, A. ACS Catal. 2015, 5, 3422.

(150) Li, L.; Duan, L. L.; Xu, Y. H.; Gorlov, M.; Hagfeldt, A.; Sun, L. C. Chem. Commun. 2010,

46, 7307.

(151) Gao, Y.; Ding, X.; Liu, J. H.; Wang, L.; Lu, Z. K.; Li, L.; Sun, L. C. J. Am. Chem. Soc. 2013,

135, 4219.

(152) Bhasikuttan, A. C.; Suzuki, M.; Nakashima, S.; Okada, T. J. Am. Chem. Soc. 2002, 124,

8398.

(153) Lv, H.; Song, J.; Geletii, Y. V.; Vickers, J. W.; Sumliner, J. M.; Musaev, D. G.; Kögerler, P.;

Zhuk, P. F.; Bacsa, J.; Zhu, G.; Hill, C. L. J. Am. Chem. Soc. 2014, 136, 9268.

(154) Hanson, K.; Brennaman, M. K.; Luo, H.; Glasson, C. R. K.; Concepcion, J. J.; Song, W.;

Meyer, T. J. ACS Appl. Mater. Interfaces 2012, 4, 1462.

(155) Ding, X.; Gao, Y.; Zhang, L.; Yu, Z.; Liu, J.; Sun, L. ACS Catal. 2014, 4, 2347.

(156) Alibabaei, L.; Brennaman, M. K.; Norris, M. R.; Kalanyan, B.; Song, W. J.; Losego, M. D.;

Concepcion, J. J.; Binstead, R. A.; Parsons, G. N.; Meyer, T. J. P Natl Acad Sci USA 2013, 110,

20008.

(157) Gabrielsson, E.; Tian, H.; Eriksson, S. K.; Gao, J.; Chen, H.; Li, F.; Oscarsson, J.; Sun, J.;

Rensmo, H.; Kloo, L.; Hagfeldt, A.; Sun, L. Chem. Commun. 2015, 51, 3858.

(158) Rüttinger, W.; Dismukes, G. C. Chemical reviews 1997, 97, 1.

(159) Swierk, J. R.; Mallouk, T. E. Chemical Society Reviews 2013, 42, 2357.

(160) Li, L.; Duan, L. L.; Wen, F. Y.; Li, C.; Wang, M.; Hagfeld, A.; Sun, L. C. Chem. Commun.

2012, 48, 988.

(161) Tada, H.; Mitsui, T.; Kiyonaga, T.; Akita, T.; Tanaka, K. Nature materials 2006, 5, 782.

(162) Kato, H.; Hori, M.; Konta, R.; Shimodaira, Y.; Kudo, A. Chem. Lett. 2004, 33, 1348.

(163) Sayama, K.; Mukasa, K.; Abe, R.; Abe, Y.; Arakawa, H. J. Photochem. Photobiol. 2002,

148, 71.

(164) Ida, S.; Yamada, K.; Matsunaga, T.; Hagiwara, H.; Matsumoto, Y.; Ishihara, T. J. Am. Chem.

Soc. 2010, 132, 17343.

(165) Nattestad, A.; Mozer, A. J.; Fischer, M. K. R.; Cheng, Y. B.; Mishra, A.; Bauerle, P.; Bach,

U. Nat. Mater. 2010, 9, 31.

92

(166) Li, L.; Gibson, E. A.; Qin, P.; Boschloo, G.; Gorlov, M.; Hagfeldt, A.; Sun, L. C. Adv Mater

2010, 22, 1759.

(167) O'Regan, B.; Gratzel, M. Nature 1991, 353, 737.

(168) Fan, K.; Li, F.; Wang, L.; Daniel, Q.; Gabrielsson, E.; Sun, L. Phys. Chem. Chem. Phys.

2014, 16, 25234.

(169) Hisatomi, T.; Kubota, J.; Domen, K. Chem. Soc. Rev. 2014, 43, 7520.