DETERMINING THE ENERGETICS AND CHARGE DISTRIBUTION IN THE MESOPOROUS SEMICONDUCTOR PHOTOANODES USED IN DYE-SENSITIZED SOLAR CELLS By Dhritabrata Mandal A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Chemistry – Doctor of Philosophy 2017 ABSTRACT DETERMINING THE ENERGETICS AND CHARGE DISTRIBUTION IN THE MESOPOROUS SEMICONDUCTOR PHOTOANODES USED IN DYE-SENSITIZED SOLAR CELLS By Dhritabrata Mandal Mesoporous wide bandgap semiconductor electrodes have been widely studied for solar energy conversion, especially in the context of dye-sensitized solar cells (DSSCs) and related systems. The knowledge of their band edge energies, trap state distribution and electron density in various electronic states are essential in order to optimize the photoanode performance and to obtain a detailed understanding of the associated electron-transfer processes. This dissertation focuses on spectroelectrochemical and electrochemical analysis of the mesoporous phoanodes to determine the band edge position, energy distribution of the electronic states and also their spatial locations. Also presented here is a preparation method of mesoporous strontium titanate (SrTiO3) photoanode and a detailed study of its energetics and the charge distribution. The goal of this work is to explore the pros and cons of SrTiO3 material as a photoanode material in DSSC. Copyright by DHRITABRATA MANDAL 2017 To Maa, Baba, Chiku, Yindrila and my dear friends For their unconditional love and support iv ACKNOWLEDGEMENTS I would like to thank my PhD advisor Prof. Thomas Hamann. Without his adept advice, continued support and mentoring throughout my PhD career, this would not have been possible. His dedication for good science and his academic rigor in research have always motivated me to overcome the struggling period of my graduate research. I would also like to thank my committee members, Dr. Rémi Beaulac, Dr. Aaron Odom, and Dr. Benjamin G. Levine for their guidance and support throughout my PhD. Jesse, Yuling and Ben, thank you for your excellent mentoring and help during my graduate life. I would like to thank all the current members and former members of the Hamann group, Kelley, Rena, Suraj, Omid, Arianna, Josh, Dan, Yuan, Hamed, Yujue, Faezeh, Qiong, Austin and Parisa, for their support and for making my time here memorable. The graduation would be much difficult without your brilliant ideas that have enriched me and also without the fun that I had in my PhD life. So, thank you all. I had a great and enjoyable PhD life for all of you. I would also like to acknowledge U.S. Department of Energy for providing the funding to continue my research. Friends from my college days, I cannot thank you enough. I feel exceptionally lucky to have you during the “early” age of my life which shaped not only my career but also my way of life. Also, the friends I have made during my Spartan life have always been helpful in numerous ways and most importantly have offered me a great time in East Lansing. The unconditional love and support that I received from my Maa, Baba and Chiku are priceless. And finally, Yindrila. I have been blessed by your presence in my life. I cannot count the innumerable ways you have contributed to my personal and professional development and I eagerly look forward to what is to come. v TABLE OF CONTENTS LIST OF TABLES………………………………………………………………………...……viii LIST OF FIGURES………………………………………………………………………………ix KEY TO ABBREVIATIONS………………………………………………………………...…xiv Chapter 1 Introduction .................................................................................................................... 1 1.1 Motivation for research on Dye-sensitized Solar Cell (DSSC) ............................................ 1 1.2 DSSC operation and charge transfer processes .................................................................... 3 1.3 Historical development of DSSC .......................................................................................... 5 1.4 Literature review on photoanode materials used in DSSCs.................................................. 7 1.5 Motivation for studying photoanode properties in DSSC................................................... 10 1.6 Overview of the thesis ........................................................................................................ 11 REFERENCES ......................................................................................................................... 13 Chapter 2 Spectroelectrochemical Determination of the Conduction Band Edge Potential ........ 16 2.1 Introduction ......................................................................................................................... 16 2.2 Experimental ....................................................................................................................... 19 2.2.1 TiO2 photoanode preparation ....................................................................................... 19 2.2.2 Electrolyte preparation ................................................................................................. 20 2.2.3 Spectroelectrochemical measurement .......................................................................... 20 2.3 Results and discussion ........................................................................................................ 22 2.3.1 Absorption spectra of the TiO2 electrode..................................................................... 22 2.3.2 Determining band gap energy ...................................................................................... 23 2.3.3 Sources of uncertainty in determining band gap energy .............................................. 25 2.3.4 Burstein-Moss shift under potentiostatic bias .............................................................. 27 2.4 Conclusion: ......................................................................................................................... 33 REFERENCES ......................................................................................................................... 34 Chapter 3 Charge Distribution in Nanostructured TiO2 Photoanode Determined by Quantitative Analysis of the Band Edge Unpinning ......................................................................................... 38 3.1 Introduction ......................................................................................................................... 38 3.2 Electrochemical techniques for determining charge distribution ....................................... 41 3.2.1 Cyclic voltammetry...................................................................................................... 42 3.2.2 Electrochemical impedance spectroscopy ................................................................... 44 3.2.3 Charge insertion/extraction method ............................................................................. 46 3.3 Experimental ....................................................................................................................... 47 3.3.1 TiO2 photoanode preparation ....................................................................................... 47 3.3.2 Electrolyte preparation ................................................................................................. 48 3.3.3 Measurements .............................................................................................................. 48 3.4 Results and discussions ....................................................................................................... 50 3.4.1 Accurate energy distribution of trap states in mesoporous TiO2 photoanode ............. 50 vi 3.4.2 Spatial location of the trap states ................................................................................. 55 3.5 Conclusion .......................................................................................................................... 59 REFERENCES ......................................................................................................................... 61 Chapter 4 Absorbance Features of Mesoporous TiO2 Electrode under Electron Accumulation .. 65 4.1 Introduction ......................................................................................................................... 65 4.2 Experimental ....................................................................................................................... 67 4.2.1 TiO2 photoanode preparation ....................................................................................... 67 4.2.2 Electrolyte preparation ................................................................................................. 68 4.2.3 Spectroelectrochemical measurement .......................................................................... 68 4.3 Results and discussion ........................................................................................................ 70 4.4 Conclusion .......................................................................................................................... 80 REFERENCES ......................................................................................................................... 81 Chapter 5 Conduction Band Energy and Charge Distribution in Strontium Titanate Nanoparticle Electrode ....................................................................................................................................... 84 5.1 Introduction ......................................................................................................................... 84 5.2 Experimental ....................................................................................................................... 88 5.2.1 Preparation of SrTiO3 electrode ................................................................................... 88 5.2.2 Characterization of SrTiO3 film ................................................................................... 90 5.2.3 Preparation of electrolyte solutions ............................................................................. 90 5.2.4 Spectroelectrochemical measurement .......................................................................... 90 5.2.5 Electrochemical measurements .................................................................................... 92 5.3 Results and Discussion ....................................................................................................... 92 5.3.1 Material characterization ............................................................................................. 92 5.3.2 Determination of the conduction band and electron distribution................................. 94 5.4 Conclusion ........................................................................................................................ 114 REFERENCES ....................................................................................................................... 116 Chapter 6 Predicting Maximum Attainable Efficiency of DSSC ............................................... 121 6.1 Introduction ....................................................................................................................... 121 6.2 Calculating the power conversion efficiency of DSSC .................................................... 122 6.3 Energy losses in a DSSC................................................................................................... 122 6.3.1 Energy loss during photoexcited electron injection, Vloss, injection ............................... 125 6.3.2 Energy loss during regeneration of the oxidized dye, Vloss, regeneration ......................... 129 6.3.3 Recombinations losses, Vloss, recombination ..................................................................... 131 6.4 Predicting the maximum power conversion efficiency in DSSCs: ................................... 134 6.5 Conclusion ........................................................................................................................ 140 REFERENCES ....................................................................................................................... 142 Chapter 7 Conclusions and future directions .............................................................................. 147 REFERENCES ....................................................................................................................... 150 vii LIST OF TABLES Table 2-1 Conduction band positions were calculated at various applied potentials from the magnitudes of the band gap shift. EG was calculated from the Tauc plots at the applied potentials mentioned below and EG = EG (negative bias) – EG (+0.445 V vs. SCE). The electrolyte solution is 0.2 M KCl at pH 12.8. ................................................................................................. 32 Table 4-1 Values of P in presence of different electrolyte solutions. The absorption spectra of TiO2 electrode in contact with different electrolyte solution were measured and fitted to Drude equation, A = CP. Global fit of several spectra measured in an electrolyte condition provided the value of P. ............................................................................................................................... 78 Table 5-1 Calculation of conduction band position from the Burstein-Moss shift. ................... 101 Table 6-1 Table of the photoconversion efficiencies at different absorbance onsets and the associated loss in potentials. ....................................................................................................... 137 viii LIST OF FIGURES Figure 1-1 Schematic of a dye-sensitized solar cell, DSSC (top) and energy diagram of a DSSC showing major processes (bottom). Negative potential represents higher energy. The mesoporous TiO2 photoanode is deposited on transparent conductive substrate (TCO). The conduction band (CB) edge potential of TiO2 photoanode is shown below the excited state potential of dye. EF is the Fermi level potential of electrons in the photoanode. R+/R represents the redox shuttles in the solution which regenerate the oxidized dye. VOC is the open circuit potential which is the potential difference between the solution potential and the Fermi level at the open circuit, EF,Voc. The charge transfer processes shown in the diagram are described below. .................................... 4 Figure 2-1 SEM image of mesoporous TiO2 photoanodes used in DSSC. The image on the right shows the cross section of a TiO2 film.......................................................................................... 17 Figure 2-2 Instrumental set-up of spectroelectrochemical measurement. ................................... 20 Figure 2-3 Change in absorbance of a TiO2 electrode in contact with a pH 12.8 aqueous electrolyte when potential is stepped from + 0.455 V to  1.245 V (red),  1.295 V (green),  1.345 V (yellow),  1.395 V (blue) and  1.445 V (pink) vs. SCE (from bottom to top). ........... 23 Figure 2-4 Top: Plot of absorption coefficient () vs. wavelength () of TiO2 film. Bottom: Tauc plot of a TiO2 film. It was measured under an external positive bias + 0.445 V vs. SCE to discharge any stored electrons from the nanoparticles. The electrolyte solution is 0.2 M KCl at pH 12.8.......................................................................................................................................... 24 Figure 2-5 Energy vs. crystal momentum diagram (E-k) of a semiconductor with an indirect band gap showing the Burstein-Moss shift due to the filling of the lowest energy states in the CB under negative (reductive) potential. Electron transition from the highest-energy state in the valence band to the lowest-energy state in the conduction band is not possible without a change in momentum. The transition occurs via the absorption or emission of a phonon (red horizontal arrow), which provides the required momentum change during the transition. The energy required for this transition mainly comes from the photon absorbed (violet vertical arrow). ...... 28 Figure 2-6 Plot of absorption coefficient () vs. wavelength () of TiO2 film showing blue shift of the band gap absorption energy under negative (reductive) potential. The electrolyte solution is 0.2 M KCl at pH 12.8. ............................................................................................................... 29 Figure 2-7 Band gap widening occurs while moving from +0.445 V (red) towards more negative potentials -1.245 V (green), -1.345 V (yellow), -1.445 V (blue) vs. SCE. The electrolyte solution is 0.2 M KCl at pH 12.8. ............................................................................................................... 29 Figure 2-8 Plot of Burstein-Moss shifts (EG) at various applied potentials. ............................. 30 Figure 3-1 (Left) Energy vs. crystal momentum diagram of a semiconductor with an indirect band gap. Electron transition from the highest-energy state in the valence band to the lowest- ix energy state in the conduction band is not possible without a change in momentum. The transition occurs via the absorption or emission of a phonon, which provides the required momentum change during the transition. Inter-band localized electronic states bellow the conduction band, which trap electrons, are shown. (Right) Depiction of the upward shift of the energetics in TiO2 nanoparticle film. ECB represents the magnitude of the upward shift of the conduction band energy. ............................................................................................................... 40 Figure 3-2. Cyclic voltammogram plot (left) and chemical capacitance plot (right) of TiO2 film in aq. 0.2 M KCl solution at pH 12.8. Scan rate was 50 mV / s. The chemical capacitance (C) was calculated using the relation J = l(1-p)sC, where J is the current density, l (= 6.95 m) is the electrode thickness, p (= 0.71) is the porosity of the TiO2 film and s is the scan rate. The peak at approximately  0.75 V vs. SCE is referred to as monoenergetic trap state distribution and the wave at increasing negative potentials corresponds to the exponential distribution of trap states. ....................................................................................................................................................... 43 Figure 3-3 Equivalent circuit based on transmission line model for fitting Electrochemical Impedance data of porous photoanode. The fit parameters are described in the text. .................. 45 Figure 3-4 Plot of capacitance (C) of TiO2 film measured by electrochemical impedance spectroscopy. Electrolyte was 0.2 M aq. KCl solution at pH 12.8. .............................................. 45 Figure 3-5 Typical current vs. time plot measured during charge insertion/extraction process under potentiostatic. The measurement was performed with mesoporous TiO2 film in contact with aq. 0.2 M KCl solution at pH 12.8. The reductive bias was  1.045 V and the oxidative bias was + 0.455 V vs. SCE. ................................................................................................................ 47 Figure 3-6. The position of the conduction band edge potential determined at pH 12.8 and 2.0. 51 Figure 3-7. Plot of the chemical capacitance (C) as a function of applied potential, E, (bottom axis, red circles) and applied potential which is corrected to include the effect of band edge unpinning (ECB/q) as shown in equation 7 (top axis, blue triangles). ........................................ 52 Figure 3-8. Plot of absorbance of TiO2 at 780 nm as a function of applied potential (bottom axis, red circles) and applied potential which is corrected to include the effect of band edge unpinning (top axis, blue triangles). ............................................................................................................... 54 Figure 3-9. Plot of the concentration of trapped electrons in the bulk (nB) and on the surface (nS) of TiO2 film in aq. 0.2 M KCl solution at pH 12.8. Extrapolation of the fitted plots to the conduction band edge gives the density of bulk trap states, NB, and surface traps, NS. ............... 56 Figure 3-10. Plot of electron density in the nanostructured TiO2 photoanode as a function of applied potential. The total trapped electron density (nTrap) is the sum of the electrons trapped on the surface (nS) and in the bulk (nB).............................................................................................. 56 Figure 4-1 Transmittance (%T) spectra of TiO2 photoanode in contact with aqueous 0.2 M KCl solution at pH 12.8. Potentials are given against SCE. ................................................................. 70 x Figure 4-2 Change in absorbance of a TiO2 electrode under reductive bias. (a) TiO2 film is in contact with a pH 12.8 aqueous electrolyte when potential is stepped from + 0.455 V to  1.245 V (red),  1.295 V (green),  1.345 V (yellow),  1.395 V (blue) and  1.445 V (pink) vs.SCE (from bottom to top). (b) TiO2 film is in contact with a pH 2.0 aqueous electrolyte, potential is stepped from + 0.955 V to  0.495 V (red),  0.545 V (green),  0.595 V (yellow),  0.645 V (blue) and  0.695 V (pink) ,  0.745 (sky blue) vs.SCE (from bottom to top). .......................... 71 Figure 4-3 Global fit results (black lines) of the absorbance spectra from Fig. 4-2a,b according to Drude equation. (a) TiO2 film is in contact with a pH 12.8 aqueous electrolyte when potential is stepped from + 0.455 V to  1.245 V (red),  1.295 V (green),  1.345 V (yellow),  1.395 V (blue) and  1.445 V (pink) vs.SCE (from bottom to top). (b) TiO2 film is in contact with a pH 2.0 aqueous electrolyte, potential is stepped from + 0.955 V to  0.495 V (red),  0.545 V (green),  0.595 V (yellow),  0.645 V (blue) and  0.695 V (pink) ,  0.745 (sky blue) vs.SCE (from bottom to top)...................................................................................................................... 72 Figure 4-4 Global fit results (black lines) of the absorption spectra of TiO2 film according to Drude equation. Electrolyte is 0.2 M KCl at pH 6.2 adjusted by Citrate buffer (top) and Phosphate buffer (bottom). Potential shown in the plot is given vs. SCE. ................................... 73 Figure 4-5 Global fit results (black lines) of the absorption spectra of TiO2 film according to Drude equation. Electrolyte is 0.2 M KCl at pH 7.8 adjusted by Phosphate buffer (top) and Imidazole buffer (bottom). Potential shown in the plot is given vs. SCE. ................................... 74 Figure 4-6 Global fit results (black lines) of the absorption spectra of TiO2 film according to Drude equation. Electrolytes are 0.2 M LiClO4 at pH 7.8 (top) and ~ aqueous 0.05 M NaPF6 solution (bottom). The pH of the LiClO4 solution was adjusted at 7.8 by Imidazole buffer. No buffer was added into aq. NaPF6 solution. Potential shown in the plot is given vs. SCE. ........... 75 Figure 4-7 Global fit results (black lines) of the absorption spectra of TiO2 film according to Drude equation. Electrolytes are 0.2 M LiClO4 (top) and TBAPF6 (bottom) in acetonitrile. Potential shown in the plot is given vs. SCE. ............................................................................... 76 Figure 4-8 Density of electrons in conduction band (nCB) is calculated from Burstein-Moss shift, described in Chapter 2, and absorbance (A780 nm) of TiO2 film is measured at 780 nm under different applied potentials mentioned in the plot. Electrolyte was 0.2 M KCl at pH 12.8 and the potentials are given vs. SCE. ........................................................................................................ 80 Figure 5-1 Crystal structures of Perovskite SrTiO3 and anatase TiO2.. ....................................... 85 Figure 5-2 Apparatus setup used for the synthesis of SrTiO3 nanocrystals via vapor diffusion sol-gel method. .............................................................................................................................. 88 Figure 5-3 Powder X-ray diffraction patterns of the SrTiO3 nanocrystals (top) and scanning electron micrographs SrTiO3 film deposited on conductive substrate (bottom)........................... 93 Figure 5-4 Plot of absorbance of SrTiO3 electrode in UV region under various applied potentials, 0.0 V (black), -1.3 V (blue), -1.4 V (red), -1.5 V (green) vs. Ag/AgNO3. The electrolyte solution xi is 0.1 M LiClO4 in acetonitrile. Increasing negative potential moves the band gap absorption edge to higher energy. ................................................................................................................... 95 Figure 5-5 Plot of visible-near IR absorbance spectra of SrTiO3 electrode under various applied potentials, 0.0 V (black), -1.3 V (blue), -1.4 V (red), -1.5 V (green) vs. Ag/AgNO3. .................. 96 Figure 5-6 Plot of absorbance difference spectra of SrTiO3 electrode under various applied potentials, -1.3 V (blue), -1.4 V (red), -1.5 V (green). This difference absorbance was calculated by subtracting the absorbance of SrTiO3 electrode at 0.0V vs. Ag/AgNO3 from the absorbance measured at negative potentials shown in the plot........................................................................ 96 Figure 5-7 Tauc plot of a SrTiO3 film in contact with 0.1 M LiClO4 in acetonitrile. Any contribution from the electrolyte, substrate and cuvette were subtracted and only the absorbance of the SrTiO3 electrode contributed to this plot. ........................................................................... 99 Figure 5-8 Plot of band gap widening while moving from 0.0 V (black) towards more negative potentials -1.3 V (blue), -1.4 V (red), -1.5 V (green) vs. Ag/AgNO3. .......................................... 99 Figure 5-9 Current vs. time plot measured during charge insertion-extraction process under various potentiostatic biases. A reductive bias (as mentioned in the plot) was applied for 3 min and then an oxidative bias (0.0 V vs. AgNO3) was applied to extract all the charges stored in SrTiO3 electrode.......................................................................................................................... 103 Figure 5-10 Plot of electron density in the nanostructured SrTiO3 electrode as a function of applied potential measured by charge insertion-extraction measurement. ................................. 103 Figure 5-11 Plot of electron density in SrTiO3 and TiO2 electrode under various applied biases calculated by charge extraction method. .................................................................................... 107 Figure 5-12 Plot of measured capacitance, C (red) and resistance, R (blue) of SrTiO3 electrode measured by electrochemical impedance spectroscopy. ............................................................. 107 Figure 5-13 Comparison of capacitance of SrTiO3 electrode measured by the electrochemical impedance spectroscopy and charge insertion-extraction method.............................................. 108 Figure 5-14 Plot of electron density in the nanostructured SrTiO3 electrode as a function of applied potential. nCB is density of free electrons in the conduction band, nS, and nBulk are density of electrons trapped on the surface and in the bulk respectively. ............................................... 108 Figure 5-15 (a,b) Current transient data of when applied bias was changed from 0.0 V to the negative potentials mentioned in the plot. The SrTiO3 electrode is in contact with 0.1 M LiClO4/Acetonitrile solution. ...................................................................................................... 111 Figure 5-16 Triexponential fit of the current transient data when applied bias was changed from 0.0 V to -1.5 V vs. AgNO3. The experimentally observed decay has the contribution from three decays (1, 2 and 3 in the plot) shown here. ................................................................................. 112 xii Figure 5-17 Plot of time constants derived from the triexponential fit of the experimentally observed current decay data. ....................................................................................................... 113 Figure 5-18 Comparison of the CB and electron distribution between TiO2 and SrTiO3. ........ 115 Figure 6-1 J–V curve and IPCE plot of the DSSC photosensitized collaboratively by ADEKA-1 and LEG4. [Co(phen)3]2+/3+ redox shuttle was used. JV was measured AM-1.5, 100 mW cm2. 123 Figure 6-2 Plot of major energy losses in a typical DSSC. Negative potential represents higher energy. The mesoporous TiO2 photoanode is deposited on transparent conductive substrate (TCO). The conduction band (CB) edge potential of TiO2 photoanode is shown below the excited state potential of dye. EF is the Fermi level potential of electrons in the photoanode. R+/R represents the redox shuttles in the solution which regenerate the oxidized dye. VOC is the open circuit potential which is the potential difference between the solution potential and the Fermi level at the open circuit, EF,Voc. The loss-in-potentials, Vloss, are described in this chapter in details. ..................................................................................................................................................... 124 Figure 6-3 Schematic of back-electron transfer (i.e. recombination) processes from the semiconductor nanoparticles and the transparent conductive substrate (TCO) into the redox shuttle (R+/R) in the electrolyte .................................................................................................. 131 Figure 6-4 Energy diagram displaying the distribution of electrons in the TiO2 photoanode and the recombination processes from the conduction band (green) and from the trap states (red). 132 Figure 6-5 Photon flux of the AM 1.5 G spectrum at 100 mW cm-2 (ASTM G173-03). .......... 134 Figure 6-6 Short-circuit photocurrent density as a function of absorption onset calculated by integrating the AM 1.5 solar spectrum. ...................................................................................... 135 Figure 6-7 Photoconversion efficiency of DSSC as a function of loss-in-potential at different absorption onsets of the sensitized photoanode .......................................................................... 138 xiii KEY TO ABBREVIATIONS DSSC: Dye-sensitized Solar Cell J: Joule TW: Terawatt CO2: Carbon dioxide ppm: parts-per-million CH4: Methane N2O: Nitrous oxide PV: Photovoltaic : Photoconversion efficiency e: Electron Dye*: Excited state of the dye molecule Red: Reduced form Ox: Oxidized form TCO: Transparent conductive oxide EF: Fermi level energy EF: Fermi level potential ECB: Energy of the conduction band edge ECB: Conduction band edge potential VOC: Open circuit voltage CB: Conduction band VB: Valence band Ti: Titanium O: Oxygen xiv Sr: Strontium Zn: Zinc Sn: Tin Zr: Zirconium Nb: Niobium Si: Silicon LUMO: Lowest unoccupied molecular orbital eV: Electronvolt OH: Hydroxyl group TiO2: Titanium oxide ZnO: Zinc oxide SrTiO3: Strontium titanate SrSnO3: Strontium stannate Zn2SnO4: Zinc stannate CdSnO3: Cadmium stannate BaSnO3: Barium stannate CB: Extinction co-efficient of the conduction band electrons ND: Dopant density NP: Nanoparticle IR: Infrared nm: nanometer m: Micrometer VTSEC: Variable temperature spectroelectrochemistry Vis: Visible NIR: Near infrared xv FTO: Fluorine doped tin oxide p: porosity mm: Millimeter g: Gram cm: Centimeter M: Megaohms HCl: Hydrogen chloride KOH: Potassium hydroxide %T: Transmittance UV: Ultraviolet Ag: Silver AgCl: Silver chloride A: Absorbance A: Difference absorbance : Absorption coefficient l: Thickness of the semiconductor film q: charge of an electron nCB: Density of electrons in the conduction band NCB: Density of states in the conduction band kb: Boltzmann constant T: Absolute temperature V: Volt : Wavelength of light a.u. : Arbitrary unit h: Plank’s constant xvi : Frequency of light k: Crystal momentum EG: Band gap energy EP: Phonon energy B-M shift: Burstein-Moss shift m*: Reduced mass of charge carrier mh*: Effective mass of electron me*: Effective mass of hole me: Mass of an electron meV: Millielectronvolt SCE: Saturated calomel electrode d: Diameter of a nanoparticle DFT: Density functional theory CV: Cyclic voltammetry EIS: Electrochemical impedance spectroscopy C: Capacitance C: Chemical capacitance Q: Charge s: Scan rate of a cyclic voltammetric scan DOS: Density of states KCl: Potassium chloride Rs: Series resistance Rt: Transport resistance Rr: Interfacial resistance RPt: Resistance at platinum counter electrode xvii Aq.: Aqueous i: current t: time H2: Hydrogen Hz: Hertz N2: Nitrogen K: Kelvin MCB: Molar concentration of conduction band electron MT: Molar concentration of trapped electron CH: Helmholtz layer capacitance CBE: Conduction band electron p: porosity min: Minute ALD: Atomic layer deposition s: second AgNO3: Silver nitrate Pt: Platinum XRD: X-ray Diffraction Pmax: Maximum output power Pin: Power of the incident light JSC: Short circuit current ff: Fill factor J: Current density V: Voltage IPCE: Incident photon to current conversion efficiency xviii AM: Air mass mW: Milliwatt xix Chapter 1 Introduction 1.1 Motivation for research on Dye-sensitized Solar Cell (DSSC) The ever increasing demand for energy is a major driving force behind the extensive research on carbon-neutral energy sources. In 2012, the worldwide energy demand was 5531018 J/year which is equivalent to an average power consumption of 17.5 TW and in 2100, the average power consumption is projected to be 43 TW.1,2 Fossil sources like natural gas, oil, coal have been estimated to provide more than 80% of the global total primary energy supply.3,4 However, the anthropogenic carbon dioxide (CO2) released during the combustion of fossil fuel is strongly correlated to global warming. CO2 concentration in the atmosphere has drastically increased in the last century due to the industrialization as well as the other human activities relying on continuous source of energies. In mid 1800s, the average growth of CO2 concentration was 280 ppm which has become 399 ppm in 2015 (~ 40% increase in less than 200 years). During this time the other human activities e.g. agriculture, domestic livestock, biomass burning, waste etc., which produce greenhouse gases like CO2, CH4, N2O, fluorinated gases, have also increased. However, the CO2 gas released to meet the huge energy demand shares about 60% of the global anthropogenic greenhouse gases.5 Moreover, the uneven geographical distribution of these fossil fuel resources has resulted in several political and economical issues. The solution lies in moving towards renewable energy resources like solar and wind which are abundant in nature and potentially cost-effective. Solar energy arguably has the highest potential to meet the growing energy demand. The sun irradiates our earth with around 1.2×105 TW of power. Covering less than 0.4% of our planet's surface with 15% efficient solar panels would meet this demand.6 However, photovoltaic (PV) devices have remained very far from attaining a 1 significant share of the energy production market, primarily, because of their production cost which is mainly associated with the extreme purity requirement for the active materials, e.g crystalline silicon PV device requires ~99.9999% purity.7 Additionally, Silicon PV does not perform well under cloudy or shaded regions and their opacity hinders them from being an attractive choice as Building-integrated photovoltaics. Dye-sensitized Solar Cells (DSSCs) have many advantages compared to conventional PV.8,9 They can be manufactured from earth abundant and cheap materials through cost effective and high throughput manufacturing methods such as reel-to-reel printing.10 They offer flexibility in shape,11 transparency and even with the color which makes them a potential candidate for solar windows.12 Because of the rough materials surface, DSSCs perform much better than the conventional PV devices under diffused light and low light intensity (cloudy day) which has expanded their versatile use in indoor applications. The major drawbacks holding DSSCs back from a competitive entry into the PV market are their stability issues and comparatively low efficiency in large modules.13 A promising laboratory record efficiency of 14.3% has been achieved so far.14 However, efficiency of the large modules is much lower and a photoconversion efficiency of 5-8% is considered to be good.15,16 The photoconversion efficiencies of Si-based photovoltaic modules generally range from 14 to 18%, much higher than DSSC modules. Making the photoconversion efficiency higher will further reduce the overall electricity production cost which would make DSSC technology more competitive in the energy supply market. DSSC has several components and interfaces and the photoconversion efficiency is hugely challenged by various energy loss mechanisms. Even after more than 20 years of research, the system has not been well understood and predicting a roadmap towards the best DSSC, in terms of photoconversion efficiency (), is still not feasible. Altering a single component from the optimized combination affect the other 2 electron transfer processes which in turn necessitates change in other components.17 This challenge will only be met through extending our understanding of the interfacial charge transfer processes and mastering our ability of systematic molecular engineering of each of the components in the dye-sensitized solar cell. 1.2 DSSC operation and charge transfer processes The main components of DSSCs are nanostructured semiconductor anode, monolayer of dye attached to on the surface of this anode, redox shuttle and counter electrode (cathode). The charge separation takes place at the dye-semiconductor interface. Separation between the solution potential, E(R+/R) and the Fermi level of electrons in the semiconductor, EF, is the maximum voltage output from a cell. At the open circuit, this separation is the highest and the corresponding voltage is called open circuit voltage, VOC. Due to the presence of various components and interfaces, there are many electron transfer processes that occur in a DSSC as briefly described below. The numbers denotes the processes as shown in the DSSC energy diagram, Figure 1-1. 1) Under illumination, the dye absorbs light and an electron is excited. 1) Relaxation of the excited dye directly into its ground state may occur before the injection process takes place. 2) The photoexcited electron is injected into the conduction band of the adjacent semiconductor nanoparticle. 3 Figure 1-1 Schematic of a dye-sensitized solar cell, DSSC (top) and energy diagram of a DSSC showing major processes (bottom). Negative potential represents higher energy. The mesoporous TiO2 photoanode is deposited on transparent conductive substrate (TCO). The conduction band (CB) edge potential of TiO2 photoanode is shown below the excited state potential of dye. EF is the Fermi level potential of electrons in the photoanode. R+/R represents the redox shuttles in the solution which regenerate the oxidized dye. VOC is the open circuit potential which is the potential difference between the solution potential and the Fermi level at the open circuit, EF,Voc. The charge transfer processes shown in the diagram are described below. 4 3) The electrons injected into the semiconductor nanostructure are transported through the mesoporous network to the conductive substrate where it is collected by the external circuit. Generally transparent conductive oxides (TCO) are used as the substrate in DSSCs. 4,5) The hole produced in the dye is transported to the cathode by the redox shuttle or hole collector and thus the dye is regenerated to participate into the next photoexcitation cycle. 6,7) Recombination of the electrons from the conduction band of semiconductor photoanode into the oxidized form of the redox shuttle and dye. 8,9) Mesoporous semiconductor photoanodes are known to contain a large density of trap states and trap state mediated recombination processes also take place. In addition to the above mentioned components and charge transfer processes, a variety of additives and blocking layers are often used to make the performance better. This also adds some additional processes in the DSSC. 1.3 Historical development of DSSC Sensitization of semiconductor to modulate its absorbance is over a century old technology used in photography.18 In 1938, Gurney and Mott proposed the electron transfer from a dye excited state into the semiconductor conduction band which is the central idea of the dyesensitization.19 Then, electron generation by sensitizing a wide band gap semiconductor was reported by Putseiko and Trenin in 1949 who were studying photoconductivity of dyes adsorbed ZnO under illumination.20 The seminal work by Gerischer and Tributsch in 1968 made a breakthrough in the understanding of producing current by sensitized semiconductor under illumination.21,22 Using an electrochemical cell constructed with single wide band gap ZnO and a solution of fluorescein, rose bengale, or pseudoisocyanine chloride, they observed “Sensitized 5 photo-currents” when irradiated with the light of a wavelength which can only be absorbed by the dyes. This is the fundamental idea of a dye sensitized solar cell. A later breakthrough came in 1976 when Osa et al. reported a system where the dye is attached to the semiconductor, hence greatly improving the quantum yield of the electron injection when compared to the previous system where the dye is dissolved in the solution and a majority of the photoexcited dye decays before it can inject an electron into the semiconductor.23 However, the attempts to make dyesensitized photoelectrochemical cells were severely hindered by the fact that only a single molecular layer of dye on a flat semiconductor can inject an electron into the semiconductor. The light harvesting efficiency of one layer of dye adsorbed on a flat semiconductor crystal was extremely small and far from producing effective photoconversion. Attempts were made to increase the internal surface area of semiconductor electrode to enhance the dye adsorption and hence improve the light harvesting efficiency.24,25 The breakthrough in modern DSSC technology was made by O’Regan and Grӓtzel in 1991 where a mesoporous semiconductor electrode was used to increase the internal surface area two orders of magnitude higher than the geometric area.8 It greatly enhanced the amount of dyes loaded in the path of the incident light. After a few optimizations with the HOMO-LUMO gap of the dye and electrolyte (redox shuttles and additives) an excellent photoconversion efficiency of 10% was achieved in 1993.26 This breakthrough demonstrated the potential of DSSC renewable electricity production and attracted enormous attention worldwide. Within a few years, engineering the absorption onset of the sensitizers lead to the photocurrent production beyond 20 mAcm-2 under 1 sun illumination.27,28 However, the overall photoconversion efficiency remained stagnant at around 10-11 % for almost 18 years.29 A new breakthrough came in 2011 when Yella and Gratzel pushed the efficiency to 12.3%. This report is very crucial not only because its record efficiency but also in 6 the context of systematic engineering of dye and redox shuttle. Iodide/triiodide redox shuttle was used in all of the champion DSSCs starting from the beginning in 1991 which was replaced by cobalt trispyridyl redox shuttles in their 2011 report. Additionally, ruthenium-based sensitizer had been an essential part of the champion DSSCs throughout the history which was replaced in their report by a successful co-sensitization of donor-π-bridge-acceptor zinc porphyrin and organic D-π-A dye (Y123). The efficiency of similar DSSC configuration has been further pushed to 13% in 2013 which has remained the record till today.14 Apart from the quest for improving photoconversion efficiency, DSSC research has also advanced our understanding of fundamental science of complex systems. The knowledge of structure/function relationships that we acquire from DSSC research can be extrapolated to other solar cell technologies leading to the establishment of the guidelines for the preparation of new material components. 1.4 Literature review on photoanode materials used in DSSCs Breakthrough in dye-sensitized photovoltaic technology came through the introduction of mesoporous photoanode structure making the internal surface area 2 orders of magnitude higher than the geometric area. A transparent mesoporous film made of wide band gap semiconductor is deposited on a conductive substrate. Incident light passes though this semiconductor film and gets absorbed by the monolayer of the dye on the nanostructure. The purpose of the semiconductor film is to make the internal area higher so that the monolayer of dye can effectively deliver good light harvesting ability. Furthermore, the nanostructure collects the photoinjected electron and transports it to the conductive substrate. However, the advantages of the nanostructures come with their drawbacks. It makes the transport difficult due to polycrystalline nature and numerous grain boundaries. Additionally, higher surface area unlocks the chances of interfacial recombination processes. 7 The research on DSSC photoanode mainly focuses on enhancing the electron transport through nanostructured photoanode and diminution of the interfacial recombination processes. At the same time, the charge injection from the excited dye to the semiconductor needs to be energetically and entropically favored. The conduction band of the semiconductor is desired to be at lower energy than the LUMO of the dye. Additionally, large excess of density of states in the conduction band than the photoexcited dye entropically favors the electron flow from dye molecular orbital to semiconductor conduction band. TiO2 has been the most studied material as DSSC photoanode. Among many crystalline forms of TiO2, rutile and anatase are widely studied. Anatase has been the preferred choice in DSSC because of its larger bandgap (3.2 vs. 3.0 eV for rutile) and higher (more negative) conduction band edge leading to larger photovoltage. Apart from TiO2, Zinc Oxide (ZnO) has received significant attention as DSSC photoanode. Higher electron mobility and ease in making desired transparent nanostructures like nanoparticles, nanowires, nanorods, nanotubes, tetrapods, nanoflowers, nanosheets and branched nanostructures are some of the great advantages of ZnO photoanodes. ZnO based DSSC has achieved a promising photoconversion efficiency of 8% so far.30 However, due to the poor chemical stability in both acidic and basic solution, ZnO has never been the best alternative to TiO2. Most of the dyes used in DSSCs have acidic carboxylic anchoring group which reacts with ZnO, form insoluble Zn2+-dye aggregates and precipitate in the pores of the photoanode film. These inactive dye complexes absorb light but do not inject electrons into the semiconductor. Modification of dye anchoring group and applying a “protective” layer on the ZnO material have been reported to improve the stability. Nguyen et al. reduced the precipitation of the Zn2+-dye aggregates by using a Ru complex with single carboxylated bipyridyl ligand and a promising 4% 8 efficiency was reported. Park and coworkers have reported a SiO2 coated ZnO system with improved chemical stability and a photoconversion efficiency of 5.2%. Very interestingly, study of the material property and transport characteristics of nanostructured ZnO photoanode system along with TiO2 has provided some valuable insights of structure/function correlation. ZnO has been reported to have lower density of trap states than TiO2 and the electron transport through nanostructured ZnO film is faster than TiO2 which is attributed to the fact that electron transport occurs through trapping-detrapping mechanism. The electron life time of ZnO is also longer than that of TiO2. Thus, a semiconductor photoanode having lower density of trap states is expected to offer superior performance in the context of diminishing recombination and attaining higher photovoltage. SnO2 is another wide band gap semiconductor studied as DSSC photoanode material. Its ECB is about 0.5 V lower than TiO2 and potentially can be utilized with a dye with low lying LUMO. New redox shuttles with lower redox potential needs to be explored in order to optimize the photovoltage with SnO2 and dyes with lower energetics. The record efficiency of a DSSC constructed with SnO2 photoanode is little over 3% so far.31 Niobium Oxide (Nb2O5) has also received some attention and generally offer higher photovoltage than TiO2 which is attributed to the higher conduction band position of Nb2O5. However, making high surface area transparent Nb2O5 photoanode is challenging along with the poor dye loading. In 2011, Zhang et al. have addressed these issues and made crystalline Nb3O7(OH) nanorod film through solvothermal process. No annealing was done to preserve the hydroxyl group on the surface which greatly enhanced the dye loading and thus improvement of the current. Their reported photoconversion efficiency of 6% has been the best with this material so far. 9 Ternary oxides have also attracted researchers’ attention as they offer a wider possibility of tuning the physical–chemical properties by altering the relative ratios of the cationic components. SrTiO3, SrSnO3, Zn2SnO4, CdSnO3 and BaSnO3 are among the ternary metal oxides utilized as photoanode.32 However, performance have been much lower than the DSSC constructed with TiO2 photoanode and further studies need to be performed to assess their potential as DSSC photoanode material. 1.5 Motivation for studying photoanode properties in DSSC The journey of DSSC progress has largely kept the initial champion components, like TiO2 photoanode, Ruthenium dye, Iodide redox shuttle and Pt counter electrode, unchanged. In spite of the huge boom in DSSC research and a large effort to maximize the overall efficiency, it remained stagnant for almost 18 years. The recent breakthrough in the efficiency came through the systematic understanding and improvement of the outer sphere redox shuttles which showed pitiful performance initially.33 Engineering the absorption characteristics, energetics and stability of sensitizers have attracted broad attention. Ruthenium bipyridine dyes have remained the best performing sensitizer for two decades starting from the first introduction in nanostructured DSSC. Through extensive research with various dyes, new types of champion dyes have been introduced and a donor-π-bridge-acceptor porphyrin dye has been found to be the new champion. In counter electrode part, the former champion platinum has been challenged by graphene. Thus among the main components, all the initially champion components have been recently changed except the photoanode material, titanium dioxide. The photoanode plays an essential role in the electron injection process from the dye, transportation of these electrons and recombination processes. Significant energy loss occurs during these processes which needs to be largely minimized in order to push the efficiency towards 20%. The ideal photoanode should 10 demonstrate quantitative charge injection with minimal energy difference between the dye excited states and the conduction band edge. Additionally, surface state mediated recombination processes are among the issues with TiO2 photoanode while using outer sphere redox shuttle with highly positive redox potential and fast self exchange kinetics.17 New photoanode needs to be explored to get rid of this recombination issue. Moreover, relying on a single photoanode material has restricted the use of many sensitizers because of the mismatch in energetics. There are several studies with alternative photoanodes like ZnO, SnO2, Nb2O5, SrTiO3, SrZnO3 etc. in literature.34 However, their performances are not comparable to TiO2. Moreover, the reason behind their poor performance has remained unexplored and needs further studies to understand the parameters which are detrimental to the photoconversion efficiency of DSSCs. 1.6 Overview of the thesis Lack of reliable methods to study the important material properties like energetics and charge distribution has been a major issue behind it. In order to obtain a detailed understanding of relevant electron-transfer processes the position of CB edge is essential. In the Chapter 2, a spectroelectrochemical method is described to determine the CB edge of mesoporous TiO2 photoanode. These mesoporous photoanodes are known to have large density of trap states and the accurate energy distribution as well as spatial location of the trap states are very important features controlling the key electron transfers processes associated with this photoanode material. A reliable method to study this trap state distribution is described in the Chapter 3. Another crucial parameter required in the spectroelectrochemical study of electron transfer kinetics in DSSC is the extinction coefficient of conduction band electrons (CB). However, due to the presence of large excess of trapped electrons, the nature of the absorbing species has been a long standing debate for more than 2 decades which prevented any accurate determination of CB. In 11 the Chapter 4, a detailed spectroeletrochemical study of the absorbance behavior of electrons in TiO2 film is presented. TiO2 is mostly used as photoanode material in DSSCs. Though TiO2 has remained the best performing DSSC photoanode so far, it has some disadvantages like sluggish charge transport; trap states mediated recombination processes; limited energy match with sensitizers etc. However, alternative materials are still far from being utilized to resolve these issues. A systemic investigation of the photoanode materials to establish a correlation between material properties and photoanode functions has been missing in literature. Chapter 5 describes the study of mesoporous SrTiO3 photoanode. In Chapter 6, the maximum attainable efficiency of DSSC technology is assessed and the necessity of systematic improvement of photoanode part is pointed out. Chapter 7 looks into the viability of DSSC technology for reliable renewable energy production along with the outlook for the future directions in this field. Overall, we have developed methods to determine crucial properties of photoanode like energetics and electron distributions. Our effort also is to explore alternative photoanode materials and assess their viability as DSSC photoanode. 12 REFERENCES 13 REFERENCES (1) Lewis, N. S.; Nocera, D. G. Proc. Natl. Acad. Sci. 2006, 103, 15729–15735. (2) Lewis, N. S. MRS Bull. 2007, 32, 808–820. (3) IEA. Key world energy statistics; 2016. (4) BP. Statistical Review of World Energy. 2016. (5) IEA. CO2 Emissions from Fuel Combustion. 2016. (6) Docampo, P.; Guldin, S.; Leijtens, T.; Noel, N. K.; Steiner, U.; Snaith, H. J. Adv. Mater. 2014, 26, 4013–4030. (7) IEA. Trends 2015 - In Photovoltaic Applications; 2015. (8) O’Regan, B.; Grätzel, M. Nature 1991, 353, 737–740. (9) Grätzel, M. J. Photochem. Photobiol. C Photochem. Rev. 2003, 4, 145–153. (10) Crossland, E. J. W.; Noel, N.; Sivaram, V.; Leijtens, T.; Alexander-Webber, J. A.; Snaith, H. J. Nature 2013, 495, 215–219. (11) Brown, T. M.; De Rossi, F.; Di Giacomo, F.; Mincuzzi, G.; Zardetto, V.; Reale, A.; Di Carlo, A. J. Mater. Chem. A 2014, 2, 10788–10817. (12) Tagliaferro, R.; Colonna, D.; Brown, T. M.; Reale, A.; Di Carlo, A. Opt. Express 2013, 21, 3235. (13) Hashmi, G.; Miettunen, K.; Peltola, T.; Halme, J.; Asghar, I.; Aitola, K.; Toivola, M.; Lund, P. Renew. Sustain. Energy Rev. 2011, 15, 3717–3732. (14) Mathew, S.; Yella, A.; Gao, P.; Humphry-Baker, R.; Curchod, B. F. E.; Ashari-Astani, N.; Tavernelli, I.; Rothlisberger, U.; Nazeeruddin, M. K.; Grätzel, M. Nat. Chem. 2014, 6, 242–247. (15) Fakharuddin, A.; Jose, R.; Brown, T. M.; Fabregat-Santiago, F.; Bisquert, J. Energy Environ. Sci. 2014, 7, 3952–3981. (16) Yamaguchi, T.; Tobe, N.; Matsumoto, D.; Nagai, T.; Arakawa, H. Sol. Energy Mater. Sol. Cells 2010, 94, 812–816. (17) Ondersma, J. W.; Hamann, T. W. Coord. Chem. Rev. 2013, 257, 1533–1543. (18) Moser, J. Monatsh. Chem. 1887, 8. 14 (19) Gurney, R. W.; Mott, N. F. Proc. R. Soc. A Math. Phys. Eng. Sci. 1938, 164, 151–167. (20) Putseiko, E. K. and Terenin, A. N. Zhur. Fiz. Khim. 1949, 23. (21) Gerischer, H.; Michel-Beyerle, M. E.; Rebentrost, F.; Tributsch, H. Electrochim. Acta 1968, 13, 1509–1515. (22) Gerischer, H.; Tributsch, H. Ber. Bunsen-Ges. Phys. Chem. 1968, 72, 437–445. (23) Osa, T.; Fujihira, M. Nature 1976, 264, 349–350. (24) Desilvestro, J.; Graetzel, M.; Kavan, L.; Moser, J.; Augustynski, J. J. Am. Chem. Soc. 1985, 107, 2988–2990. (25) Vlachopoulos, N.; Liska, P.; Augustynski, J.; Graetzel, M. J. Am. Chem. Soc. 1988, 110, 1216–1220. (26) Nazeeruddin, M. K.; Kay, A.; Rodicio, I.; Humphry-Baker, R.; Mueller, E.; Liska, P.; Vlachopoulos, N.; Graetzel, M. J. Am. Chem. Soc. 1993, 115, 6382–6390. (27) Nazeeruddin, M. K.; Péchy, P.; Grätzel, M. Chem. Commun. 1997, 1, 1705–1706. (28) Nazeeruddin, M. K.; Péchy, P.; Renouard, T.; Zakeeruddin, S. M.; Humphry-Baker, R.; Comte, P.; Liska, P.; Cevey, L.; Costa, E.; Shklover, V.; Spiccia, L.; Deacon, G. B.; Bignozzi, C. a.; Grätzel, M. J. Am. Chem. Soc. 2001, 123, 1613–1624. (29) Yella, A.; Lee, H.-W.; Tsao, H. N.; Yi, C.; Chandiran, A. K.; Nazeeruddin, M. K.; Diau, E. W.-G.; Yeh, C.-Y.; Zakeeruddin, S. M.; Gratzel, M. Science 2011, 334, 629–634. (30) He, Y.; Hu, J.; Xie, Y. Chem. Commun. 2015, 51, 16229–16232. (31) Birkel, A.; Lee, Y.-G.; Koll, D.; Meerbeek, X. Van; Frank, S.; Choi, M. J.; Kang, Y. S.; Char, K.; Tremel, W. Energy Environ. Sci. 2012, 5, 5392–5400. (32) Li, Y.; Zhang, H.; Guo, B.; Wei, M. Electrochim. Acta 2012, 70, 313–317. (33) Gregg, B. a.; Pichot, F.; Ferrere, S.; Fields, C. L. J. Phys. Chem. B 2001, 105, 1422–1429. (34) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Chem. Rev. 2010, 110, 6595– 6663. 15 Chapter 2 Spectroelectrochemical Determination of the Conduction Band Edge Potential Adapted from: Band energies of nanoparticle semiconductor electrodes determined by spectroelectrochemical measurements of free electrons, Dhritabrata Mandal and Thomas W. Hamann, Phys. Chem. Chem. Phys. 2015, 17, 11156–11160 2.1 Introduction Nanostructured semiconductor electrodes have garnered intense recent interest for their use in various solar energy conversion systems since they offer the possibilities of circumventing low efficiencies associated with short diffusion length bulk materials as well as optimizing light absorption in dye-sensitized solar cells.1–7 In such context, knowledge of the conduction (ECB) and valence band (EVB) edge positions are the most important electronic properties in order to optimize the performance and obtain a detailed understanding of relevant electron-transfer processes.8 However, there is no reliable direct method to measure the band edges in nanostructured semiconductor electrodes. Efforts to determine the conduction band edges in these nanostructured photoelectrodes by electrochemical methods have failed. Mott–Schottky analysis is widely used to determine the band edge positions in a compact semiconductor electrode. The central idea of Mott-Schottky equation is modeling the depletion region capacitance as a parallel plate capacitor. However, the mesoporous photoanodes used in DSSCs are generally made of small nanoparticles as shown in Figure 2-1. The Mott-Schottky analysis cannot be used here as the dimensions of the nanoparticles are usually smaller than the depletion region.9,10 The nanoparticles surrounded by 16 the electrolyte solution are unable to support traditional band bending required for the MottSchottky analysis to be valid. On the other hand, this depletion region capacitance is associated with the dopant density (ND) of the semiconductor. The typical value of ND in anatase TiO2 material is generally in the range of 1016  1017 cm-3.11–13 Assuming this value is valid in nanoparticles having diameter of 10 nm, one dopant can be present in every 20 nanoparticles or even in 200 nanoparticles. Thus, a homogenous distribution of depletion region capacitance throughout the nanoparticle network is not possible and Mott-Schottky equation cannot be applied. Band edge position can also be determined from the photocurrent onset potential.14 However, determining the exact potential where the photocurrent can be detected in the presence of the dark current is uncertain. Figure 2-1 SEM image of mesoporous TiO2 photoanodes used in DSSC. The image on the right shows the cross section of a TiO2 film. Fitzmaurice et al. developed a spectroeletrochemical method to determine the flat band positions in mesoporous semiconductor films.15–21 The absorption of a transparent film was measured while the film was charged by applying large negative bias (reductive) during the cyclic voltammetric scan. This absorption in the visible-near IR region was considered to be 17 associated with the intra-band absorption of the free electrons in the conduction band. Their model is based on the assumption of an accumulation layer, created within each nanocrystalline particle, under reductive potential more negative than the flat-band potential. However, their model and the associated assumptions have been debated and the validity of their approach has been questioned.9,22–27 Many of their key assumptions e.g. flat interface, Schottky junction, formation of accumulation layer are very unlikely to occur in these small nanoparticles (diameter 10- 20 nm). Additionally, the large negative bias can cause unwanted change in the electrode especially in the presence of small cations. Scan rate dependence of the measured absorption is the other drawback of this method, which implies that the system is not under equilibrium and all the electronic states are not filled during the absorption measurement.28 Their model also assumes that the conduction band edge does not change while raising the Fermi energy. However such nanostructured TiO2 films with high roughness factor should have significant electroactive states on the surface which can cause band edge shift.29 Moreover, it has been argued that the absorption spectrum arises due to the absorption of the trapped electrons instead of free electrons in the CB.30–34 A new variable temperature spectroelectrochemistry (VTSEC) method was introduced as a way to overcome these issues.28 The VTSEC method is not ideal, however, since it relies on two assumptions: the dominant absorbing species of the visible and near IR (Vis/NIR) wavelengths measured are free conduction band electrons and the only effect of the variable temperature is the free electron concentration. The first assumption of the absorbing species has not been firmly established and is the subject of continued disagreement, with alternative assignments of a trapped electron absorption or a combination of free and trapped electrons (i.e. total electron concentration).22,31,34,35 Regarding the second assumption, it is likely that there is a temperature 18 dependence of the band gap, dielectric constant, phonons (thus extinction coefficient) of TiO2 employed that was not accounted for and may lead to a significant error. Further, VTSEC is fairly tedious to control and not amenable to being a routine measurement. The motive of this work is primarily to accurately determine the conduction band edge position of nanostructured semiconductor electrodes. A TiO2 nanoparticle electrode was biased under negative (reductive) potential which increases the electron population in the TiO2 electrode and the optical absorption was measured. The change in the band gap absorption was correlated to the electron density in the conduction band and the position of the CB edge potential was determined. 2.2 Experimental 2.2.1 TiO2 photoanode preparation The transparent mesoporous anatase TiO2 electrode was prepared by spreading HT/SP TiO2 paste (Solaronix) on fluorine-doped tin oxide (FTO) conducting glass (Hartford Glass) by the doctor-blade method, followed by sintering at 500°C for 15 min in air. The thickness of the TiO2 film was measured by a Dektak3 Surface Profiler and found to be 6.950.05 μm. The porosity, p, of the mesoporous TiO2 film was also determined. 15 TiO2 films (geometric area 64 mm2 and thickness 6.95 m) were collected from the FTO by scraping out with a sharp blade. The total mass of the collected TiO2 material was measured. Total volume of the 15 films was then calculated which is 64 mm2  6.95 m 15 and the mass of the solid TiO2 material without any pores was calculated from the density of anatase material, 3.93 g / cm-3. Porosity is calculated using the equation, p = 1 – (measured mass of porous film / calculated mass of solid material). 19 2.2.2 Electrolyte preparation Aqueous electrolyte solutions were made by dissolving recrystallized potassium chloride (KCl) salts into Millipore water (18.2 M-cm). A Fischer Scientific Accumet pH meter was used to measure the pH of the solutions. Drops of hydrochloric acid (HCl) and potassium hydroxide (KOH) were added to adjust the pH to the desired values, pH 12.8 and 2.0 used here. 2.2.3 Spectroelectrochemical measurement Figure 2-2 Instrumental set-up of spectroelectrochemical measurement. In spectroelectrochemical measurements, the transmittance (%T) of the film was measured by a Perkin Elmer Lamda 35 UV-Vis spectrophotometer under an external bias provided by a potentiostat (-Autolab, Metrohm). A three electrode electrochemical cell, made by customizing a commercial spectroscopic cuvette (Quartz Spectrophotometer Cell, Starna Cells, Inc), was placed into the UV-Vis spectrophotometer such that the TiO2 film remained perpendicular to the 20 optical path as shown in Figure 2-2. A platinum wire (diameter 1.5 mm) was used as counter electrode and a commercial “Leakless Miniature” Ag/AgCl reference electrode (EDAQ ET072) was used. The TiO2 films were aged overnight in contact with the electrolyte solutions prior to any experiment. The temperature of the electrochemical cells was 220.5 ⁰C during the measurements. During spectroelectrochemical measurements, each film was stabilized at an applied bias for 5 minutes and %T of the electrode was measured. A control experiment was done with bare FTO electrode to confirm that the FTO has no significant contribution into the change in %T with varying potential. Stability of the film as well as reproducibility of the results were checked both spectroscopically as well as electrochemically to make sure that there is no irreversible change in the film during the measurements. The difference absorbance spectrum, A, is the absorbance change under external bias. At first, the transmittance was measured under a significantly positive potential to discharge all the stored electrons from the film. Then a negative potential was applied to increase the electron population and the transmittance was measured. Absorbance, A, was calculated from the %T by using the relation, A = 2  log10%T. Absorbance at positive potential was subtracted from the absorbance at negative bias to calculate A. While determining the band gap using a Tauc plot, the absorption spectra measured while using bare FTO electrode was subtracted from the spectra measured with TiO2 film on FTO. Thus the absorption of TiO2 film was calculated by eliminating the contribution of everything else in the optical path (FTO, cuvette and electrolyte). The contribution of the free electron absorption was calculated in the wavelength range of 350 – 400 nm by extrapolating the absorption spectra measured in the visible range (discussed in Chapter 4). This calculated absorption was then excluded from the total absorption measured in this region to determine the 21 absorption of the band gap transition only. Absorption co-efficient () was calculated from absorbance by using the relation  = (2.303  A) / l, where A is the absorbance and l is the thickness of the film.36,37 2.3 Results and discussion 2.3.1 Absorption spectra of the TiO2 electrode Due to the good electronic connectivity among nanoparticles and equipotential environment surrounding the nanoparticles, the Fermi level (EF) within the TiO2 nanoparticle electrode can be controlled by applying an external bias to the conductive substrate.38–41 The measured absorption at a certain applied potential (E) can be correlated to the electron population in the semiconductor film at the Fermi level EF which is equal to qE (q is the energy of an electron). Here, E represents energy and E (=E/q) represents the electrochemical potential. This potentiostatically controlled Fermi level is directly related to the density of electron in the conduction band. When the Fermi level moves closer to the conduction band (CB) the density of the electrons (nCB) increases exponentially according to the following relation: nCB  N CB exp (EF  ECB ) k bT (1) where ECB is the conduction band minima and NCB is the density of states (DOS) in the conduction band. Figure 2-3 shows typical difference absorption, A, spectra of a TiO2 film immersed in a pH 12.8 aqueous electrolyte. The absorbance increases in the visible and the near-IR regions with increasing (negative) applied potential (E) with a concomitant decrease at wavelengths shorter than ~380 nm. The characteristics of this absorption in visible  near IR region are discussed in Chapter 4 in details. 22 The band gap transition in anatase material occurs below 380 nm and the decrease in A can be explained by the blue shift of the band gap absorption.22 This shift is attributed to the filling of the low energy states of conduction band which block the lowest energy band gap transition.42 Blue shift in the band gap absorption has been observed in the highly doped semiconductor and is known as Burstein-Moss shift.42–47 This shift has been quantitatively analyzed in our study to determine the density of the CB electrons and the position of CB edge. Figure 2-3 Change in absorbance of a TiO2 electrode in contact with a pH 12.8 aqueous electrolyte when potential is stepped from + 0.455 V to  1.245 V (red),  1.295 V (green),  1.345 V (yellow),  1.395 V (blue) and  1.445 V (pink) vs. SCE (from bottom to top). 2.3.2 Determining band gap energy The absorption edge due to the bandgap transition can be determined from a Tauc plot using the relation,37 ( hv)  A' (EG  hv) 23 (2) where  is the absorption co-efficient, h is energy of incident light, EG is the band-gap energy, A’ is a constant and the value of  depends on the characteristics of the transition process. Figure 2-4 Top: Plot of absorption coefficient () vs. wavelength () of TiO2 film. Bottom: Tauc plot of a TiO2 film. It was measured under an external positive bias + 0.445 V vs. SCE to discharge any stored electrons from the nanoparticles. The electrolyte solution is 0.2 M KCl at pH 12.8. 24 For the indirect allowed transition,  is 1/2 and it is 2 for the direct allowed transition. The band gap transition in anatase is generally accepted to be indirect allowed transition and  = 1/2 is used to determine the band-gap energy.48–51 Figure 2-4 shows a typical Tauc plot used to determine the band gap energy of the anatase nanoparticles. The steep fraction of the (h)1/2 vs. h curve fitted between 3.444 eV to 3.547 eV (360 - 350 nm) was extrapolated to determine the band gap and it was found to be 3.209 eV which excellently matches to the literature values for EG of anatase TiO2.48–51 The range of photon energies utilized to fit the Tauc plot was chosen where a good straight line fit was observed. The energy region should not be too close to the band gap as various factors (e.g. Urbach tail) can dominate the overall absorption spectra where the absorption of band gap transition is still relatively low.37 On the other hand, in a scenario when the photon energy is very much higher the band gap energy, the parabolic approximation of E-k plot is no longer valid. Also, the direct band gap transition (which occurs at higher energy than the indirect transition in anatase material) can contribute to the absorption. The maximum energy applied here is 3.547 eV ( = 350 nm) which is in the energy range where the Tauc plot is generally fitted in literature for the indirect transition in anatase and at this energy the direct band gap transition is very unlikely to occur.52 Additionally, we found a value ~ 3.2 eV for band gap which excellently matched with the literatures values. Participation of the absorption due to direct band gap transition at higher energy is expected to abruptly increase the absorption rise which was not observed here. 2.3.3 Sources of uncertainty in determining band gap energy In the indirect transition, both the energy and the momentum are changed. As a result, phonon interaction (resulting in thermal vibrations in the crystal lattice) is required.37 Therefore the band gap transition becomes a two-step process assisted by either phonon absorption or 25 phonon emission which provides the required momentum change during the transition. The energy required for this transition mainly comes from the photon absorbed. So the energy required to excite an electron from the valence band to the conduction band can be written as,37 h  EG  Ep (3) where EP is the energy of the phonon required to compensate the difference in momentum between the minima and maxima of the conduction and valence bands, respectively. The absorption co-efficient for a phonon assisted indirect transition can be given by the following equations,  a (hv)   e (hv)  A" (hv  EG  E p ) 2 E exp p  1 k bT A" (hv  EG  E p ) 2  E  1  exp   p   k bT  (4a) (4b) where a(h) and e(h) are the absorption co-efficients corresponding to phonon absorbed and phonon emitted process, respectively, A” is a constant, kb is Boltzmann constant and T is absolute temperature. Now, both phonon absorption and emission are possible as the range of applied photon energy considered to calculate the band edge is much higher than EG. So, the total absorption co-efficient is given by,  (hv)   a (hv)   e (hv) (5) These equations imply that phonon assisted absorption co-efficient is temperature dependent. The phonon energy is expected to be constant while changing the potential at constant temperature. It should also be mentioned that it is not straightforward to determine the absolute band gap in case of phonon assisted indirect transition. Additionally, in presence of a high 26 dopant density and/or impurity, charge-charge and/or charge-ionic impurity scattering can also provide the momentum conservation required for the indirect transition.37 The dopant density as well as impurity concentration in nanostructured TiO2 films has not been precisely determined so far. Broadening of the absorption curves also contributes to the uncertainty in determining the band gap energy precisely. Band tail, which is expected to be present in our sample due to disorder effects and/or defects below the conduction band, can also contribute to this uncertainty. Significant broadening of absorption curve has been observed in our study (Figure 2-4). It is very difficult to get rid of this broadening which might cause some error in our band gap determination. At a constant temperature and electrolyte composition, however, the magnitude of the broadening remains constant and does not contribute to the shift in the band gap absorption, which is of primary interest here. In summary, at constant temperature and electrolyte environment the error in determining the band gap transition remains constant which has allowed us to accurately determine the magnitude of the shift in band gap energy under external potential bias. 2.3.4 Burstein-Moss shift under potentiostatic bias Band gap transition is defined as the lowest energy required to excite an electron from the top of the VB to the lowest energy states of the CB. Increasing electron population in the CB fills the lowest energy states and band gap transition requires some extra energy causing a blue shift in the band gap absorption. This blue shift has been reported to occur due to increasing CB electron population by doping semiconductor material and is generally described as Burstein-Moss shift.42–47 In our experiment, reductive bias has been applied to increase the electron population in the CB and the similar phenomenon has been observed. 27 Figure 2-5 Energy vs. crystal momentum diagram (E-k) of a semiconductor with an indirect band gap showing the Burstein-Moss shift due to the filling of the lowest energy states in the CB under negative (reductive) potential. Electron transition from the highest-energy state in the valence band to the lowest-energy state in the conduction band is not possible without a change in momentum. The transition occurs via the absorption or emission of a phonon (red horizontal arrow), which provides the required momentum change during the transition. The energy required for this transition mainly comes from the photon absorbed (violet vertical arrow). When the lowest energy states of the conduction band become occupied by the potentiostatically injected electrons, the energy to excite electrons from valence band to the unoccupied states of the conduction band increases. As a result, a blue shift in the band gap absorption is observed as shown in Figure 2-6. The magnitude of this band gap widening (EG) is correlated to the density of electron in the CB according to the following equation: EG  h 2  3nCB  8m*    2/3 (6) where m* is the reduced effective mass of the charge carriers and nCB is the concentration of electrons in the conduction band. 28 Figure 2-6 Plot of absorption coefficient () vs. wavelength () of TiO2 film showing blue shift of the band gap absorption energy under negative (reductive) potential. The electrolyte solution is 0.2 M KCl at pH 12.8. Figure 2-7 Band gap widening occurs while moving from +0.445 V (red) towards more negative potentials -1.245 V (green), -1.345 V (yellow), -1.445 V (blue) vs. SCE. The electrolyte solution is 0.2 M KCl at pH 12.8. 29 Figure 2-8 Plot of Burstein-Moss shifts (EG) at various applied potentials. Correlating the observed blue shift in the band gap absorption with the density of the electrons in the CB is based on a few assumptions. The parabolic approximation of energy bands were made here, which basically describes that the charge carriers (electrons, holes) behave as free carriers with a parabolic momentum (k) dependence. This parabolic approximation provides values for the effective mass of the electron (me*) from the bottom energy level of the conduction band and hole effective mass (mh*) from the top of the valence band. However, at energies far from the band edges this approximation loses its validity and the effective masses become energy dependent. In our experiment, the increase in the band gap transition energy is  17.8 meV when the bottom of the CB is filled which is less than the 1kbT value at room temperature. It is reasonable to assume that the parabolic approximation is valid in this energy range which provides energy independent values for me* and mh*. Also, a constant value for the DOS at CB, NCB, was calculated from the effective mass of electron according to the following equation, 30  me*k bT  NCB  2  2   2  3/2 (7) Burstein-Moss shift was determined at a range of applied potentials and the corresponding nCB was calculated from equation 6. A range of values (1me to 10me) are found in literature for me*.51,53,54 Liu and Wang et al. studied the electron-transport in mesoporous TiO2 photoanode in a working DSSC which depends on the DOS in the CB.55 They estimated an effective electron mass of 2.3me from the measured electron density in the CB. In our study, me* = 2.3me was used. Thulin et al. calculated a value of ~ 1.8me for mh* and Fitzmaurice reported a value of 0.80.2me.52,56 These values were determined based on several assumptions and the accurate value for the mh* in anatase nanoparticle is uncertain. In our study, we used 1.0me for mh*. The uncertainty in mh* should not have significant effect in the values of ECB/q (<10 mV in a range of 1me to 5me). Now, once nCB is known at a given potential, E (= EF/q), the conduction band edge energy can be calculated directly from equation 1. Table 2-1 shows the values of ECB/q calculated under a range of applied potentials. These results indicate an upward shift of the conduction band edge when the applied potential is increased i.e. more negative. This phenomenon can be explained by the potential drop across the Helmholtz layer. Accumulation of electrons in the electrically active states on the surface or within close vicinity of surface of the mesoporous TiO2 films can cause a change in the electric field of the Helmholtz layer resulting in this potential drop.22,57,58 This high surface area mesoporous TiO2 electrode is made of small TiO2 nanoparticles (10 nm diameter) connected to each other. Various types of the electroactive states exist on the surface which are energetically located below the CB. Truncation of the crystal lattice on the particle surface as well as the point defects are the possible reasons behind the origin of these surface states. 31 Detailed discussions about these electroactive states and the potential drop are provided in Chapter 3. Table 2-1 Conduction band positions were calculated at various applied potentials from the magnitudes of the band gap shift. EG was calculated from the Tauc plots at the applied potentials mentioned below and EG = EG (negative bias) – EG (+0.445 V vs. SCE). The electrolyte solution is 0.2 M KCl at pH 12.8. E vs. SCE / V EG / meV nCB / cm–3 ECB/q / V -1.045 0.22  0.42 2.97  8.47 × 1016 -1.249  0.073 -1.095 0.47  0.42 0.93  1.23 × 1017 -1.270  0.034 -1.145 1.05  0.42 3.09  1.84 × 1017 -1.289  0.015 -1.195 2.25  0.42 9.70  2.70 × 1017 -1.310  0.007 -1.245 4.52  0.42 2.76  0.38 × 1018 -1.334  0.004 -1.295 7.03  0.41 5.36  0.47 × 1018 -1.367  0.002 -1.345 10.07  0.41 9.18  0.56 × 1018 -1.403  0.002 -1.395 13.74  0.41 1.46  0.07 × 1019 -1.441  0.001 -1.445 17.78  0.41 2.16  0.07 × 1019 -1.481  0.001 Another interesting point here is that when the Fermi level of electrons moves very close to the CB edge, EF  ECB becomes as small as 36 meV. Now, equation 1 is considered to be valid when EF  ECB  3kbT where kbT is about 26 meV because of the breakdown of the Boltzmann statistics. Liu and Wang et al. observed that EF eventually reaches to the transport energy in mesoporous TiO2 suggesting that EF is too close to ECB.55 Thus, in Table 2-1, the values for ECB/q, calculated when EF/q is more negative than -1.295 vs. SCE, are questionable. The validity of the parabolic approximation of E-k diagram when the energy is close to the band edges and 32 the applicability of equation 1 (Boltzmann statistics) are two primary conditions for our method to be valid. 2.4 Conclusion: Spectroelectrochemical measurement has been performed on transparent polycrystalline TiO2 films. Under negative potential, a blue shift of the band gap transition has been observed. The observed blue shift can be explained by the filling of the lowest energy states of the CB under negative bias. This type of blue shift i.e. an apparent band gap widening is known as BursteinMoss shift. Quantitative analysis of this blue shift has provided the density of electrons in the CB under a range of applied potentials. Position of CB edge has been calculated at different Fermi level and an upward shift of the CB has been observed. This shift is explained by the potential drop across the Helmholtz layer causing unpinning of the band positions. The electroactive surface states on these high surface area electrodes accumulate electrons on the surface of the nanoparticles under negative bias.33,57–59 To maintain the charge neutrality, positive charges, i.e. the cations in the electrolyte, are adsorbed at the semiconductor/electrolyte interface which causes potential drop in the Helmholtz layer. This potential drop is discussed in Chapter 3. The band edge unpinning has been pointed out while studying the electron transfer processes in DSSCs.55 However, due to the absence of any reliable methods for quantitative analysis of the band edge shift, it has been either ignored or only estimated indirectly. Thus, this work solves a long standing problem of being able to determine the band edge positions in nanoparticle semiconductor electrodes. 33 REFERENCES 34 REFERENCES (1) Ardo, S.; Meyer, G. J. Chem. Soc. Rev. 2009, 38, 115–164. (2) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Chem. Rev. 2010, 110, 6595– 6663. (3) Youngblood, W. J.; Lee, S.-H. A.; Maeda, K.; Mallouk, T. E. Acc. Chem. Res. 2009, 42, 1966–1973. (4) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q.; Santori, E. a.; Lewis, N. S. Chem. Rev. 2010, 110, 6446–6473. (5) Chen, X.; Shen, S.; Guo, L.; Mao, S. S. Chem. Rev. 2010, 110, 6503–6570. (6) Gust, D.; Moore, T. A.; Moore, A. L. Acc. Chem. Res. 2009, 42, 1890–1898. (7) Kay, A.; Cesar, I.; Grätzel, M. J. Am. Chem. Soc. 2006, 128, 15714–15721. (8) Ondersma, J. W.; Hamann, T. W. J. Am. Chem. Soc. 2011, 133, 8264–8271. (9) Hagfeldtt, A.; Gratzel, M. Chem. Rev. 1995, 95, 49–68. (10) Peter, L. M. Phys. Chem. Chem. Phys. 2007, 9, 2630. (11) Cameron, P. J.; Peter, L. M. J. Phys. Chem. B 2005, 109, 7392–7398. (12) O’Hayre, R.; Nanu, M.; Schoonman, J.; Goossens, A. J. Phys. Chem. C 2007, 111, 4809– 4814. (13) Redmond, G.; O’Keeffe, A.; Burgess, C.; MacHale, C.; Fitzmaurice, D. J. Phys. Chem. 1993, 97, 11081–11086. (14) Beranek, R. Adv. Phys. Chem. 2011, 2011, 1–20. (15) Brian, O.; Grstzel, M.; Fitzmaurice, D. Chem. Phys. Lett. 1991, 183, 89–93. (16) Rothenberger, G.; Fitzmaurice, D.; Graetzel, M. J. Phys. Chem. 1992, 96, 5983–5986. (17) O’Regan, B.; Graetzel, M.; Fitzmaurice, D. J. Phys. Chem. 1991, 95, 10525–10528. (18) Enright, B.; Redmond, G.; Fitzmaurice, D. J. Phys. Chem. 1994, 98, 6195–6200. (19) FITZMAURICE, D. Sol. Energy Mater. Sol. Cells 1994, 32, 289–305. (20) Boschloo, G.; Fitzmaurice, D. J. Phys. Chem. B 1999, 103, 3093–3098. (21) Redmond, G.; Fitzmaurice, D. J. Phys. Chem. 1993, 97, 1426–1430. (22) Cao, F.; Oskam, G.; Searson, P. C.; Stipkala, J. M.; Heimer, T. a; Farzad, F.; Meyer, G. J. 35 J. Phys. Chem. 1995, 99, 11974–11980. (23) Lyon, L. A.; Hupp, J. T. J. Phys. Chem. 1995, 99, 15718–15720. (24) Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S.-E. J. Phys. Chem. B 1997, 101, 7710–7716. (25) Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S.-E. J. Phys. Chem. B 1997, 101, 7717–7722. (26) Ohta, M.; Koumura, N.; Hara, K.; Mori, S. Electrochem. commun. 2011, 13, 778–780. (27) Liu, Y.; Jennings, J. R.; Huang, Y.; Chem. C 2011, 115, 18847–18855. (28) Ondersma, J. W.; Hamann, T. W. Energy Environ. Sci. 2012, 5, 9476. (29) Kopidakis, N.; Neale, N. R.; Zhu, K.; Van De Lagemaat, J.; Frank, a. J. Appl. Phys. Lett. 2005, 87, 1–3. (30) Berger, T.; Anta, J. A.; Morales-flo, V. J. Phys. Chem. C 2012, 116, 11444–11455. (31) Berger, T.; Anta, J. a; Morales-Flórez, V. Phys. Chem. Chem. Phys. 2013, 15, 13790. (32) Berger, T.; Lana-villarreal, T.; Go, R. J. Phys. Chem. C 2007, 111, 9936–9942. ang, .; Zakeeruddin, S. M.; Grätzel, M. J. Phys. (33) Cao, F.; Oskam, G.; Searson, P. C.; Stipkala, J. M.; Heimer, T. A.; Farzad, F.; Meyer, G. J. J. Phys. Chem. C 1995, 99, 11974–11980. (34) Meekins, B. H.; Kamat, P. V. ACS Nano 2009, 3, 3437–3446. (35) Berger, T.; Monllor-Satoca, D.; Jankulovska, M.; Lana-Villarreal, T.; Gómez, R. ChemPhysChem 2012, 13, 2824–2875. (36) Houng, B.; Liu, C. C.; Hung, M. T. Ceram. Int. 2013, 39, 3669–3676. (37) Jacques I. Pankove. Optical Processes in Semiconductors; 1975th ed.; Dover Publications, New York, 1975. (38) O’Regan, B.; Moser, J.; Anderson, M.; Graetzel, M. J. Phys. Chem. 1990, 94, 8720–8726. (39) Fabregat-santiago, F.; Mora-sero, I.; Bisquert, J. J. Phys. Chem. B 2003, 107, 758–768. (40) Bisquert, J. Phys. Chem. Chem. Phys. 2003, 5, 5360. (41) Kytin, V. G.; Bisquert, J.; Abayev, I.; Zaban, A. Phys. Rev. B 2004, 70, 193304. (42) Burstein, E. Phys. Rev. 1954, 93, 632–633. (43) Gibbs, Z. M.; LaLonde, A.; Snyder, G. J. New J. Phys. 2013, 15, 75020. (44) Sarkar, a.; Ghosh, S.; Chaudhuri, S.; Pal, A. K. Thin Solid Films 1991, 204, 255–264. 36 (45) Kim, S. S.; Park, J.; Kim, S. S.; Lee, Y.; Ahn, S.; Cho, J.; Ju, M.; Lakshminarayan, N.; Ai, V.; Yi, J. Vaccum 2014, 108, 39–44. (46) Hamberg, I.; Granqvist, C. G.; Berggren, K. F.; Sernelius, B. E.; Engström, L. Phys. Rev. B 1984, 30, 3240–3249. (47) Kamat, P. V.; Dimitrijevic, N. M.; Nozik, a. J. J. Phys. Chem. 1989, 93, 2873–2875. (48) Reddy, K. M.; Manorama, S. V; Reddy, A. R. Mater. Chem. Phys. 2002, 78, 239–245. (49) Valencia, S.; Marin, J. M.; Restrepo, G. Open Mater. Sci. J. 2010, 4, 9–14. (50) Welte, A.; Waldauf, C.; Brabec, C.; Wellmann, P. J. Thin Solid Films 2008, 516, 7256– 7259. (51) Tang, H.; Prasad, K.; Sanjinès, R.; Schmid, P. E.; Lévy, F. J. Appl. Phys. 1994, 75, 2042– 2047. (52) Thulin, L.; Guerra, J. Phys. Rev. B - Condens. Matter Mater. Phys. 2008, 77, 1–5. (53) Toyoda, T.; Tsuboya, I. Rev. Sci. Instrum. 2003, 74, 782–784. (54) Asahi, R.; Taga, Y.; Mannstadt, W.; Freeman, A. J. Phys. Rev. B 2000, 61, 7459–7465. (55) Liu, Y.; Jennings, J. R.; Zakeeruddin, S. M.; Grätzel, M.; Wang, Q. J. Am. Chem. Soc. 2013, 135, 3939–3952. (56) Enright, B.; Fitzmaurice, D. J. Phys. Chem. 1996, 100, 1027–1035. (57) Bard, A. J.; Bocarsly, A. B.; Fan, F. R. F.; Walton, E. G.; Wrighton, M. S. J. Am. Chem. Soc. 1980, 102, 3671–3677. (58) Schlichthörl, G.; Huang, S. Y.; Sprague, J.; Frank, a J. J. Phys. Chem. B 1997, 101, 8141– 8155. (59) Lyon, L. A.; Hupp, J. T. J. Phys. Chem. 1995, 99, 15718–15720. 37 Chapter 3 Charge Distribution in Nanostructured TiO2 Photoanode Determined by Quantitative Analysis of the Band Edge Unpinning Adapted with permission from: Charge Distribution in Nanostructured TiO2 Photoanode Determined by Quantitative Analysis of the Band Edge Unpinning, Dhritabrata Mandal and Thomas W. Hamann, ACS Appl. Mater. Interfaces 2016, 8, 419−424. Copyright 2016 American Chemical Society. 3.1 Introduction Mesoporous wide bandgap semiconductor electrodes have been widely studied for solar energy conversion, especially in the context of dye-sensitized solar cells (DSSC) and related systems.1,2 The conduction band (CB) edge position, trap state distribution and electron concentration are arguably the most important properties for detailed understanding of the various charge transfer processes associated with them.3 The mesoporous anatase titanium oxide (TiO2) electrodes are generally used as the photoanode in DSSCs. Anatase TiO2 crystal is constructed with chains of distorted TiO6 octahedra. It has a tetragonal lattice with four TiO2 units per unit cell. The CB is formed predominantly by Ti-3d orbital, with dxy states at the lower edge and the valence band is formed mainly by O-2p states.4–9 These TiO2 photoanodes are known to have large density of electroactive states energetically located below the CB (Figure 3-1) and may participate into the interfacial charge transfer processes.3 The spatial location and nature of these trap states have been debated. They are believed to be present predominantly on the surface of the nanoparticle (NP) although trapping of electrons in the bulk nanoparticle is also reported.10–13 The nature of these trap states is also not clear yet. Localization of excess electron at the Ti4+ centers in the 38 bulk or on the surface are generally considered as trapped electrons and the trapped states are frequently described as Ti3+ centers. Valentin et al. reported electronic structure calculations based on density functional theory (DFT) and suggested the presence of undercoordinated Ti 5c3+ which can potentially create a localized electronic center.14,15 Distribution of inter-band trap states below the CB have been reported in many n-type semiconductors like TiO2, SrTiO3 even in the single crystals.16,17 The inherent n-type conductivity in these metal oxides is the result of oxygen vacancies in their crystal structures.18 Missing of one lattice oxygen generate three undercoordinated (five-fold) Ti centers. Oxygen deficiency in the TiO2 crystal may result in the localization of an excess electron at a Ti site resulting in the formation of Ti3+ centers.13 On the surface, Ti3+ surface cations were reported to interact with OH ions and water molecules forming Ti3+OH bonds which create some electronic states below the CB. Hydroxyl groups were suggested to favor the localization of the Ti-3d states.14 When a proton is bound to a neighboring O center (Ti−OH species), a localized extra electron in 3d orbital of Ti center was calculated to be ~ 180 mV more stable than the scenario when this extra was electron fully delocalized on a 3dxy CB state.14 It is still a topic of discussion whether the “trapped electron” resides on a single Ti center or on few Ti centers. Concentration of the electrons in each of these states, i.e. CB, surface trap and bulk trap, changes differently with the displacement of the Fermi level. Determining their electron distributions is thus challenging but needed because they influence carrier transport and recombination differently.3,19,20 These distributions are known to depend on the film morphology,21 crystallinity,22 surface area23,24 and contacting solvent25,26. However, absence of any reliable method to determine the band energetics in nanostructured semiconductor photoanodes has also hindered an accurate determination of the distribution as well as the nature 39 of the various trap states. Using both spectroelectrochemical and electrochemical techniques, we have successfully determined the electron densities in each of these states along with the spatial location of the trap states.27,28 A spectroeletrochemical method is described in Chapter 2 which helped us to determine the conduction band edge from the apparent shift in the band gap transition, known as Burstein-Moss shift (B-M shift), in response to an applied bias.27 A very interesting phenomenon observed in this work was the upward shift or unpinning of the CB under negative bias which is generally ignored while studying the nanostructured photoanode and the charge transfer processes related to it. Figure 3-1 (Left) Energy vs. crystal momentum diagram of a semiconductor with an indirect band gap. Electron transition from the highest-energy state in the valence band to the lowestenergy state in the conduction band is not possible without a change in momentum. The transition occurs via the absorption or emission of a phonon, which provides the required momentum change during the transition. Inter-band localized electronic states bellow the conduction band, which trap electrons, are shown. (Right) Depiction of the upward shift of the energetics in TiO2 nanoparticle film. ECB represents the magnitude of the upward shift of the conduction band energy. In this chapter, we utilized spectroelectrochemical measurements to determine the magnitude of the band edge unpinning in the nanostructured TiO2 photoanode and its effect on the overall 40 charge distribution. From these results, we were also able to accurately calculate the total trap state density and estimate the spatial location of the localized trap states in TiO2. 3.2 Electrochemical techniques for determining charge distribution Standard electrochemical techniques like cyclic voltammetry (CV),29 electrochemical impedance spectroscopy (EIS),30 potential step-chronoamperometry (charge insertion- extraction)31,32 have been established as effective methods for the determination of the DOS in nanostructured metal oxide photoanodes. They are based on the measurement of the variation of the electron density (dn) with displacement of the Fermi Level (dEF) caused by changing the applied potential (dE). In this chapter, E represents energy and E (= E/q) represents the electrochemical potential. Any variation of the Fermi level has homogeneous effect on the electron concentration throughout the film as these photoanodes are made of a network of small sized nanoparticles, generally 10-20 nm in diameter, and is surrounded homogeneously by electrolyte.29,33,34 In the absence of significant IR drop along the nanoparticle network, dE = dEF is well accepted. The distribution of electrons is depicted by its chemical capacitance which describes the capability of the system to store or release additional electrons during a variation of chemical potential.33 Change in the Fermi level, causes a change in the electron density and hence in the chemical potential of the electrons. Thus, the chemical capacitance can be written as the variation of the electron density as a function of Fermi level,33 Cμ  q 41 dn dEF (1) 3.2.1 Cyclic voltammetry Cyclic voltammetry has been established as an effective method to determine the chemical capacitance (i.e. the distribution of the stored electrons) at different Fermi levels. Under cyclic voltammetric scan, the voltage-injected electrons distribute through the nanoparticle network until their chemical potential equalizes and the Fermi level becomes homogeneously distributed throughout the porous film. This capacitance is written as C = dQ/dE which gives us J = l(1p)sC, where J is the current density, l is the film thickness, p is the porosity of the mesoporous photoanode and s is the scan rate.29 Figure 3-2 shows the cyclic voltammogram and the calculated C of a mesoporous TiO2 photoanode. A prominent peak appears (region A in Figure 3-2) at around -0.8 V vs. SCE, which is identified as monoenergetic trap states.3,35 At more negative potentials (region B in Figure 3-2) an exponential distribution of density of states (DOS) approaching the conduction band was observed which is given by the expression,33,36 g Trap (E)  N Trap k bTC exp (qE  ECB ) k bTC (2) where NTrap is the total density of exponentially distributed trap states, TC is a parameter having the unit of temperature which implies the depth of the exponentially distributed trap states. Another interesting feature is the deviation from exponential distribution that appears at higher negative potential (region C) which will be discussed in details later. The CV plot and the onset of this capacitance increase (point x in Figure 3-2) is sometimes used to estimate the conduction band position.37–39 However, we would like to point out that, along with the electron distribution, this onset also depends on the conductivity of film i.e. the electron diffusion through the nanoparticle network.29,40 TiO2, along with other wide band gap semiconductors used in DSSCs, behaves as an insulator in the absence of sufficient reductive bias. When the conductive substrate 42 (which holds the porous film) is biased, initially a gradient of the Fermi level will exist. Then, the electrons spread by trapping-detrapping process from localized states into CB or/and hopping between trap sites until a homogeneous occupancy is attained though the film. Now, trappingdetrapping method requires certain electron density in the CB and the hopping process needs large enough density of electrons in the trap states for fast quantum-mechanical tunneling transition.23 Figure 3-2. Cyclic voltammogram plot (left) and chemical capacitance plot (right) of TiO2 film in aq. 0.2 M KCl solution at pH 12.8. Scan rate was 50 mV / s. The chemical capacitance (C) was calculated using the relation J = l(1-p)sC, where J is the current density, l (= 6.95 m) is the electrode thickness, p (= 0.71) is the porosity of the TiO2 film and s is the scan rate. The peak at approximately  0.75 V vs. SCE is referred to as monoenergetic trap state distribution and the wave at increasing negative potentials corresponds to the exponential distribution of trap states. Accurate and homogeneous distribution of chemical capacitance will only be determined when the charge equilibration is fast enough. A very common observation of this CV analysis is that, under very fast scan, the onset moves to more negative potential suggesting that this equilibration has not been reached. However, a slower scan is not always better. The 43 characteristic capacitance filling disappears when the scan rate is slow enough where the current loss through Faradaic processes dominates.29 Proper range of scan rates need to be chosen where the Fermi level equilibration occurs through the nanoparticle network and current loss through the Faradaic processes are negligible. 3.2.2 Electrochemical impedance spectroscopy Electrochemical impedance spectroscopy (EIS) is another widely used technique to study the capacitive processes in the semiconductor.30,41 A transmission line model needs to be used in case of porous photoanodes instead of simple Randal circuit. The equivalent circuit suitable for this mesoporous photoanode film is shown in Figure 3-3. This circuit illustrates the internal distribution of chemical capacitance in response to the modulated small perturbation of the steady state bias. Distribution of C with energy, determined by EIS using transmission line model, is shown in Figure 3-4. Rs is series resistance accounting for the transport resistance of the conductive substrate, Rt is the electron transport resistance through photoanode nanostructure, Rr is the resistance at the nanoparticle/electrolyte interface. Simple Randal circuit is still being used by some researchers to fit the impedance data. However, it should be taken into account that Randal circuit can only be applicable when this nanoparticle network is sufficiently conductive i.e. Rt is negligible. To study this porous semiconductor photoanode in a wide range of bias, transmission line model needs to be used. Also, at low cathodic potentials, measured capacitance is attributed to the substrate/electrolyte interface and should not be confused with NP/electrolyte interface.42,43 Moreover, the apparent onset in the capacitance plot is not the flat band potential; it is the point where the chemical capacitance begins to dominate the measured capacitance. 44 Figure 3-3 Equivalent circuit based on transmission line model for fitting Electrochemical Impedance data of porous photoanode. The fit parameters are described in the text. Figure 3-4 Plot of capacitance (C) of TiO2 film measured by electrochemical impedance spectroscopy. Electrolyte was 0.2 M aq. KCl solution at pH 12.8. 45 3.2.3 Charge insertion/extraction method Charge insertion/extraction measurement provides a direct method to determine the total electrons stored in the photoanode at a certain Fermi level.31,32 In this method, the electrons are inserted into the photoanode under a bias, e.g. illumination or potentiostatic bias. Then, after the Fermi level stabilizes and all the states below that energy level are filled, the electrons are extracted by removing the previous bias or by applying discharging (oxidative) bias. Using Chronoamperometric setup, the current (i) is measured with time (t) and the total charge, Q, is calculated by integrating the area under current-time curve (Q = i.t). Figure 3-5 shows a typical current-time plot of charge insertion and extraction under potentiostatic bias. The charge accumulated under that negative (reductive) bias can be calculated from the current decay curve in both charging region and discharging region. Ideally, they should provide the same value. However, in presence of any Faradaic process (e.g. H2 evolution) or any irreversible ion insertion process, the extracted charge is lower than inserted charge.44 Charge calculated during the extraction process is more reliable to give information about the amount of electrons stored at any energy level. We can write, EF Q  q  N ( E )dE (3) E0 where N(E) is the electron density as a function of the Fermi level (EF = E), E0 is the discharging (oxidative) bias or the solution potential. Thus, we can directly get the distribution of the density of states from this relation, N (E)  46 1 dQ q dE (4) The above methods are effective to calculate the overall electron density in the mesoporous photoanodes. However, they cannot separate the density of delocalized CB electrons from the trapped electrons and, also, are unable to differentiate the surface trapped electrons from the electrons trapped in the bulk. Each of these states responds to the electron transfer mechanisms very differently and, therefore, needs to be determined separately.3 Figure 3-5 Typical current vs. time plot measured during charge insertion/extraction process under potentiostatic. The measurement was performed with mesoporous TiO2 film in contact with aq. 0.2 M KCl solution at pH 12.8. The reductive bias was  1.045 V and the oxidative bias was + 0.455 V vs. SCE. 3.3 Experimental 3.3.1 TiO2 photoanode preparation The transparent mesoporous anatase TiO2 electrode was prepared by spreading HT/SP TiO2 paste (Solaronix) on fluorine-doped tin oxide (FTO) conducting glass (Hartford Glass) by the doctor-blade method, followed by sintering at 500°C for 15 min in air. The thickness of the TiO2 film was measured by a Dektak3 Surface Profiler and found to be 6.950.05 μm. The porosity, p, 47 of the mesoporous TiO2 film was also determined. 15 TiO2 films (geometric area 64 mm2 and thickness 6.95 m) were collected from the FTO by scraping out with a sharp blade. The total mass of the collected TiO2 material was measured. Total volume of the 15 films was then calculated which is 64 mm2  6.95 m 15 and the mass of the solid TiO2 material without any pores was calculated from the density of anatase material, 3.93 g / cm-3. Porosity is calculated using the equation, p = 1 – (measured mass of porous film / calculated mass of solid material). 3.3.2 Electrolyte preparation Aqueous electrolyte solutions were prepared by dissolving recrystallized potassium chloride into Millipore water (18.2 M-cm). pH of the electrolytes were adjusted to 12.8 and 2.0 by adding KOH and HCl, respectively. 3.3.3 Measurements In spectroelectrochemical measurements, the transmittance (%T) of the film was measured by a Perkin Elmer Lamda 35 UV-Vis spectrophotometer under an external bias provided by a potentiostat (-Autolab, Metrohm). A three electrode electrochemical cell, made by customizing a commercial spectroscopic cuvette (Quartz Spectrophotometer Cell, Starna Cells, Inc), was placed into the UV-Vis spectrophotometer such that the TiO2 film remained perpendicular to the optical path as shown in Figure 2-2. A platinum wire (diameter 1.5 mm) was used as counter electrode and a commercial “Leakless Miniature” Ag/AgCl reference electrode (EDAQ ET072) was used. All electrochemical measurements done here were three-electrode measurements. The TiO2 electrode was sealed to an opening of an electrochemical cell with a Viton O-ring. TiO2 films were aged overnight in contact with the electrolyte solutions and N2 gas was purged into the aqueous electrolytes for 30 minutes, prior to any measurement. During charge extraction 48 method, the electrode was charged for 5 minutes followed by applying a positive bias (+0.455 V vs. SCE) to extract all the charges stored during charging. The current was measured against time and the total charge was calculated by integrating the area under the current vs. time curve. Prior to the cyclic voltammetry scan, a significant positive potential (+0.455 V vs. SCE) was applied to the film to discharge all the electrons from the film. Electrochemical Impedance (EIS) measurements were done in the dark. At each potential, the frequencies were equally spaced in logarithmic steps between 2×10-2 to 1×105 Hz. The counter electrode used in the spectroelectrochemical measurements was a platinum wire and a high surface area platinum mesh was used in electrochemical experiments. A homemade Ag/AgCl reference electrode was used during all the measurements. Its potential was measured against a commercial saturated calomel electrode (SCE); all the applied potentials are reported vs. SCE. The TiO2 films were aged overnight in contact with the electrolyte solutions prior to any experiments. The temperature of the electrochemical cells was 220.5 ⁰C during the measurements. During spectroelectrochemical measurements, each film was stabilized at an applied bias for 5 minutes and %T of the electrode was measured. A control experiment was done with bare FTO electrode to confirm that the FTO has no contribution into the change in %T with varying potential. Stability of the film as well as reproducibility of the results was checked both spectroscopically as well as electrochemically to make sure that there is no irreversible change of the film during the measurements. The difference absorbance spectrum, A, is the absorbance change under external bias. At first, the absorbance was measured under a significantly positive potential to discharge all the stored electrons from the film. Then a negative potential was applied to increase the electron population and the absorbance was measured. Absorbance measured at positive potential was subtracted from the absorbance at negative bias and thus A was calculated. 49 3.4 Results and discussions 3.4.1 Accurate energy distribution of trap states in mesoporous TiO2 photoanode The upward shift of ECB when the applied potential is increased to more negative values was shown in Chapter 2. This shift was attributed to the potential drop across the Helmholtz layer.45– 47 Change in the electric field of the Helmholtz layer occurs due to the accumulation of electrons in electrically active states on the surface or in close vicinity of the surface of mesoporous TiO2 films. In order to calculate the magnitude of the upward shift of conduction band, ECB, an initial position of the conduction band, E*CB, needs to be determined. Figure 3-6 shows a plot of EG determined as a function of E. A linear extrapolation of this plot to the intercept, corresponding to no bandgap shift, produced values of –1.17 V vs. SCE at pH 12.8 and –0.55 V at pH 2.0 for E*CB. Though no theoretical background of this linear behavior can be found, it gives an estimate of the initial value for CB edge potential. The expected Nernstian shift of the conduction band with pH was observed, which can be expressed by: E*CB / q  0.435V(vs.SCE)  0.058  pH (5) This Nernstian behavior is well known for metal oxide semiconductors in aqueous electrolytes.48 50 Figure 3-6. The position of the conduction band edge potential determined at pH 12.8 and 2.0. Knowledge of the band edge shift, ECB/q (= E*CB/q  ECB/q), is required for accurate determination of the charge distribution by electrochemical methods. Now, equation 2 is expected to provide the actual distribution of localized states only if ECB is fixed, which is not the case. At more negative applied potentials, the parameter ECB/q, i.e. the potential drop across the Helmholtz layer, needs to be included. Thus, equation 2 should be rewritten as: g ETS ( E )  N ETS (qE  ECB  ECB ) exp k bTC k bTC (6) Therefore, C is expected to increase exponentially with (E  ECB/q) instead of E, as depicted in Figure 3-7. Inclusion of ECB/q causes two major modifications to the plot of the chemical capacitance. 1) The deviation of C from the exponential behavior at high negative potentials disappears and 2) the exponential slope as well as the value of TC changes significantly. 51 At lower potentials, the magnitude of the potential drop is small and there is an overlap between the tail of the Gaussian distribution and the exponential distribution of trap states which may be the reason behind the exponential like increase of C. However, at those potentials the lnC vs. E plot does not necessarily represent the actual distribution of the localized states as the corrected TC (682 K) is significantly different than the values of TC (975 – 1667 K) calculated from the lnC vs. E plot without considering band edge unpinning. Therefore, the common practice of interpreting electrochemical data assuming a fixed band edge position without the necessary consideration of the band edge shift gives rise to a significant error in the determination of the electron distribution in nanostructured semiconductor photoanodes. The monoenergetic trap states shows a peak at ~0.8 V vs. SCE in CV scan and is not observed in the potential range where the chemical capacitance (Figure 3-7) is measured by EIS measurements. Figure 3-7. Plot of the chemical capacitance (C) as a function of applied potential, E, (bottom axis, red circles) and applied potential which is corrected to include the effect of band edge unpinning (ECB/q) as shown in equation 7 (top axis, blue triangles). 52 Another example where the band edge shift needs to be considered is in the interpretation of the visible absorbance (A) of the TiO2 photoanodes for the spectroelectrochemical analysis of the charge transfer processes in DSSCs.49,50 The calculated nCB showed an excellent linear relationship with the measured A in accordance with the Beer-Lambert law: A   CBl (1  p)M CB (7) where CB is the extinction coefficient of delocalized CB electrons and MCB is the molar concentration (mole L-1) of CB electrons.27 However, participation of the trapped electrons in the absorbance spectra has been invoked by many researchers, primarily because of the close resemblance of the slope of absorbance (A) vs. E with the slope of nTrap vs. E as well as the slope of nTotal, total concentration of the electrons in the film, vs. E.13 Density of both free and trapped electrons increase with negative potential, in addition to conduction band edge movement. Explaining the correlation between the intra-bandgap trapped electrons and the measured absorbance based on the resemblance of their slopes is not straightforward and might be misleading. Assuming a fixed conduction band, the absorbance of the CB electrons should increase exponentially with E (= EF/q) according to the following equation, nCB  N CB exp (EF  ECB ) k bT (8) Thus, a semi-logarithmic plot of A vs. E, should be linear with a slope of 1/kbT = 39.2 V-1. However, the experimental slope of ln(A780nm) vs. E was found to be ~11 V-1, as shown in Figure 3-8. Because of the upward movement of ECB, the magnitude of change in (ECB/q – E) is actually smaller than the change in the applied bias, E. Thus the slope of ln(A780nm) vs. E is much lower 53 than the ideal slope 39.2 V-1. Considering the band edge movement, however, equation 9 can be rewritten as: nCB  N CB exp (EF  ECB  ECB ) k bT (9) Combining equations 7 and 9, the following expression is produced to describe the expected dependence of absorbance with applied potential: ln( A)  ln( CBEl (1  p) N CB )  ECB (qE  ECB )  k bT k bT (10) Figure 3-8 shows excellent agreement between the slope of ln(A780nm) vs. (E  ECB/q) and the ideal slope of 39.2 V-1 (1/kT). Figure 3-8. Plot of absorbance of TiO2 at 780 nm as a function of applied potential (bottom axis, red circles) and applied potential which is corrected to include the effect of band edge unpinning (top axis, blue triangles). 54 3.4.2 Spatial location of the trap states The band edge shift can also provide information about the spatial location of the trap states. Trap states are reported to be present predominantly on the surface of the nanoparticle (NP)11 although trapping of electrons in the bulk nanoparticle is also reported.10 The band edge unpinning suggests that the nanostructured TiO2 film has a high density of surface states which can accumulate electrons under negative bias. To maintain the charge neutrality, this electron accumulation is accompanied by positive charges, i.e. the cations in the electrolyte, at the semiconductor/electrolyte interface which causes potential difference in the Helmholtz layer. The potential drop across the Helmholtz layer i.e. the upward shift of the conduction band, ECB/q, and the surface charge concentration (nS) can be related by the following equation: ECB / q  qnS CH (11) Here CH is the Helmholtz layer capacity which was determined by measuring the capacitance of a flat TiO2 film by electrochemical impedance spectroscopy and found to be ~6 Fcm-2. The total concentration of electrons in the film can be expressed as nTotal = nCB + nTrap. Charge extraction was performed to determine nTotal as a function of potential, and nCB was determined from the B-M shift, thus allowing calculation of nTrap. The total concentration of trapped electrons can be expressed as the sum of surface trapped electrons and the bulk trapped electrons, nB. 55 Figure 3-9. Plot of the concentration of trapped electrons in the bulk (nB) and on the surface (nS) of TiO2 film in aq. 0.2 M KCl solution at pH 12.8. Extrapolation of the fitted plots to the conduction band edge gives the density of bulk trap states, NB, and surface traps, NS. Figure 3-10. Plot of electron density in the nanostructured TiO2 photoanode as a function of applied potential. The total trapped electron density (nTrap) is the sum of the electrons trapped on the surface (nS) and in the bulk (nB). 56 Figure 3-10 shows a plot of the concentration of electrons in surface traps calculated from equation 12 and bulk traps determined by subtracting the surface traps from the total traps. Both nS and nB increase with negative bias, but their slopes are different. NS, density of surface states, and NB, density of bulk traps, are found to be 2.4  1019 cm-3 and 1.4 1021 cm-3, respectively (Figure 3-9). The trap states are localized Ti3+(3d) states which can accept (trap) electrons and releasing (detrapping) this electrons needs some driving force. They are energetically located below CB edge. Above this edge, electrons are delocalized “free” electrons. The surface states are the localized sites on the surface or in close vicinity to surface those can trap electrons. Presence of under coordinated TiIV/TiIII centers or any terminal oxo/hydroxo sites have been observed on the surface of the nanoparticles.51,52 Additionally, electrons are also trapped in the bulk.48,53 When a negative (reductive) bias is applied to the photoanode the electron is first injected into the CB and transported through the delocalized CB states distributed uniformly in the bulk photoanode material. Thus, the availability of electrons to be captured by the trap centers in the bulk and on the surface is the same. The excess negative charge of these injected electrons in the photoanode must be neutralized by the equivalent charge of the cations. At low applied potentials, the trapped electron density is dominated by the electrons in surface traps. The negative charge on the surface can be compensated by electrolyte cations such as K + or H+ present in the aqueous solutions. Since the electrolyte contains ~1012 times more K+ than H+ in a pH 12.8 solution with 0.2 M KCl, we suggest K+ is the more likely surface charge compensating cation. Considering NS = 2.4 1019 cm-3, the area available for each compensating cation adsorbed on the NP surface is ~7 nm2. K+ ions has an effective ion radius of 1.38 Å54 which implies that sufficient space is present to accommodate the solvated K+ ions compensating the surface charge. As shown in Figure 3-10, at low reductive region, the trapped electron density is 57 dominated by the electrons in the surface traps because they are easily charge compensated by K+ and/or H+. More negative bias forces the H+ to insert into the nanoparticle to compensate the charge which involves bond length adjustment or/and breaking of Ti-O bonds. Another very interesting observation is, while the density of surface states does not change much with the increasing negative potential, the bulk trap states increases rapidly with an exponential dependence on the applied potential. We have explained it by the fact that the available sites to adsorb K+(solvated) is limited by their steric hindrance as well as columbic repulsion. Although K+ is unlikely to penetrate into the bulk and compensate the electron charge arising from trapping electrons in bulk states,55 H+ insertion into the TiO2 material under negative bias has been found to be reversible in the potential range used here.48,56 Under increasing negative potential, the bulk trapped electrons are therefore compensated by H+ insertion. This H+ can reside in the interstitial space in an O6 octahedral site or can form a covalent bond with a neighboring oxygen.57 Additionally, under reductive bias, proton insertion is known to create TiIII centers which trap electrons.48 With increasing negative potential, the bulk trap starts dominating the total trap density and at the negative end it has become an order of magnitude higher than the surface trap as shown in Figure 3-9. It can be explained by the ease of H+ insertion into anatase NPs and availability of more space as well as Ti-centers in the bulk than the surface (~ 6 times more Ti atoms in the bulk than on the surface). The total density of states of bulk traps below the CB edge is two orders of magnitude higher than the density of surface traps (NS = 2.4  1019 cm-3 and NB = and 1.4 1021 cm-3). A very interesting point needs to be discussed here is the “surprisingly” high densities of the trap states. The value of NS and NB translates to ~ 4e/nanoparticle and ~ 224e/nanoparticle, respectively. Berger et al. reported ~ 650e/nanoparticle based on the charge extraction method.13 The diameter (d) of the nanoparticle 58 (NP) used in their study was 20 nm and the diameter of the nanoparticles used in our study is 10 nm.13 Their findings are quite similar to the values we found for total trapped electron density. This huge number of electrons is unlikely to be captured by the defect states. Storage of electrons inside a nanoparticle may have different mechanisms. Though it is not the scope of this study, we can discuss some of the possibilities based on the findings and hypotheses in the literature. Hupp et al. showed intercalation of charge-compensating cations in the TiO2 NPs under electrochemically reduced bias.48,53,58 Interestingly, they observed reversible proton insertion into the NPs even under conditions where surface adsorption of protons is not plausible. It suggests two interesting points, 1) charge neutralization processes on the surface and in the bulk are very different and 2) insertion of electron in the NPs can occur even without the availability of protons in the aqueous solution like highly basic solution (pH 12.8 was used in our study). The high density of electrons stored and discharged during the charge insertion-extraction process can be explained by the processes observed in intercalation-based battery materials. A possible explanation behind the observed exponential increase in bulk trapped electron density with applied reductive bias is the higher driving force for electron insertion accompanied by proton intercalation. This process can electrochemically generate Ti(III) centers even in the absence of any prior “defect” centers. The proton bounds to a lattice oxygen and the inserted electron is localized on the adjacent Ti center. Point to be noted that extent of localization of this electron is a still a topic of debate.48 3.5 Conclusion A spectroelectrochemical method has been described that allows for the quantitative analysis of the band edge unpinning in nanostructured semiconductor photoanodes. The position of the conduction band edge in porous TiO2 photoanode was determined from the Burstein-Moss shift 59 observed under negative potential. The upward shift of the conduction band edge has been explained by the presence of large density of the electrons accumulated on the surface of nanoparticles causing potential drop across the Helmholtz layer. The magnitude of the band edge unpinning should be included while studying the nanostructured photoanode and the charge transfer processes related to it. Moreover, the trap states are found distributed across the surface as well as in the bulk and with increasing negative bias bulk traps dominate the overall chemical capacitance. Knowledge of the spatial as well as the accurate energy distribution of the trap states are required to study their participation into the electron-transfer processes. The trap states present on the surface or in close vicinity of the surface are more likely to participate in the surface mediated charge transfer processes than the bulk traps. Such detailed understanding is especially important in describing recombination reactions which generally limit the performance of DSSCs. 60 REFERENCES 61 REFERENCES (1) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Chem. Rev. 2010, 110, 6595– 6663. (2) Kay, A.; Cesar, I.; Grätzel, M. J. Am. Chem. Soc. 2006, 128, 15714–15721. (3) Ondersma, J. W.; Hamann, T. W. J. Am. Chem. Soc. 2011, 133, 8264–8271. (4) Zhu, T.; Gao, S.-P. J. Phys. Chem. C 2014, 118, 11385–11396. (5) Han, X.; Song, K.; Lu, L.; Deng, Q.; Xia, X.; Shao, G. J. Mater. Chem. C 2013, 1, 3736. (6) Dou, M.; Persson, C. J. Appl. Phys. 2013, 113, 83703. (7) Zhang, J.; Zhou, P.; Liu, J.; Yu, J. Phys. Chem. Chem. Phys. 2014, 16, 20382–20386. (8) Patrick, C. E.; Giustino, F. J. Phys. Condens. Matter 2012, 24, 202201. (9) Baldini, E.; Chiodo, L.; Dominguez, A.; Palummo, M.; Moser, S.; Yazdi-Rizi, M.; Auböck, G.; Mallett, B. P. P.; Berger, H.; Magrez, A.; Bernhard, C.; Grioni, M.; Rubio, A.; Chergui, M. Nat. Commun. 2017, 8, 13. (10) Westermark, K.; Henningsson, A.; Rensmo, H.; Södergren, S.; Siegbahn, H.; Hagfeldt, A. Chem. Phys. 2002, 285, 157–165. (11) Kopidakis, N.; Neale, N. R.; Zhu, K.; Van De Lagemaat, J.; Frank, a. J. Appl. Phys. Lett. 2005, 87, 1–3. (12) Berger, T.; Lana-Villarreal, T.; Monllor-Satoca, D.; Gómez, R. J. Phys. Chem. C 2007, 111, 9936–9942. (13) Berger, T.; Anta, J. a; Morales-Flórez, V. J. Phys. Chem. C 2012, 116, 11444–11455. (14) Di Valentin, C.; Pacchioni, G.; Selloni, A. J. Phys. Chem. C 2009, 113, 20543–20552. (15) Di Valentin, C.; Pacchioni, G.; Selloni, A. Phys. Rev. Lett. 2006, 97, 166803. (16) Vandermolen, J. J. Electrochem. Soc. 1980, 127, 324. (17) Siripala, W. J. Electrochem. Soc. 1981, 128, 2491. (18) Gutierrez, C.; Salvador, P. J. Electroanal. Chem. Interfacial Electrochem. 1982, 138, 457–463. (19) Peter, L. Acc. Chem. Res. 2009, 42, 1839–1847. (20) Meekins, B. H.; Kamat, P. V. ACS Nano 2009, 3, 3437–3446. (21) Mercado, C. C.; Knorr, F. J.; Mchale, J. L.; Usmani, S. M.; Ichimura, A. S.; Saraf, L. V. J. 62 Phys. Chem. C 2012, 116, 10796–10804. (22) Mercado, C.; Seeley, Z.; Bandyopadhyay, A.; Bose, S.; McHale, J. L. ACS Appl. Mater. Interfaces 2011, 3, 2281–2288. (23) J. Frank, A.; Kopidakis, N.; Lagemaat, J. Van De. Coord. Chem. Rev. 2004, 248, 1165– 1179. (24) Zhu, K.; Kopidakis, N.; Neale, N. R.; van de Lagemaat, J.; Frank, A. J. J. Phys. Chem. B 2006, 110, 25174–25180. (25) Knorr, F. J.; Mercado, C. C.; McHale, J. L. J. Phys. Chem. C 2008, 112, 12786–12794. (26) Schwanitz, K.; Weiler, U.; Hunger, R.; Mayer, T.; Jaegermann, W. J. Phys. Chem. C 2007, 111, 849–854. (27) Mandal, D.; Hamann, T. W. Phys. Chem. Chem. Phys. 2015, 17, 11156–11160. (28) Mandal, D.; Hamann, T. W. ACS Appl. Mater. Interfaces 2016, 8, 419–424. (29) Fabregat-Santiago, F.; Mora-Seró, I.; Garcia-Belmonte, G.; Bisquert, J. J. Phys. Chem. B 2003, 107, 758–768. (30) Fabregat-Santiago, F.; Garcia-Belmonte, G.; Bisquert, J.; Zaban, A.; Salvador, P. J. Phys. Chem. B 2002, 106, 334–339. (31) Bailes, M.; Cameron, P. J.; Lobato, K.; Peter, L. M. J. Phys. Chem. B 2005, 109, 15429– 15435. (32) Wang, H.; He, J.; Boschloo, G.; Lindström, H.; Hagfeldt, A.; Lindquist, S.-E. J. Phys. Chem. B 2001, 105, 2529–2533. (33) Bisquert, J. Phys. Chem. Chem. Phys. 2003, 5, 5360. (34) Kytin, V. G.; Bisquert, J.; Abayev, I.; Zaban, A. Phys. Rev. B 2004, 70, 193304. (35) Boschloo, G.; Fitzmaurice, D. J. Phys. Chem. B 1999, 103, 2228–2231. (36) Peter, L. M. J. Phys. Chem. C 2007, 111, 6601–6612. (37) Le Viet, A.; Jose, R.; Reddy, M. V.; Chowdari, B. V. R.; Ramakrishna, S. J. Phys. Chem. C 2010, 114, 21795–21800. (38) Jankulovska, M.; Berger, T.; Lana-Villarreal, T.; Gómez, R. Electrochim. Acta 2012, 62, 172–180. (39) Berger, T.; Monllor-Satoca, D.; Jankulovska, M.; Lana-Villarreal, T.; Gómez, R. ChemPhysChem 2012, 13, 2824–2875. (40) Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S.-E. J. Phys. Chem. B 1997, 101, 7710–7716. 63 (41) Bisquert, J.; Garcia-Belmonte, G.; Fabregat-Santiago, F.; Ferriols, N. S.; Bogdanoff, P.; Pereira, E. C. J. Phys. Chem. B 2000, 104, 2287–2298. (42) Zaban, A.; Meier, A.; Gregg, B. a. J. Phys. Chem. B 1997, 101, 7985–7990. (43) Bisquert, J.; Garcia-Belmonte, G.; Fabregat-Santiago, F. J. Solid State Electrochem. 1999, 3, 337–347. (44) van de Krol, R.; Goossens, A.; Schoonman, J. J. Phys. Chem. B 1999, 103, 7151–7159. (45) Cao, F.; Oskam, G.; Searson, P. C.; Stipkala, J. M.; Heimer, T. a; Farzad, F.; Meyer, G. J. J. Phys. Chem. 1995, 99, 11974–11980. (46) Bard, A. J.; Bocarsly, A. B.; Fan, F. R. F.; Walton, E. G.; Wrighton, M. S. J. Am. Chem. Soc. 1980, 102, 3671–3677. (47) Schlichthörl, G.; Huang, S. Y.; Sprague, J.; Frank, a J. J. Phys. Chem. B 1997, 101, 8141– 8155. (48) Lyon, L. A.; Hupp, J. T. J. Phys. Chem. B 1999, 103, 4623–4628. (49) Anderson, A. Y.; Barnes, P. R. F.; Durrant, J. R.; O’Regan, B. C. J. Phys. Chem. C 2010, 114, 1953–1958. (50) Barnes, P. R. F.; Miettunen, K.; Li, X.; Anderson, A. Y.; Bessho, T.; Gratzel, M.; O’Regan, B. C. Adv. Mater. 2013, 25, 1881–1922. (51) Vittadini, A.; Casarin, M.; Selloni, A. Theor. Chem. Acc. 2007, 117, 663–671. (52) Nunzi, F.; Mosconi, E.; Storchi, L.; Ronca, E.; Selloni, A.; Grätzel, M.; De Angelis, F. Energy Environ. Sci. 2013, 6, 1221. (53) Lyon, L. A.; Hupp, J. T. J. Phys. Chem. 1995, 99, 15718–15720. (54) Heyrovská, R. Chem. Phys. Lett. 1989, 163, 207–211. (55) Kavan, L.; Kratochvilová, K.; Grätzel, M. J. Electroanal. Chem. 1995, 394, 93–102. (56) Halverson, A. F.; Zhu, K.; Erslev, P. T.; Kim, J. Y.; Neale, N. R.; Frank, A. J. Nano Lett. 2012, 12, 2112–2116. (57) Zhang, J.; Steigerwald, M.; Brus, L.; Friesner, R. a. Nano Lett. 2014, 14, 1785–1789. (58) Lemon, B. I.; Hupp, J. T. J. Phys. Chem. 1996, 100, 14578–14580. 64 Chapter 4 Absorbance Features of Mesoporous TiO2 Electrode under Electron Accumulation 4.1 Introduction Mesoporous TiO2 films are widely studied for their potential in photovoltaic, photocatalytic and electrochromic applications. The behavior of the charge carriers in these films under external bias has been an important field of study. Electrochemical techniques are used to study the populated states as well as the associated charge transfer processes which rely on the measured relationships between potential, current, and charge. Furthermore, spectroscopic techniques are often combined with the electrochemical measurements to get further insight into the electrochemical processes that cannot be obtained from the electrochemical experiments alone. In spectroelectrochemical methods, the charge carriers in the bulk electrode or at the electrodeelectrolyte interface are spectroscopically monitored as a function of applied potential. The measured absorbance or reflectance is then translated into the carrier density provided the corresponding extinction coefficient is known. However, two major challenges in this process are 1) assigning a certain absorbing species that contribute to the measured spectra and 2) determining the extinction coefficient of the species of interest. Fitzmaurice et al. studied the absorbance of the mesoporous TiO2 film under accumulation conditions.1,2 In this method, TiO2 films, made of nanoparticles (NPs) with an average diameter of 15 nm, were deposited on a transparent conductive substrate. Due to the good electronic connectivity among nanoparticles and equipotential environment surrounding the nanoparticles, the Fermi level (EF) within the TiO2 nanoparticle electrode can be controlled by applying an external bias to the conductive substrate.3–6 The absorbance of the film was measured while 65 applying large negative bias during the cyclic voltammetric scan. The measured absorption at a certain applied potential (E) was correlated to the electron population in the semiconductor film at the Fermi level EF which is equal to qE (q = charge of an electron). Here, E represents energy and E (=E/q) represents the electrochemical potential. They observed that the absorbance in the visible region was increased while raising the Fermi level which was attributed to the absorption by the increasing population of the electrons. The characteristic of the absorbance spectra shows resemblance with the free carrier absorption generally described as Drude absorption.7 The density of free electrons, flat band position and the extinction co-efficient were determined from this absorbance data assuming a Schottky junction model. The key idea of this model was the formation of an accumulation layer within each nanoparticle as the applied potential was shifted negative of the flat-band potential. Assumption of a flat interface and Schottky junction is unlikely in these small nanoparticles and the validity of the model has been debated.8–10 Moreover, the nature of the absorbing species has also been a long debated issue. Large density of trapped electron is known to be present in the mesoporous TiO2 electrodes. Many researchers have assigned this absorption to the trapped electrons primarily because of the close resemblance between the slope of absorbance vs. E and the slope of trapped electron density vs. E or the slope of total electron density vs. E as well.11–15 However, it is not that straightforward and might be misleading as both free and trapped electrons increase with negative potential. Additionally, conduction band edge shift, as described in Chapter 2, makes it difficult to determine the accurate slope. Thus analyzing localized electronic states, such as Ti(III) created under reduced conditions, from the measured absorption might lead us to erroneous interpretation. Moreover, extinction coefficient of the CB electrons (CB) needs to be known in various spectroscopic studies of the electron transfer processes in DSSCs. While performing 66 spectroscopic studies on various DSSC components other than the photoanode, the absorption contribution of the photoanode needs to be subtracted from the measured absorption to monitor the spectroscopic behavior of the desired component. The literature values of CB are typically determined by correlating the absorbance with the concentration of electrons in the TiO2 electrode measured either by chronoamperometry or titration. These techniques, however, measure the total concentration of electron in the TiO2 film, both trapped and free electrons. Considering that trapped electron concentration is orders of magnitude higher than the free CB electrons, as shown in Chapter 3, the measured extinction coefficients of free CB electrons are underestimated by one or two orders of magnitude. In this work, the absorbance characteristics of mesoporous TiO2 film under reductive bias have been revisited. Detailed spectroeletrochemical study was performed with nanostructured TiO2 photoanode films in contact with various electrolyte solutions. The extinction coefficient of the electrons has been determined by spectroelectrochemical method. 4.2 Experimental 4.2.1 TiO2 photoanode preparation The transparent mesoporous anatase TiO2 electrode was prepared by spreading HT/SP TiO2 paste (Solaronix) on fluorine-doped tin oxide (FTO) conducting glass (Hartford Glass) by the doctor-blade method, followed by sintering at 500°C for 15 min in air. The thickness of the TiO2 film was measured by a Dektak3 Surface Profiler and found to be 6.950.05 μm. The porosity, p, of the mesoporous TiO2 film was also determined. 15 TiO2 films (geometric area 64 mm2 and thickness 6.95 m) were collected from the FTO by scraping out with a sharp blade. The total mass of the collected TiO2 material was measured. Total volume of the 15 films was then 67 calculated which is 64 mm2  6.95 m 15 and the mass of the solid TiO2 material without any pores was calculated from the density of anatase material, 3.93 g / cm-3. Porosity is calculated using the equation, p = 1 – (measured mass of porous film / calculated mass of solid material). 4.2.2 Electrolyte preparation Aqueous electrolyte solutions were made by dissolving recrystallized salts into Millipore water (18.2 M-cm). Anhydrous acetonitrile (Sigma Aldrich) was used as received. Tetrabutylammonium hexafluorophosphate (TBAPF6), lithium hexafluorophosphate (LiPF6), lithium perchlorate (LiClO4) and sodium hexafluorophosphate (NaPF6) were recrystallized and dried under vacuum at 80 C for 48 hours prior to any use. All the aqueous and acetonitrile electrolytes contain 0.2 M inert salts i.e. KCl, TBAPF6, LiPF6, LiClO4, NaPF6. A Fischer Scientific Accumet pH meter was used to measure the pH of the aqueous electrolytes. pH were maintained by different types of buffer systems like citrate buffer (pH 6.2), imidazole buffer (pH 6.2 and pH 7.8) and phosphate buffer (pH 7.8). Drops of hydrochloric acid (HCl) and potassium hydroxide (KOH) were added to adjust the pH to the desired value. pH 12.8 was made with KOH only and pH 2.0 was made with HCl only. 4.2.3 Spectroelectrochemical measurement In spectroelectrochemical measurements, the transmittance (%T) of the film was measured by a Perkin Elmer Lamda 35 UV-Vis spectrophotometer under an external bias provided by a potentiostat (-Autolab, Metrohm). A three electrode electrochemical cell, made by customizing a commercial spectroscopic cuvette (Quartz Spectrophotometer Cell, Starna Cells, Inc), was placed into the UV-Vis spectrophotometer such that the TiO2 film remained perpendicular to the optical path as shown in Figure 2-2. A platinum wire (diameter 1.5 mm) was used as counter 68 electrode and a commercial “Leakless Miniature” Ag/AgCl reference electrode (EDAQ ET072) was used. The TiO2 films were aged overnight in contact with the electrolyte solutions prior to any experiment. The temperature of the electrochemical cells was 22  0.5 ⁰C during the measurements. During spectroelectrochemical measurements, each film was stabilized at an applied bias for 5 minutes and %T of the electrode was measured. A control experiment was done with bare FTO electrode to confirm that the FTO has no significant contribution into the change in %T with varying potential. Stability of the film as well as reproducibility of the results were checked both spectroscopically as well as electrochemically to make sure that there is no irreversible change in the film during the measurements. The difference absorption spectrum, A, is the absorbance change under external bias. At first, the absorbance was measured under a significantly positive potential to discharge all the stored electrons from the film. Then a negative potential was applied to increase the electron population and the absorbance was measured. When the Fermi level moves closer to the conduction band (CB) the density of the electrons (nCB) increases exponentially according to the following relation: nCB  N CB exp (EF  ECB ) k bT (1) where ECB is the conduction band minima and NCB is the density of states (DOS) in the conduction band. Absorbance measured at positive potential was subtracted from the absorbance at negative bias and thus A was calculated. The wavelength range was selected from 450 nm to 800 nm considering the appearance of a broad peak at ~400 nm and a strong solvent absorbance in the near IR region starting at ~ 900 nm. The peak ~400 nm was correlated to the surface trap states in literature.16 69 4.3 Results and discussion Figure 4-1 shows the change of transmittance (%T) when reductive potential was applied. %T shown here is the transmittance of the electrochemical cell which includes mesoporous TiO2 film, FTO glass substrate, cuvette and electrolyte in the optical path. In the absence of a TiO2 film, %T of the other components in the optical path did not change in response to the applied potential range of interest herein. Therefore, we attribute this change in %T exclusively to the TiO2 film. Figure 4-1 Transmittance (%T) spectra of TiO2 photoanode in contact with aqueous 0.2 M KCl solution at pH 12.8. Potentials are given against SCE. The absorbance difference (A), in 4-2a,b, shows the change in the absorbance of the TiO2 film when electron population is changed under applied potential. The absorbance increases in the visible and the near-IR regions with increasing (negative) applied potential with a concomitant decrease at wavelengths shorter than ~380 nm. The absorption bleaching below 70 ~380 nm is attributed to the filling of conduction band states which gives rise to the Burstein– Moss shift.17–22 (a) (b) Figure 4-2 Change in absorbance of a TiO2 electrode under reductive bias. (a) TiO2 film is in contact with a pH 12.8 aqueous electrolyte when potential is stepped from + 0.455 V to  1.245 V (red),  1.295 V (green),  1.345 V (yellow),  1.395 V (blue) and  1.445 V (pink) vs.SCE (from bottom to top). (b) TiO2 film is in contact with a pH 2.0 aqueous electrolyte, potential is stepped from + 0.955 V to  0.495 V (red),  0.545 V (green),  0.595 V (yellow),  0.645 V (blue) and  0.695 V (pink) ,  0.745 (sky blue) vs.SCE (from bottom to top). 71 (a) (b) Figure 4-3 Global fit results (black lines) of the absorbance spectra from Fig. 4-2a,b according to Drude equation. (a) TiO2 film is in contact with a pH 12.8 aqueous electrolyte when potential is stepped from + 0.455 V to  1.245 V (red),  1.295 V (green),  1.345 V (yellow),  1.395 V (blue) and  1.445 V (pink) vs.SCE (from bottom to top). (b) TiO2 film is in contact with a pH 2.0 aqueous electrolyte, potential is stepped from + 0.955 V to  0.495 V (red),  0.545 V (green),  0.595 V (yellow),  0.645 V (blue) and  0.695 V (pink) ,  0.745 (sky blue) vs.SCE (from bottom to top). 72 (a) (b) Figure 4-4 Global fit results (black lines) of the absorption spectra of TiO2 film according to Drude equation. Electrolyte is 0.2 M KCl at pH 6.2 adjusted by Citrate buffer (top) and Phosphate buffer (bottom). Potential shown in the plot is given vs. SCE. 73 (a) (b) Figure 4-5 Global fit results (black lines) of the absorption spectra of TiO2 film according to Drude equation. Electrolyte is 0.2 M KCl at pH 7.8 adjusted by Phosphate buffer (top) and Imidazole buffer (bottom). Potential shown in the plot is given vs. SCE. 74 (a) (b) Figure 4-6 Global fit results (black lines) of the absorption spectra of TiO2 film according to Drude equation. Electrolytes are 0.2 M LiClO4 at pH 7.8 (top) and ~ aqueous 0.05 M NaPF6 solution (bottom). The pH of the LiClO4 solution was adjusted at 7.8 by Imidazole buffer. No buffer was added into aq. NaPF6 solution. Potential shown in the plot is given vs. SCE. 75 (a) (b) Figure 4-7 Global fit results (black lines) of the absorption spectra of TiO2 film according to Drude equation. Electrolytes are 0.2 M LiClO4 (top) and TBAPF6 (bottom) in acetonitrile. Potential shown in the plot is given vs. SCE. The increasing visible and near-IR absorption with bias is due to the increasing electron population in the TiO2. We attribute this broad absorption to free (delocalized) electrons in the TiO2 conduction band, CB. The conduction band electrons (CBEs) are delocalized and they have 76 a continuum of possible electronic excited states resulting in a broad absorption in visible and near IR region. It showed a monotonic increase of absorptivity of TiO2 film with the wavelength According to the Drude-Zener theory, the absorption of delocalized (free) electrons gives rise to a wavelength dependent absorption spectrum according to, A  C P (2) where C is a scaling parameter,  is the wavelength of light and the exponent, P, is a value from 1.5 – 3.5, depending on the scattering mechanism. Scattering by acoustic phonons gives rise to P = 1.5, scattering by optical phonons produces P = 2.5 a scattering by ionic impurities gives P = 3 – 3.5.7 Figure 4-2a,b shows the difference absorption spectra in the wave length range of 450 nm to 800 nm and the results of a global fit of P to all spectra in equation 1. The fit produces a value of P = 1.66 and 1.85 when the TiO2 electrode is in contact with 2.0 and 12.8, respectively, which are well matched to the theoretical value of P for the absorption by free electrons coupled with acoustic phonon scattering. Analogous measurements were performed in a variety of pHs and non-aqueous electrolytes. The results of fitting the absorption spectra under these different conditions to equation 1 are shown in Table 4-1. In all cases, the P value is close to 1.5. This is consistent with the assignment of the absorption in this wavelength range being dominated by free conduction band electrons since this absorbance is a property of the nanoparticle and independent of contacting electrolyte. Panayotov et al. have done detailed study of the free electron absorption in TiO2.23–28 The free CB electrons were increased by UV- photoexcitation, n-doping (by atomic hydrogen) and thermal reduction, and the characteristic absorption spectra showed excellent agreement with the Drude equation (equation 2). Values of P were found to be 1.5 – 1.7 in their study which matches closely to the values we observed when we increased the 77 CB electrons by applying reductive potential. In a recent report, Trizio et al. observed similar free electron absorption in visible and IR region when the CB electron population in TiO2 nanoparticles was increased by niobium doping.29 It implies that the free CB electrons have major contribution to the visible absorbance of mesoporous TiO2 film under reduced bias. Table 4-1 Values of P in presence of different electrolyte solutions. The absorption spectra of TiO2 electrode in contact with different electrolyte solution were measured and fitted to Drude equation, A = CP. Global fit of several spectra measured in an electrolyte condition provided the value of P. Electrolyte solution P Aq. 0.2 M KCl, pH 2.0, pH adjusted by HCl 1.66 Aq. 0.2 M KCl, pH 6.2, Citrate Buffer 1.66 Aq. 0.2 M KCl, pH 6.2, Imidazole Buffer 1.57 Aq. 0.2 M KCl, pH 7.8, Imidazole Buffer 1.47 Aq. 0.2 M KCl, pH 7.8, Phosphate Buffer 1.63 Aq. 0.2 M LiClO4, pH 7.8, Imidazole Buffer 1.41 Aq. 0.2 M KCl, pH 12.8, pH adjusted by KOH 1.85 Aq. 0.2 M NaPF6 1.58 0.2 M TBAPF6 / Acetonitrile 1.56 0.2 M LiClO4 / Acetonitrile 1.99 Now, as discussed in Chapter 2, filling the lowest energy states of the CB blocks the band gap transition i.e. the minimum energy required to excite an electron from the top of the valence band to the bottom of the CB. The extra energy required to find the available energy states in the CB results in a blue shift in the band gap absorption which is generally known as Burstein-Moss shift. The magnitude of this shift translates into the density of electrons in the CB. These CB 78 electrons can absorp light in the visible – near IR region. Figure 4-8 shows the plot of absorbance measured at 780 nm vs. the density of electron calculated from the Burstein-Moss shift observed in UV region. The linear correlation of nCB and A further supports our assignment of absorbing species to the free CB electrons and also allows for the calculation of the extinction coefficient (CB) of these electrons from the slope according to Beer’s law: A   CBl (1  p) M CB (3) where l is the electrode thickness, p is the porosity and MCB is the molar concentration (mol L-1). Using values of l = 6.95 m and p = 0.71, CB was calculated to be 1.7  104 M-1cm-1 at 780 nm when TiO2 is in contact with pH 12.8. Our calculated value of CB is approximately an order of magnitude higher than the values found in literature measured using different methods.30–34 Prior methods did not discriminate free conduction band electrons from the large total concentration, ntotal, of trapped electrons. Thus, a large fraction of electrons in TiO2 are not participating in the absorption pattern in the visible range, which supports the higher value of CB determined here. When we use charge extraction method (described in Chapter 3) to quantify ntotal in TiO2, a valued of CB = 7  102 M-1 cm-1 is determined, consistent with literature underestimates. 79 Figure 4-8 Density of electrons in conduction band (nCB) is calculated from Burstein-Moss shift, described in Chapter 2, and absorbance (A780 nm) of TiO2 film is measured at 780 nm under different applied potentials mentioned in the plot. Electrolyte was 0.2 M KCl at pH 12.8 and the potentials are given vs. SCE. 4.4 Conclusion Detailed study of the absorbance pattern of the TiO2 was performed in contact with various electrolytes. Free electrons in the conduction band have the major contribution to the monotonic absorbance pattern of TiO2 semiconductor film in the visible – near IR region. The visible absorbance showed excellent linear co-relation with the density of the electrons in the CB. From this correlation the extinction co-efficient of the free CB (CB) electrons was calculated using Beer’s law. Determination of CB can provide a valuable insight into the electron transfer processes in various photoelectrochemical cells. 80 REFERENCES 81 REFERENCES (1) Brian, O.; Grstzel, M.; Fitzmaurice, D. Chem. Phys. Lett. 1991, 183, 89–93. (2) O’Regan, B.; Graetzel, M.; Fitzmaurice, D. J. Phys. Chem. 1991, 95, 10525–10528. (3) O’Regan, B.; Moser, J.; Anderson, M.; Graetzel, M. J. Phys. Chem. 1990, 94, 8720–8726. (4) Fabregat-santiago, F.; Mora-sero, I.; Bisquert, J. J. Phys. Chem. B 2003, 107, 758–768. (5) Bisquert, J. Phys. Chem. Chem. Phys. 2003, 5, 5360. (6) Kytin, V. G.; Bisquert, J.; Abayev, I.; Zaban, A. Phys. Rev. B 2004, 70, 193304. (7) Jacques I. Pankove. Optical Processes in Semiconductors; 1975th ed.; Dover Publications, New York, 1975. (8) Cao, F.; Oskam, G.; Searson, P. C.; Stipkala, J. M.; Heimer, T. a; Farzad, F.; Meyer, G. J. J. Phys. Chem. 1995, 99, 11974–11980. (9) Hodes, G. J. Electrochem. Soc. 1992, 139, 3136. (10) Albery, W. J. J. Electrochem. Soc. 1984, 131, 315. (11) Berger, T.; Anta, J. A.; Morales-flo, V. J. Phys. Chem. C 2012, 116, 11444–11455. (12) Berger, T.; Anta, J. a; Morales-Flórez, V. Phys. Chem. Chem. Phys. 2013, 15, 13790. (13) Berger, T.; Lana-villarreal, T.; Go, R. J. Phys. Chem. C 2007, 111, 9936–9942. (14) Cao, F.; Oskam, G.; Searson, P. C.; Stipkala, J. M.; Heimer, T. A.; Farzad, F.; Meyer, G. J. J. Phys. Chem. C 1995, 99, 11974–11980. (15) Meekins, B. H.; Kamat, P. V. ACS Nano 2009, 3, 3437–3446. (16) Boschloo, G.; Fitzmaurice, D. J. Phys. Chem. B 1999, 103, 2228–2231. (17) Gibbs, Z. M.; LaLonde, A.; Snyder, G. J. New J. Phys. 2013, 15, 75020. (18) Sarkar, a.; Ghosh, S.; Chaudhuri, S.; Pal, A. K. Thin Solid Films 1991, 204, 255–264. (19) Kim, S. S.; Park, J.; Kim, S. S.; Lee, Y.; Ahn, S.; Cho, J.; Ju, M.; Lakshminarayan, N.; Ai, V.; Yi, J. Vaccum 2014, 108, 39–44. (20) Hamberg, I.; Granqvist, C. G.; Berggren, K. F.; Sernelius, B. E.; Engström, L. Phys. Rev. B 1984, 30, 3240–3249. (21) Burstein, E. Phys. Rev. 1954, 93, 632–633. (22) Kamat, P. V.; Dimitrijevic, N. M.; Nozik, a. J. J. Phys. Chem. 1989, 93, 2873–2875. 82 (23) Panayotov, D. a.; Yates, J. T. Chem. Phys. Lett. 2007, 436, 204–208. (24) Panayotov, D. a.; Burrows, S. P.; Morris, J. R. J. Phys. Chem. C 2012, 116, 4535–4544. (25) Nanoparticles, T.; Panayotov, D. a; Burrows, S. P.; Morris, J. R. J. Phys. Chem. 2012. (26) Panayotov, D. a.; Desario, P. a.; Pietron, J. J.; Brintlinger, T. H.; Szymczak, L. C.; Rolison, D. R.; Morris, J. R. J. Phys. Chem. C 2013, 117, 15035–15049. (27) Panayotov, D. a.; Yates, J. T. J. Phys. Chem. C 2007, 111, 2959–2964. (28) Panayotov, D. a.; Burrows, S. P.; Yates, J. T.; Morris, J. R. J. Phys. Chem. C 2011, 115, 22400–22408. (29) De Trizio, L.; Buonsanti, R.; Schimpf, A. M.; Llordes, A.; Gamelin, D. R.; Simonutti, R.; Milliron, D. J. Chem. Mater. 2013, 25, 3383–3390. (30) Rothenberger, G.; Fitzmaurice, D.; Graetzel, M. J. Phys. Chem. 1992, 96, 5983–5986. (31) Safrany, A.; Gao, R.; Rabani, J. J. Phys. Chem. B 2000, 104, 5848–5853. (32) Boschloo, G.; Fitzmaurice, D. J. Phys. Chem. B 1999, 103, 3093–3098. (33) Liu, Y.; Jennings, J. R.; Huang, Y.; Chem. C 2011, 115, 18847–18855. ang, .; Zakeeruddin, S. M.; Grätzel, M. J. Phys. (34) Kuciauskas, D.; Monat, J. E.; Villahermosa, R.; Gray, H. B.; Lewis, N. S.; McCusker, J. K. J. Phys. Chem. B 2002, 106, 9347–9358. 83 Chapter 5 Conduction Band Energy and Charge Distribution in Strontium Titanate Nanoparticle Electrode 5.1 Introduction Nanostructured wide bandgap semiconductor photoanodes are being widely used in solar energy conversion systems like Dye-sensitized Solar Cell (DSSC) as their high surface area (2-3 orders of magnitude higher than geometric area) can be utilized to optimize the light harvesting efficiency.1–7 Additionally, they offer the possibilities of overcoming the challenges associated with short diffusion length of various materials in bulk structure. In the context of DSSC and related solar energy conversion systems, TiO2 is extensively used as a photoanode material. Relying on a single photoanode material has restricted the use or optimization of many sensitizers because of the mismatch in energetics. In DSSCs, the photoanode directly participates in the charge injection processes from the photoexcited dye, transporting these electrons and also in the recombination processes. In order to properly understand these processes, photoanode materials having different energetics and charge distributions need to be studied. Though other metal oxides like ZnO, SnO2, Nb2O5, SrTiO3 have been used, their performances have been much lower than the DSSCs constructed with TiO2 photoanodes.8 Those reports are primarily based on the optimization of photoconversion efficiencies. Proper correlation between the material properties (e.g. energetics and charge distribution) and electron transfer processes (e.g. recombination processes) are inadequate in literature. Thus further study needs to be performed to assess their potential and to understand the photoanode related charge transfer processes. Strontium titanate (SrTiO3) has the potential to be a good photoanode material in DSSCs.9–11 It has Perovskite structure and has structural similarities with anatase TiO2 as shown in Figure 5- 84 1.12 Perovskite structure is described by the chemical formula ABO3 where the B atoms (Ti here) have a 6-fold coordination and O atoms form octahedra surrounding the B atoms. Each A atom (Sr here) is surrounded by four TiO6 octahedra. SrTiO3 Perovskite structure contains titanium atoms in 6-fold octahedral coordination, similar to the titanium atoms in anatase TiO2. They both are wide band gap semiconductor and do not absorb light in the Visible spectra which is essential requirements for a semiconductor to be used in Dye-sensitized photoconversion technologies.13,14 Additionally, similar to TiO2, the conduction band in SrTiO3 is mainly made of the Ti 3d states with a small component from the O 2p states.15,16 It has been reported that d-orbitals provide more effective overlap with the sensitizer and facilitate more efficient electron transfer from the Ruthenium based sensitizers. Unlike TiO2, the CB of SnO2 and ZnO are sp-type.17,18 Additionally, CB band composed of empty d-orbitals has high density of states (DOS). The injection efficiency in sensitized photoanodes has been found in the following order: TiO2>SnO2~ZnO. The DOS in the CB of TiO2 and SrTiO3 are similar (~ 1020 cm-3).19 Figure 5-1 Crystal structures of Perovskite SrTiO3 and anatase TiO2. This figure is reprinted with permission from reference 12 (Bera, A.; Wu, K.; Sheikh, A.; Alarousu, E.; Mohammed, O. F.; Wu, T. J. Phys. Chem. C 2014, 118, 28494–28501). Copyright 2014 American Chemical Society. 85 Sensitized SrTiO3 anode has been reported to provide higher photovoltage than TiO2 which implies higher conduction band (CB) energy or/and lower photoanode related recombination processes in SrTiO3 than TiO2.10,20,21 Under illumination, the photoexcited dye injects electron into the semiconductor. It increases the electron population in the semiconductor and the Fermi level of electrons (EF) moves closer to the CB according to the following equation: nCB  N CB exp (EF  ECB ) k bT (1) where nCB is the density of electrons in the CB, ECB is the CB energy minimum, kb is the Boltzmann constant, T is the absolute temperature and NCB is the density of states (DOS) in the CB which is calculated by,  m*k T  NCB  2  e b2   2  3/2 (2) where me* is the effective mass of CB electrons. The charge separation in DSSC takes place at the dye-semiconductor interface.13 Separation between the solution potential and the Fermi level of the electrons in the semiconductor is the voltage output from a cell. At open circuit, this separation is the highest and the corresponding voltage is called open circuit voltage, VOC. Thus, if the CB of a material is higher than TiO2, higher photovoltage can be generated by pushing the Fermi level potential, EF, at higher potential. SrTiO3 could potentially offer higher VOC if the recombination losses are minimized. There are reports on DSSCs constructed with SrTiO3 photoanode.9–11,22 The quality of those photoanodes, however, are not ideal for systematic physical studies of the material properties. In 1999, Burnside et al. showed the potential of SrTiO3 as a DSSC photoanode and observed about 100 mV higher VOC comparing to TiO2 based DSSC.10 However, the size and shape of the SrTiO3 86 particles have large distribution and even have bimodal particle distribution. Additionally, trace amounts of SrCO3 was present in their sample.10 The following reports also suffered from the presence of SrCO3.11 In 2014, Jayabal et al. reported a method to eliminate the impurities by heating the SrTiO3 sample at a very high (900 C) temperature. However, well defined nanoparticles were not observed possibly because the particle morphology was changed during the heating at such a high temperature.9 Nanostructured SrTiO3 has also been used in Hybrid Perovskite (CH3NH3PbI3−xClx) Solar Cells and the open circuit voltage, VOC, was found to be ~ 120 mV higher that solar cell constructed with TiO2.12 One of the major issues, in the context of exploring new nanostructured photoanode materials, is the lack of reliable methods to determine important material properties of the photoanodes such as the conduction band edge position, distribution of electrons with the energetics and, also, the spatial locations of the trap states. Recently, we have developed a spectroelectrochemical method to determine the conduction band edge and charge distribution in nanostructured photoanodes.19,23 In this work, we report a reliable method to make mesoporous transparent SrTiO3 photoanode and, in order to characterize its potential for photoanode applications, a detailed spectroelectrochemical and electrochemical studies of its energetics and the charge distribution have been performed. The spectroelectrochemical method is described in Chapter 2 and 3 where the CB edge of TiO2 has been determined in presence of aqueous electrolyte (0.2 M KCl, pH 12.8). The CB edge of SrTiO3 electrode was determined in contact with dry acetonitrile solution (0.1 M LiClO4). Aqueous solution could not be used with SrTiO3 because the light absorption measured in the UV range (band gap transition) under negative bias goes through irreversible change. SrTiO3 has been reported to be capable of photoassisted electrolysis of water 87 and the generated gaseous bubbles, trapped in the mesoporous structure, might be the reason behind the irreversibility of measured light absoption.20 5.2 Experimental 5.2.1 Preparation of SrTiO3 electrode Synthesis of SrTiO3 nanoparticles: SrTiO3 nanoparticles were made through a vapor diffusion sol−gel (VDSG) method developed by Brutchey el al.24,25 Strontium titanium alkoxide (SrTi[OCH2CH(CH3)OCH3]6; 0.7 M in n-butanol/3-methoxypropanol) was used as precursor and were used as received from Gelest, Inc. It was stored in a N2 glove box. ACS grade hydrochloric acid was used. B A C Figure 5-2 Apparatus setup used for the synthesis of SrTiO3 nanocrystals via vapor diffusion sol-gel method. All procedures were done under N2 atmosphere using a standard Schlenk-line system. A 100 mL three-neck flask was connected to vacuum adapter-glass stopcocks. Two of the adapters were 88 used as N2 purging line and vent, respectively and the third one was used to flow acidic vapor into the reaction flask. A glass bubbler was filled with 150 mL of 0.75 M HCl solution. A flowmeter was used to control the N2 flow though the bubbler. The flask was initially purged with N2 for 30 minutes. In this configuration, adapters B and C (one for N2 in and one for vent) are connected to the 3-neck flask and one of the necks, A is closed by a septum. N2 gas was then bubbled into the HCl/H2O to remove the dissolved air (CO2 specifically) and to saturate the N2 gas flow with H2O vapor. After 30 min N2 purging, the strontium titanium alkoxide precursor has been injected into the 3-neck flask. Then the bubbler was connected to the 3-neck flask (A) and the adapter which connects the flask with the N2 flow (B) was closed. Thus, in this setup, two adaptors were open, one for flowing HCl/H2O/N2 into the flask (A) and other for venting (C). The flow rate was 0.02 L min-1. A venting adapter was also connected to a bubbler to prevent any air exposure into the reaction flask. After ~7 hours a gel appeared which eventually cracked. After 5 days, the flask was opened and the gel was collected and washed with ethanol. During the washing process, the gel was sonicated with 5 mL ethanol for 15 min and centrifuged at 6000 rpm for 1 hour. This washing was repeated 3 times. The resulting powder was dried under vacuum for overnight at 60 C and was grounded in a mortar pestle to make SrTiO3 powder. Preparation of SrTiO3 paste: SrTiO3 powder was mixed with ethanol and dispersed using an ultrasonic horn. The dispersion was then centrifuged at 2000 rpm to remove the larger particles and aggregates. The decant was then mixed with ethyl cellulose and -terpineol and was ultrasonicated again. The excess solvent was then evaporated using a rotavap at 60 C until a viscous paste was formed. This paste was kept overnight before further use. 89 Preparation of SrTiO3 electrode: The transparent mesoporous SrTiO3 film was prepared by spreading SrTiO3 paste on transparent conductive substrate. This substrate was made by depositing ~10 nm thick layer of TiO2 on fluorine-doped tin oxide glass (FTO) (Hartford Glass) by atomic layer deposition (ALD). This film was first annealed at 360°C for 30 min in the air and then it was annealed at 560 C under hydrogen/nitrogen mixture (5% H2). 5.2.2 Characterization of SrTiO3 film Powder X-ray diffraction (XRD) of SrTiO3 was measured by Bruker Davinci Diffractometer (Cu K radiation ( = 1.5406 Å). The 2 range was 20-70, step size was 0.02 and the collection time was 3 s per step. The size, shape of SrTiO3 nanoparticles and the morphology of the SrTiO3 film were characterized by using a scanning electron microscope (Hitachi S-4700 II FESEM). 5.2.3 Preparation of electrolyte solutions Lithium perchlorate (Sigma-Aldrich, battery grade, dry, 99.99% trace metals basis) was dissolved in anhydrous acetonitrile to make a 0.1M solution. 5.2.4 Spectroelectrochemical measurement In spectroelectrochemical measurements, the transmittance (%T) of the film was measured by a Perkin Elmer Lambda 35 UV-Vis spectrophotometer under an external bias provided by a potentiostat (-Autolab, Metrohm). A three electrode electrochemical cell setup was made by making a rectangular hole in a cuvette which can be sealed by a cap. A platinum mesh was used as counter electrode and a homemade Ag/AgNO3 was used as reference electrode. This cuvette was placed into the UV-Vis spectrophotometer such that the SrTiO3 electrode remained perpendicular to the optical path. The SrTiO3 electrodes were aged overnight in the glove box in 90 contact with the electrolyte solutions prior to any experiment. The temperature of the electrochemical cells was 220.5 ⁰C during the measurements. During spectroelectrochemical measurements, each film was stabilized at an applied bias for 3 minutes and %T of the electrode was measured. Stability of the film as well as reproducibility of the results were checked both spectroscopically as well as electrochemically to make sure that there is no irreversible change in the film during the measurements. The cuvette was sealed in the glove box and all the measurements were done under sealed condition i.e. in dry N2 atmosphere. The absorbance of SrTiO3 electrode was determined by subtracting the absorbance of the cell having a bare conductive substrate without a SrTiO3 film from the absorbance measured with SrTiO3 film to eliminate the contribution of everything else in the optical path (FTO, cuvette and electrolyte) and provides only the changes in the SrTiO3 film. Absorbance, A, was calculated from the measured %T by using the relation, A = 2  log10%T. The difference absorbance spectrum, A, is the absorbance change under external bias. At first, the absorbance was measured under a sufficiently positive potential (0.0 V vs. Ag/AgNO3) to discharge all the stored electrons from the film. Then a negative potential was applied to increase the electron population and the absorbance was measured. The absorbance measured at positive potential was subtracted from the absorbance at negative bias to calculate A. During the study of the band gap transitions in the UV region (330 – 400 nm) under potential bias, the absorption in the visible-near IR (400 - 1100 nm) was fitted to a Gaussian fit using OriginPro 9.0 and was extrapolated in the UV region. It was then subtracted from the absorbance measured in the UV region to determine the absorbance of the band gap transition only. Absorption co-efficient () was calculated from absorbance by using the relation  = (2.303  A) / l, where A is the absorbance and l is the thickness of the film.26,27 The thickness of the films 91 varied from 1.9 to 2.1 m. The thicknesses of the films were measured after each spectroelectrochemical measurement for accurate calculation of the absorption properties of the SrTiO3 electrode. 5.2.5 Electrochemical measurements Cyclic voltammetry (CV), Electrochemical Impedance Spectroscopy (EIS) and Charge Extraction methods were performed to study the electron distribution in SrTiO3 electrodes. A three electrode set up was used. A Pt mesh was used as counter electrode and a homemade Ag/AgNO3 (in anhydrous Acetonitrile) reference electrode was used. All the measurements were done under dry N2 atmosphere. 5.3 Results and Discussion 5.3.1 Material characterization SrTiO3 nanoparticles were synthesized through a vapor-diffusion sol–gel route at room temperatures, in the absence of any structure-directing template.28–30 Nanoparticles were formed through a kinetically controlled coalescence processes by slowly reacting acidic water vapor with the metal alkoxides. The chemical reaction can be expressed as, HClcat . SrTi[OCH 2CH3OCH3 ]6  3H 2O   SrTiO3  6HOCH 2CH (CH3 )OCH3 These nanoparticles were used to make films which were annealed at 560 C under a reducing atmosphere (5% H2 / N2). Powder XRD analysis showed that the nanoparticle film is comprised of pure SrTiO3 with a cubic Perovskite phase.25 No other peaks associated to other phases or impurities were observed. Scanning electron microscope analysis revealed the nanocrystal diameter of about 152 nm. 92 Figure 5-3 Powder X-ray diffraction patterns of the SrTiO3 nanocrystals (top) and scanning electron micrographs SrTiO3 film deposited on conductive substrate (bottom). 93 5.3.2 Determination of the conduction band and electron distribution Spectroelectrochemical analysis of SrTiO3 electrode: The Fermi level of electron within this semiconductor nanoparticle film can be controlled by the potential applied to the conductive substrate.31–33 Figure 5-4 and 5-5 show how the absorbance of the SrTiO3 electrode changed when the electron population was increased by applying a negative (i.e. reductive) potential. In the visible-near IR region an absorbance peak appeared at ~894 nm under negative bias which increased gradually with the increasing bias. A visible-near IR absorbance was also observed for TiO2 electrodes under negative bias, however no peak was observed and the absorbance spectra was dominated by free electron absorbance.23 The difference absorbance spectra of SrTiO3 (Figure 5-6), shows the change in absorbance when the electron population was changed. Under, negative potential, the densities of both trapped electrons and delocalized (also called “free”) CB electrons were increased. According to the Beer’s law, the total absorbance, A, can be defined as, A / l   CB M CB   T M T where l is the electrode thickness, MCB and MT are the molar concentrations (mole L-1) of CB electrons and the trapped electrons, respectively. CB and T are the extinction coefficients of the free CB electrons and trapped electrons, respectively. According to the Drude-Zener theory, the absorption of delocalized (free) electrons gives rise to a wavelength dependent absorption spectrum according to A  C P , where C is a scaling parameter,  is the wavelength of light and the exponent, P, is a value from 1.5 – 3.5, depending on the scattering mechanism. Scattering by acoustic phonons gives rise to P = 1.5, scattering by optical phonons produces P = 2.5 and a scattering by ionic impurities gives P = 3 – 3.5. In addition to the absorbance peak at ~894 nm (Figure 5-6) a featureless absorbance was also observed in the wavelength range of 1500-2200 nm. Both the peak and the featureless absorbance increased with increasing electron population. A possible 94 assignment of the absorbance peak is the inter-band localized electronic states though it is not possible to explain the nature and distribution of these localized states, based on this observation.34 The featureless absorbance can be explained by the intra-band absorption of the free electrons in the CB. Figure 5-4 Plot of absorbance of SrTiO3 electrode in UV region under various applied potentials, 0.0 V (black), -1.3 V (blue), -1.4 V (red), -1.5 V (green) vs. Ag/AgNO3. The electrolyte solution is 0.1 M LiClO4 in acetonitrile. Increasing negative potential moves the band gap absorption edge to higher energy. 95 Figure 5-5 Plot of visible-near IR absorbance spectra of SrTiO3 electrode under various applied potentials, 0.0 V (black), -1.3 V (blue), -1.4 V (red), -1.5 V (green) vs. Ag/AgNO3. Figure 5-6 Plot of absorbance difference spectra of SrTiO3 electrode under various applied potentials, -1.3 V (blue), -1.4 V (red), -1.5 V (green). This difference absorbance was calculated by subtracting the absorbance of SrTiO3 electrode at 0.0V vs. Ag/AgNO3 from the absorbance measured at negative potentials shown in the plot. 96 In the UV region, the band gap transition of SrTiO3 occurs below 380 nm. Interestingly, a gradual blue shift of the absorption onset was observed with increasing negative potential i.e. with the increasing electron population in the SrTiO3 film. Band gap energy is the lowest energy required to excite an electron from the valance band top to the CB bottom. This blue shift of the band gap transition can be explained by the filling of the bottom of the CB which blocks the lowest energy band gap transitions. If the lowest energy states in CB are filled, band gap transition needs higher energy to find the lowest empty states. Similar blue shift in the band gap absorption onset has been observed in the highly doped semiconductors where the introduced dopant increases the population of the CB electrons. This phenomenon is generally described as the Burstein-Moss shift.35–40 Determination of the conduction band edge: This apparent band gap widening (EG) is correlated to the density of electrons in the CB (nCB) according to the following equation: EG  h 2  3nCB  8m*    2/3 (3) where m* is the reduced effective mass of the charge carriers and h is the Plank’s constant. Knowledge of nCB leads to the position of ECB according to the equation 1. The band gap energy (EG) of a semiconductor material can be determined from a Tauc plot using the relation,27 ( hv)  A' (EG  hv) (4) where  is the absorption coefficient, h is the energy of the incident light, A’ is a constant and the value of  describes the characteristics of the transition process. The band gap transition in SrTiO3 material is known to be indirect in nature i.e. electron transition from the highest-energy state in the valence band to the lowest-energy state in the conduction band is not possible without 97 a change in momentum.16,41,42 The transition occurs via the absorption or emission of a phonon which provides the required momentum change during the transition. The energy required for this transition mainly comes from the photon absorbed.  in equation 4 is 1/2 in case of indirect semiconductor. Figure 5-7 shows the Tauc plot of SrTiO3 film. The range of photon energies utilized to fit the Tauc plot was chosen where a good straight line fit was observed beyond the band gap transition energy. The energy region should not be too close to the band gap as various factors (e.g. Urbach tail) can dominate the overall absorption spectra where the absorption of band gap transition is still relatively low. On the other hand, in a scenario when the photon energy is very much higher the band gap energy, the parabolic approximation of E-k plot is no longer valid. Also, the direct band gap transition (which occurs at higher energy than the indirect transition in anatase material) can contribute to the absorption. The maximum energy applied here is 3.55 eV ( = 350 nm) and the direct interband direct transition (i.e. trasition from VB to CB without change in momentum) has been reported to occur above this energy range. 43,44 We observed a sudden rise of the absorption spectra at energies higher than ~ 3.6 eV presumably due to the contribution of direct inter band transition. Direct allowed transition shows sharp rise of the absorption co-efficient with beyond the energy required to excite an electron from VB to CB without changing the momentum.27 The (h)1/2 vs. h curve fitted between 3.543 eV to 3.647 eV (350 - 340 nm) was extrapolated to determine the band gap and it was found to be 3.066 eV. Though 3.2 eV is frequently reported as the band gap energy of SrTiO3 material, Li et al. reported a band gap energy of 3.01 eV for SrTiO3 micro-structures.11,16,20,41,43–46 The value of EG we observed is in the lower end of the literature values for the band gap of SrTiO3. 98 Figure 5-7 Tauc plot of a SrTiO3 film in contact with 0.1 M LiClO4 in acetonitrile. Any contribution from the electrolyte, substrate and cuvette were subtracted and only the absorbance of the SrTiO3 electrode contributed to this plot. Figure 5-8 Plot of band gap widening while moving from 0.0 V (black) towards more negative potentials -1.3 V (blue), -1.4 V (red), -1.5 V (green) vs. Ag/AgNO3. 99 There are some possible sources of uncertainty in determining the band gap energy in our experiment. In the indirect transition, both the energy and the momentum are changed. As a result, phonon interaction, a quantum mechanical demonstration of thermal vibrations in the crystal lattice, is required to conserve the momentum. Therefore the band gap absorbance becomes a two-step process assisted by either phonon absorption or phonon emission. So the energy required to excite an electron from the valence band to the conduction band can be written as,27 h  EG  Ep (5) where EP is the energy of the phonon required to compensate the difference in momentum between the minima and maxima of the conduction and valence bands, respectively. Now, both phonon absorption and emission are possible as the range of applied photon energy considered to calculate the band edge is much higher than EG. The phonon energy is expected to be constant while changing the potential at constant temperature. It should also be mentioned that it is not straightforward to determine the absolute band gap in case of phonon assisted indirect transition. Additionally, in presence of a high dopant density and/or impurity, charge-charge and/or chargeionic impurity scattering can also provide the momentum conservation required for the indirect transition.27 The dopant density as well as impurity concentration in nanostructured SrTiO3 films has not been accurately determined so far. Broadening of the absorption curves also contributes to the uncertainty in determining the band gap energy accurately. Band tail, which is expected to be present in our sample due to the disorder effects and/or defects below the conduction band, can also contribute to this uncertainty. Significant broadening of the absorption is observed in our measurements (Figure 5-7). It is very difficult to get rid of this broadening which might cause some error in our band gap determination though the magnitude of this broadening should 100 remain constant for a specific measurement setup, which we observe. Thus, at constant temperature and electrolyte environment, the error in determining the band gap transition remains constant and does not contribute to the magnitude of the shift in the band gap absorbance which is of primary interest here. Figure 5-8 shows a continuous apparent increase of the measured band gap. The magnitudes of Burstein-Moss shift were calculated at a range of negative potentials and nCB was calculated from equation 3. me* = 5me and mh* = 1.2me have been used in this study.41,47–52 The position of the conduction band edge ECB/q was then calculated according to equation 1 as shown in table 1. Table 5-1 Calculation of conduction band position from the Burstein-Moss shift. E vs. AgNO3/V EG/meV nCB/cm-3 ECB/q/V -1.5 72.07  1.91 6.66  0.265  1019 -1.537  0.001 -1.45 60.08  1.93 5.07  0.244  1019 -1.494  0.001 -1.4 49.48  1.95 3.79  0.224  1019 -1.451  0.002 -1.3 33.44  1.98 2.10  0.187  1019 -1.366  0.002 -1.2 20.05  2.0 9.78  1.46  1018 -1.286  0.004 -1.1 11.32  2.02 4.15  1.11  1018 -1.207  0.007 A very interesting feature is observed in this table. The position of the conduction band is shifting upwards (i.e. more negative) while raising the Fermi level, EF/q (which is equal to the applied potential, E). This result is consistent with our prior report (described in Chapter 2) where similar upward shift of CB was observed in nanostructured TiO2 photoanodes.23 This upward shift of band energies occurs due to the potential drop across the double layer (nanoparticle-electrolyte interface). The electron accumulation is accompanied by positive 101 charges, i.e. the cations in the electrolyte. Accumulation of electrons in the electrically active states on the surface or within close vicinity of the surface of SrTiO3 nanoparticle changes ionic distribution and hence the electric field in this interfacial double layer. Though the physical nature of these electrically active states is uncertain, they are most likely to be localized states arising from the under coordinated TiIV or TiIII centers, hydroxyl group or any other crystal ‘defects’ on the surface or within very close vicinity of the surface that can trap electrons. 53 We refer to these states as “surface states” in the following sections. Determination of the total electron density, nTotal: nTotal was calculated over a range of Fermi levels by a charge insertion-extraction method. The applied bias was switched from 0.0 V vs. AgNO3 to a negative bias, E, to initiate a charge insertion process and the reverse process (E  0.0 V) was performed to extract the charge. The charge accumulated under a negative bias was calculated by integrating the current-time curves as shown in Figure 5-9. A reasonable match between the inserted charge and the extracted charge was observed which suggests that the charge insertion is reversible and no significant redox reaction occurred under negative bias. 102 Figure 5-9 Current vs. time plot measured during charge insertion-extraction process under various potentiostatic biases. A reductive bias (as mentioned in the plot) was applied for 3 min and then an oxidative bias (0.0 V vs. AgNO3) was applied to extract all the charges stored in SrTiO3 electrode. Figure 5-10 Plot of electron density in the nanostructured SrTiO3 electrode as a function of applied potential measured by charge insertion-extraction measurement. 103 Determination of the density of trap states on the surface (nS) and in the bulk (nB): The upward shift of the CB provided a method to calculate the density of surface states, nS, since it should be correlated to ΔECB by following relationship: ECB / q  qnS CH (6) Here CH is the capacitance of the Helmholtz layer formed on the surface of the nanoparticles. This capacitance depends on the width of the Helmholtz layer and its corresponding dielectric constant. To the best of our knowledge, CH has not been measured for SrTiO3 electrode in contact with dry acetonitrile solvent. CH was reported to be 52 F cm-2 for TiO2 electrode in presence of dry acetonitrile.54,55 Considering the structural similarity between SrTiO3 and TiO2, CH of SrTiO3 electrode in dry acetonitrile solvent was assumed to be 52 F cm-2 in our calculation.53 Additionally, the density of the bulk trap states, nB, can also be calculated from the following equation, nTotal  nCB  nS  nB (7) In order to calculate nS, the magnitude of the CB shift, ΔECB, needed to be determined which required the information of the minimum CB position, E*CB, without the effect of the potential drop. However, it is not straightforward to determine where the surface state distribution starts to cause significant potential drop in the Helmholtz layer. Thus, we need to rely on a rational estimation based on the electrochemical data. The charge insertion and extraction data (Figure 510) suggested that the electron density in the SrTiO3 starts increasing when the Fermi level is pushed beyond -1.0250.025 V vs. AgNO3. The measured capacitance by electrochemical impedance measurement also supported it as shown in Figure 5-12. Thus the position of CB at 104 this Fermi level provides a good estimation of E*CB and was calculated to be  1.123 V vs. AgNO3. ECB (= E*CB  ECB) was calculated over a range of Fermi levels, EF/q (= E) to determine the energy distribution of surface trap states. nB was calculated by subtracting nCB + nS from nTotal. Figure 5-13 shows the distribution of nCB, nS and nB as a function of applied potential (= EF/q). Interestingly, nS does not increase with applied bias as much as nB does. The Li+ cations are the primary charge compensating ion here. The electrons trapped on the surface can easily be charge compensated by Li+ ions in the electrolyte solution. However, to compensate bulk electrons, Li+ ions are likely to be inserted into the nanoparticle which involves reorientation of the bond lengths in the nanocrystals. Thus, initially, when the negative bias is not high enough, more electrons are trapped on the surface because they can be easily charge–compensated by the available Li+ cations. However, increasing Coulombic repulsions among the surface adsorbed Li+ cations limits the capacity to trap a large density of electrons on the surface. In the bulk of the nanoparticle, more electrons can be stored if charge-compensating Li+ can reside in the interstitial space in an O6 octahedral site similar to the anatase crystal. Thus, when the potential was negative enough to overcome the energy requirement for Li+ insertion, a large amount of electrons were trapped in the bulk.19 Determination of the density of states from electrochemical methods: The DOS can be calculated from the measured charge density while changing the Fermi level by the following relations, N (E)  105 dn dE (8) n EF  N ( E )dE (9) E0F The electron distribution is also described by its chemical capacitance (C) which describes the capacity of the nanostructured film to store/release electrons during a change in the chemical potential, n.32 When Fermi level, EFn (=E) changes, it causes a change in the electron density which changes the chemical potential of the electrons. This can be written as dEFn = dn and the chemical capacitance can then be written as the variation of the electron density as a function of Fermi level,32 Cμ  q dn dEFn (10) Note that, equation 8 and 10 are essentially describing the same thing. Figure 5-13 shows the C calculated from the charge insertion-extraction data and from EIS measurements.56,57 A transmission line model was used to fit the EIS data in case of porous photoanodes, as shown in Figure 3-3. This circuit illustrates the internal distribution of chemical capacitance in response to the modulated small perturbation of the steady state bias.58,59 106 Figure 5-11 Plot of electron density in SrTiO3 and TiO2 electrode under various applied biases calculated by charge extraction method. Figure 5-12 Plot of measured capacitance, C (red) and resistance, R (blue) of SrTiO3 electrode measured by electrochemical impedance spectroscopy. 107 Figure 5-13 Comparison of capacitance of SrTiO3 electrode measured by the electrochemical impedance spectroscopy and charge insertion-extraction method. Figure 5-14 Plot of electron density in the nanostructured SrTiO3 electrode as a function of applied potential. nCB is density of free electrons in the conduction band, nS, and nBulk are density of electrons trapped on the surface and in the bulk respectively. 108 Analysis of current decay (charge insertion and extraction): Wang et al. investigated the trap states in nanoparticle TiO2 electrode based on current decay measurements.60 Lindstrom et al. studied the current decay characteristics of nanoparticle TiO2 films in the presence of Li+ cations.61 They found that the shape of the current decay is controlled by the filling of trap states. The current decay was studied on nanoparticle ZnO films as well.62 Current decay characteristics of TiO2 film was also reported under photoexcitation.63,64 Here, a detailed analysis of the current decay characteristics of nanoparticle SrTiO3 film is presented. When a reductive bias is applied to the photoanode, the conductive substrate (FTO here) injects electrons into the nanoparticles directly attached to FTO. These electrons then diffuse through the nanoparticle network. At the same time, the cations from the electrolyte (Li + here) are either adsorbed on the surface of the nanoparticle or diffuse into the nanoparticles to maintain charge neutrality. Weppner et al. studied the movement of these electrons and suggested a cation coupled electron transport.65 Figure 5-9 shows that as soon as the potential was shifted from 0.0V vs. AgNO3 to more negative potentials, a sharp cathodic current peak was observed which then decayed with time. The characteristics of the decay remained almost same from -0.8 to -1.05 V vs. AgNO3 and the current decays were notably faster than the decays observed when the applied bias is more negative than -1.05 V (Figure 5-15). Moreover, the charge, Q, calculated from the decay plot (i-t) remained almost similar from -0.8 to -1.05 V. Beyond, ~ -1.05 V, the decay became noticeably slower and Q increased exponentially towards more negative potential. Now, the SrTiO3 is a wide band gap semiconductor and generally behave as an insulator.20,66 Moderate electron conductivity can be achieved under reduced condition. When the negative (reductive) potential was applied, the electron density in the material and hence the electron conductivity was 109 increased. At the potentials more positive than -1.05 V vs. AgNO3, the electron transport through the SrTiO3 nanoparticles were very low and the observed current decays were possibly due to the reorganization of electrolyte at the conductive substrate (FTO). Figure 5-11 shows that the electron distribution in the nanostructured SrTiO3 effectively starts from -1.05 V vs. AgNO3 which was also supported by the electrochemical impedance data (Figure 5-12). The current decay characteristics, observed when the applied bias is more negative than -1.05 V, were analyzed to extract information about the density of states (DOS) of the trap states. In nanoparticle TiO2 and ZnO, the characteristics of the current decay were observed to be controlled by the filling of the trap states and similar explanation can be provided in case of SrTiO3.61–64 When a negative (reductive) potential was applied the electrons were injected from the conductive substrate into the SrTiO3 nanoparticles. The injected electrons were trapped by the localized trap states followed by detrapping (i.e. releasing the electrons). Interestingly, we observed that the current decay is significantly influenced by the applied potentials which implies that some type of driving force was associated to the filling of the empty states. Also, more than one characteristic were observed in the current decay plot (Figure 5-15) which suggested that multiples processes were going on when a reductive potential was applied. We tried to deconvolute the decay plot in order to extract information about distributions of the density of states in this mesoporous SrTiO3 photoanode. Our previous measurements suggested that the mesoporous SrTiO3 film has mainly three types of electronic states, CB states, bulk trap states and surface trap states. Based on this idea and the observation of three decay characteristics, we fitted the current transient (i vs. t) plot as a tri-exponential decay. After deconvoluting the parameters associated to decay from each of the electronic states, the total charge inserted into those states were calculated. We would also like to mention the uncertainty 110 (a) (b) Figure 5-15 (a,b) Current transient data of when applied bias was changed from 0.0 V to the negative potentials mentioned in the plot. The SrTiO3 electrode is in contact with 0.1 M LiClO4/Acetonitrile solution. in extracting the parameters from a triexponential fit. Relying only on this fitting to determine the DOS of different states could lead to erroneous results. It requires a careful analysis of the 111 time constants and the DOS calculated at each electronic state throughout a range of potentials. The results were also compared with the spectroelectrochemical results. Figure 5-16 Triexponential fit of the current transient data when applied bias was changed from 0.0 V to -1.5 V vs. AgNO3. The experimentally observed decay has the contribution from three decays (1, 2 and 3 in the plot) shown here. 112 Figure 5-17 Plot of time constants derived from the triexponential fit of the experimentally observed current decay data. Figure 5-17 shows three distinctive trends of time constants. Assigning these time constants to certain electronic states and establishing a quantitative co-relation between the time constant and the driving force of electron trapping are complicated and beyond the scope of our current study. Multiple parameters like the charge mobilities, ease of charge neutrality by Li+ adsorption or insertion, distribution of the DOS etc. can have contributions into the characteristics of the current decay. However, we can make some reasonable hypotheses about the time constant. Presence of trap filling was reported to slow down the current decay during the current transient measurements.60 The electrons trapped on the surface can easily be charge compensated by Li+ present in the electrolyte whereas to compensate the charge in the bulk the Li+ is most likely need to be inserted into the crystal which involves bond reformation or/and bond elongation. Thus, an assumption can be made that the fastest current decay (curve 3 in Figure 5-16) was associated with the trapping of electrons on the surface. The corresponding electron density (4.61019 cm-3) was calculated and found to match well with the surface state density (nS) 113 calculated from spectroelectrochemical method (nS = 5.21019 cm-3). For the curve 1 in figure 516, this method provided a value of 7.61019 cm-3 whereas under the curve 2 the electron density was found to be 8.91019 cm-3. The spectroelectrochemical method gave us nCB = 6.71019 cm-3 and nB = 9.31019 cm-3 which are close to the values calculated under the curve 1 and 2. In all cases, the decay time shortens while moving towards more negative potential which can be explained by the larger driving force for the trap filling. 5.4 Conclusion SrTiO3 nanocrystals were synthesized by a vapor diffusion sol-gel method and mesoporous electrode was made. It was studied to compare its photoanode properties with anatase TiO2. The energetics and the electron distribution in mesoporous SrTiO3 electrode were measured by combining spectroelectrochemical method with electrochemical techniques like charge insertionextraction method and electrochemical impedance method. The band gap was found to be ~3.1 eV (~ band gap of anatase TiO2 is 3.2 eV) and the conduction band was found to be ~ 200 mV higher (more negative) than anatase. In SrTiO3 the free electron density in CB is comparable to the trapped electron density whereas in TiO2 the trapped electron density is orders of magnitude higher than CB electrons. -18 shows an energy level diagram of DSSC constructed with TiO2 and SrTiO3. The higher CB edge position in SrTiO3 provide an opportunity to push the electron Fermi level in semiconductor to a higher position than in the TiO2 photoanode. Also, the lower trap states in SrTiO3 can potentially offer lower possibility of trap state mediated recombination processes. Moreover, exploring a photoanode material having different energetics and electron distribution than TiO2 will help to understand the photoanode related electron transfer processes. 114 Figure 5-18 Comparison of the CB and electron distribution between TiO2 and SrTiO3. 115 REFERENCES 116 REFERENCES (1) Ardo, S.; Meyer, G. J. Chem. Soc. Rev. 2009, 38, 115–164. (2) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Chem. Rev. 2010, 110, 6595– 6663. (3) Youngblood, W. J.; Lee, S.-H. A.; Maeda, K.; Mallouk, T. E. Acc. Chem. Res. 2009, 42, 1966–1973. (4) Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q.; Santori, E. a.; Lewis, N. S. Chem. Rev. 2010, 110, 6446–6473. (5) Chen, X.; Shen, S.; Guo, L.; Mao, S. S. Chem. Rev. 2010, 110, 6503–6570. (6) Gust, D.; Moore, T. A.; Moore, A. L. Acc. Chem. Res. 2009, 42, 1890–1898. (7) Kay, A.; Cesar, I.; Grätzel, M. J. Am. Chem. Soc. 2006, 128, 15714–15721. (8) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Chem. Rev. 2010, 110, 6595– 6663. (9) Jayabal, P.; Sasirekha, V.; Mayandi, J.; Jeganathan, K.; Ramakrishnan, V. J. Alloys Compd. 2014, 586, 456–461. (10) Burnside, S.; Moser, J.; Brooks, K.; Gra, M.; Cahen, D. J. Phys. Chem. B 1999, 103, 9328–9332. (11) Yang, S.; Kou, H.; Wang, H.; Cheng, K.; Wang, J. J. Phys. Chem. C 2010, 114, 815–819. (12) Bera, A.; Wu, K.; Sheikh, A.; Alarousu, E.; Mohammed, O. F.; Wu, T. J. Phys. Chem. C 2014, 118, 28494–28501. (13) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. 2010, 6595–6663. (14) Jose, R.; Thavasi, V.; Ramakrishna, S. J. Am. Ceram. Soc. 2009, 92, 289–301. (15) Salehi, H. 2011, 2011, 934–943. (16) Cardona, M. Phys. Rev. 1965, 140. (17) Ai, X.; Guo, J.; Anderson, N. a.; Lian, T. J. Phys. Chem. B 2004, 108, 12795–12803. (18) Ai, X.; Anderson, N. A.; Guo, J.; Lian, T. J. Phys. Chem. B 2005, 109, 7088–7094. (19) Mandal, D.; Hamann, T. W. ACS Appl. Mater. Interfaces 2016, 8, 419–424. (20) Wrighton, M. S.; Ellis, A. B.; Wolczanski, P. T.; Morse, D. L.; Abrahamson, H. B.; Ginley, D. S. J. Am. Chem. Soc. 1976, 98, 2774–2779. 117 (21) Dabestani, R.; Bard, A. J.; Campion, A.; Fox, M. A.; Mallouk, T. E.; Webber, S. E.; White, J. M. J. Phys. Chem. 1988, 92, 1872–1878. (22) Yang, S.; Kou, H.; Wang, J.; Xue, H.; Han, H. J. Phys. Chem. C 2010, 114, 4245–4249. (23) Mandal, D.; Hamann, T. W. Phys. Chem. Chem. Phys. 2015, 17, 11156–11160. (24) Rabu, F. a; Lee, J. S.; Brutchey, R. L. 2012, 3114–3116. (25) Beier, C. W.; Cuevas, M. a.; Brutchey, R. L. J. Mater. Chem. 2010, 20, 5074. (26) Houng, B.; Liu, C. C.; Hung, M. T. Ceram. Int. 2013, 39, 3669–3676. (27) Jacques I. Pankove. Optical Processes in Semiconductors; 1975th ed.; Dover Publications, New York, 1975. (28) Rabuffetti, F. a; Brutchey, R. L. Dalton Trans. 2014, 43, 14499–14513. (29) Rabuffetti, F. a.; Brutchey, R. L. Chem. Commun. 2012, 48, 1437. (30) Rabuffetti, F. a.; Lee, J. S.; Brutchey, R. L. Adv. Mater. 2012, 24, 1434–1438. (31) Fabregat-santiago, F.; Mora-sero, I.; Bisquert, J. J. Phys. Chem. B 2003, 107, 758–768. (32) Bisquert, J. Phys. Chem. Chem. Phys. 2003, 5, 5360. (33) Kytin, V. G.; Bisquert, J.; Abayev, I.; Zaban, A. Phys. Rev. B 2004, 70, 193304. (34) Lenzmann, F.; Krueger, J.; Burnside, S.; Brooks, K.; Grätzel, M.; Gal, D.; Rühle, S.; Cahen, D. J. Phys. Chem. B 2001, 105, 6347–6352. (35) Gibbs, Z. M.; LaLonde, A.; Snyder, G. J. New J. Phys. 2013, 15, 75020. (36) Sarkar, a.; Ghosh, S.; Chaudhuri, S.; Pal, A. K. Thin Solid Films 1991, 204, 255–264. (37) Kim, S. S.; Park, J.; Kim, S. S.; Lee, Y.; Ahn, S.; Cho, J.; Ju, M.; Lakshminarayan, N.; Ai, V.; Yi, J. Vaccum 2014, 108, 39–44. (38) Hamberg, I.; Granqvist, C. G.; Berggren, K. F.; Sernelius, B. E.; Engström, L. Phys. Rev. B 1984, 30, 3240–3249. (39) Burstein, E. Phys. Rev. 1954, 93, 632–633. (40) Kamat, P. V.; Dimitrijevic, N. M.; Nozik, a. J. J. Phys. Chem. 1989, 93, 2873–2875. (41) Wunderlich, W.; Ohta, H.; Koumoto, K. Phys. B Condens. Matter 2009, 404, 2202–2212. (42) Wang, Y.; Yu, T.; Chen, X.; Zhang, H.; Ouyang, S.; Li, Z.; Ye, J.; Zou, Z. J. Phys. D. Appl. Phys. 2007, 40, 3925–3930. (43) Frye, A.; French, R. H.; Bonnell, D. A. Zeitschrift f?r Met. 2003, 94, 226–232. (44) Van Benthem, K.; Elsässer, C.; French, R. H. J. Appl. Phys. 2001, 90, 6156–6164. 118 (45) Waser, R.; Baiatu, T.; Hardtl, K.-H. J. Am. Ceram. Soc. 1990, 73, 1654–1662. (46) Li, Y.; Gao, X. P.; Pan, G. L.; Yan, T. Y.; Zhu, H. Y. J. Phys. Chem. C 2009, 113, 4386– 4394. (47) Kozuka, Y.; Kim, M.; Ohta, H.; Hikita, Y.; Bell, C.; Hwang, H. Y. Sci. Technol. 2010, 97, 1–14. (48) Choi, M.; Posadas, A. B.; Rodriguez, C. A.; O’Hara, A.; Seinige, H.; Kellock, A. J.; Frank, M. M.; Tsoi, M.; Zollner, S.; Narayanan, V.; Demkov, A. A. J. Appl. Phys. 2014, 116, 43705. (49) Perkins, P. G.; Winter, D. M. J. Phys. C Solid State Phys. 1983, 16, 3481–3492. (50) Calvani, P.; Capizzi, M.; Donato, F.; Lupi, S.; Maselli, P.; Peschiaroli, D. Phys. Rev. B 1993, 47, 8917–8922. (51) Wunderlich, W.; Ohta, S.; Ohta, H.; Koumoto, K.; Agency, T. 25th Int. Conf. Thermoelctrics 2005, 237–240. (52) Review, P. 1966, 147, 1956–1957. (53) Sopiha, K.; Malyi, O. I.; Persson, C.; Wu, P. Phys. Chem. Chem. Phys. 2017. (54) Schumacher, R.; Teschner, D. M.; Heinzel, a B. Berichte Der Bunsen-GesellschaftPhysical Chem. Chem. Phys. 1982, 86, 1153–1156. (55) Kavan, L.; Kratochvilová, K.; Grätzel, M. J. Electroanal. Chem. 1995, 394, 93–102. (56) Fabregat-Santiago, F.; Garcia-Belmonte, G.; Bisquert, J.; Zaban, A.; Salvador, P. J. Phys. Chem. B 2002, 106, 334–339. (57) Bisquert, J.; Garcia-Belmonte, G.; Fabregat-Santiago, F.; Ferriols, N. S.; Bogdanoff, P.; Pereira, E. C. J. Phys. Chem. B 2000, 104, 2287–2298. (58) Zaban, A.; Meier, A.; Gregg, B. a. J. Phys. Chem. B 1997, 101, 7985–7990. (59) Bisquert, J.; Garcia-Belmonte, G.; Fabregat-Santiago, F. J. Solid State Electrochem. 1999, 3, 337–347. (60) Wang, H.; He, J.; Boschloo, G.; Lindström, H.; Hagfeldt, A.; Lindquist, S.-E. J. Phys. Chem. B 2001, 105, 2529–2533. (61) Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S.-E. J. Phys. Chem. B 1997, 101, 7710–7716. (62) Hoyer, P.; Weller, H. J. Phys. Chem. 1995, 99, 14096–14100. (63) Schwarzburg, K.; Willig, F. Appl. Phys. Lett. 1991, 58, 2520–2522. (64) Solbrand, A.; Lindström, H.; Rensmo, H.; Hagfeldt, A.; Lindquist, S.-E.; Södergren, S. J. Phys. Chem. B 1997, 101, 2514–2518. 119 (65) Weppner, W. J. Electrochem. Soc. 1977, 124, 1569. (66) Call, R. W.; Alibabaei, L.; Dillon, R. J.; Knauf, R. R.; Nayak, A.; Dempsey, J. L.; Papanikolas, J. M.; Lopez, R. ACS Appl. Mater. Interfaces 2016, 8, 12282–12290. 120 Chapter 6 Predicting Maximum Attainable Efficiency of DSSC 6.1 Introduction The maximum theoretical power-conversion efficiency of single-junction solar cells is generally described by the Shockley–Queisser limit of ~33%.1 In addition to this, several multijunction technologies have been developed and as a result, power conversion efficiencies as large as 46% have recently been reported. The single junction photovoltaic cells constructed with Silicon and GaAs have both exceeded efficiencies beyond 25%.2 The small differences between the record photoconversion efficiencies and the Shockley–Queisser limit are attributed to a few practical concerns. In spite of the textured surface and use of an anti-reflective coating on the surface of solar cell, a small fraction of the sunlight is always reflected back. Some current loss occurs at the junctions and electrical contacts. The maximum power conversion efficiency of a Dye-sensitized Solar Cell (DSSC) reported so far (under 1 Sun illumination in laboratory) is 14.3%.3 The highest certified power conversion efficiency of a DSSC is 11.9% (Sharp, 2012), which is much lower than for Si-photovoltaics or thin film solar cells.4 The working principle and energy loss pathways of DSSCs are very different from the Si-solar cells and thin film solar cells. The mechanisms of the energy losses in DSSCs are yet to be optimized. The main advantage of dye-sensitized technology is that the photogenerated electrons and holes reside in different phases which hinders the electron-hole recombination mechanisms in the semiconductor. It dramatically increases the charge lifetime and the DSSC performs strikingly well in spite of the fact that the charge-carrier mobility across the nanostructured semiconductor photoanode is about one million times slower than in crystalline silicon.5,6 However, the DSSC system is complex and even after more than 25 years of research, predicting a roadmap towards 121 the best DSSC, in terms of photoconversion efficiency (), is still not feasible. A DSSC has many components and interfaces. Altering a single component from the optimized combination affects the other electron transfer processes which in turn necessitates the change of other components.7 In this chapter, our primary focus is to predict a reasonable target for the maximum photoconversion efficiency in the light of the recent advancement in our understanding of minimizing the energy-losses. 6.2 Calculating the power conversion efficiency of DSSC The photoconversion efficiency () of a solar cell is the ratio of the maximum electrical power obtained, Pmax, to the power of the incident light, Pin. %  Pmax J scVoc ff  Pin Pin (1) where JSC is the short-circuit current, VOC is the open circuit voltage and ff is the fill factor. Thus, instead of maximizing the current and voltage individually, an optimization should be made to get maximum value of current  voltage. 6.3 Energy losses in a DSSC The maximum power conversion efficiency of DSSC (Figure 6-1) achieved so far is 14.3% where JSC is 18.36 mA cm-2, VOC is 1.013 V and ff is 0.770.3 The fill factor, mainly attributed to the recombination losses and resistive losses, has nearly reached the optimal value.5 Incident photon-to-current conversion efficiency (IPCE) was reported to be ~91%. Remaining ~9% is due to reflection loss of incident light which can potentially be decreased by utilizing the textured surface and use of anti reflection coating. However, reflection loss always exists even in the record performing photovoltaics and large improvement in the IPCE beyond 91% is quite unlikely to be achieved. 122 Figure 6-1 J–V curve and IPCE plot of the DSSC photosensitized collaboratively by ADEKA-1 and LEG4. [Co(phen)3]2+/3+ redox shuttle was used. JV was measured AM-1.5, 100 mW cm2. This figure is reprinted from reference 3 (Kakiage, K.; Aoyama, Y.; Yano, T.; Oya, K.; Fujisawa, J.; Hanaya, M. Chem. Commun. 2015, 51, 15894–15897) with permission from Royal Society of Chemistry. 123 Figure 6-2 Plot of major energy losses in a typical DSSC. Negative potential represents higher energy. The mesoporous TiO2 photoanode is deposited on transparent conductive substrate (TCO). The conduction band (CB) edge potential of TiO2 photoanode is shown below the excited state potential of dye. EF is the Fermi level potential of electrons in the photoanode. R+/R represents the redox shuttles in the solution which regenerate the oxidized dye. VOC is the open circuit potential which is the potential difference between the solution potential and the Fermi level at the open circuit, EF,Voc. The loss-in-potentials, Vloss, are described in this chapter in details. Thus, this is very close to the maximum photocurrent possibly achieved with this absorption onset of the sensitized photoanode. The photoanode was co-sensitized with an alkoxysilyl anchor dye ADEKA-1 and a carboxy-anchor organic dye LEG4. Interestingly, the optical band gap of the dyes (LEG: 1.91 eV and ADEKA: 1.85 eV) used here is about 0.8 V higher than the VOC. Now, looking at the equation 1, all the parameters (JSC, ff) except VOC were optimized. Thus, the major losses in “efficient” DSSCs are reflected in V. We would like to use the parameter “lossin-potential, Vloss” here which describes the difference in energy between the optical bandgap of the sensitized photoanode and the open-circuit voltage produced by the solar cell. Vloss  Eopt q  VOC (2) where Eopt is the optical band gap of the dye, q is the charge of an electron. Vloss in the DSSC showed in the Figure 6-1 is ~0.8 V. Vloss for the previously champion DSSC was 0.88 V ( = 124 13%, 2014) which used a porphyrin dye and [Co(bpy)3]2+/3+ redox couple.4 The optical band gap of the dye was 1.79 eV, VOC = 0.91 V, JSC = 18.1 mA cm–2, ff = 0.78. The iodide/iodine based highly efficient DSSCs also suffer from >0.8 V voltage losses. Nazeeruddin et al. reported a 11.18% efficient DSSC (JACS, 2005) based on N719 dye (Di-tetrabutylammonium cisbis(isothiocyanato)bis(2,2’-bipyridyl-4,4’-dicarboxylato)ruthenium(II)) and iodine/iodide based redox shuttle where JSC = 17.730.5 mAcm-2, VOC = 0.8460.020 V, ff = 0.75 and Vloss was >0.9 V.8 These are probably the maximum efficiencies that can be achieved with this much of loss-inpotential.9 The loss-in-potential for state-of-the-art Si and GaAs solar cell have been reduced to ~0.3 V. In order to push the photoconversion efficiency of DSSC closer to Si or GaAs solar cells, the potential losses must be reduced significantly. The sources behind the overall potential losses are shown in Figure 6-2 and are described below. 6.3.1 Energy loss during photoexcited electron injection, Vloss, injection The ultrafast electron injection from the excited dye into the conduction band of the semiconductor photoanode is far from a straightforward electron transfer process between two energy states.10–18 Efficient electron injection depends on several factors including the position of the conduction band edge in the semiconductor (ECB), excited state energy of the dye, overlap between the occupied states of the dye and the empty density of states (DOS) of the semiconductor.19–22 Several injection processes take place from a range of excited states during the thermal relaxation pathway, e.g. injection from singlet states, intersystem crossing to the triplet states followed by injection into the semiconductor, etc. The injection rate can be optimized in several ways. Ardo and co-workers observed an increase in the electron injection from the excited state of Ru(bpy)2(dcb) (where dcb = 2,2′-bipyridyl-4,4′-dicarboxylate) to TiO2 in the presence of Li+ ions.22 Presence of small cations like Li+ lowers the ECB of TiO2 and as a 125 result, more acceptor states (TiO2 DOS) are thermodynamically accessible.23 The injection rate has been found to be increased in the presence of cations, Ca2+ > Ba2+ ≈ Sr2+ > Li+ > Na+ > K+ > Rb+ ≈ Cs+ ≈ TBA+ and also with a range of solution auto-ionization constant, MeOH > EtOH > aqueous pH 8 > DMF which is explained by the downward (more positive) shift of ECB in the same order.15,24–28 McCusker and co-workers studied the injection kinetics with cisRu(dcb)2(CN)2/TiO2 along with their osmium analogues and they observed that the injection is directly related to the dye excited-state reduction potential.29 Degree of photoanode crystallinity was also found to have an effect on this injection process. ‘Hot’ injection from upper vibrational excited states of the dye has been observed when the semiconductor conduction band edge is above the excited-state energy of the dye.30–32 Although this type of system is far from the ideal solar cell, it may compensate for the injection yield losses while minimizing the difference between the conduction band and the excited state energy of the dye. These observations imply that knowledge of the energetics in the semiconductor/dye interface is essential to optimize the “loss-in-potential”. A significant difficulty in studying the energetics at the semiconductor/dye interface has been the absence of effective method to determine the ECB accurately. A new spectroelectrochemical method to accurately measure ECB is presented in Chapter 2.33 Though predicting the minimum “loss-in-potential” in the injection process is difficult and complicated, we can, however, estimate the minimum difference between the CB band position in TiO2 and the excited state of the dye in an optimized DSSC. Figure 6-2 shows the energy alignment of a DSSC. The difference between the Fermi level of electron in the semiconductor under illumination and the solution potential defines the voltage output. At open circuit, the difference is the maximum and the corresponding voltage is called open circuit voltage, VOC. Using, iodine/iodide redox shuttle and ruthenium bipyridine sensitizer 126 (N719), more than 800 mV of open circuit voltage can be generated without compromising with the current output. Nazeeruddin et al. reported a DSSC where JSC = 17.73 mA cm-2, VOC = 84620 mV, ff = 0.75 and  = 11.18%.8 The IPCE was reported to be ~ 90%. Considering the reflection-loss of the incident light (~10%) this is the maximum possible current output based on this sensitizer. Thus, it can be assumed that the injection rate of the photoexcited electron from the dye excited state into the empty density of states (DOS) of TiO2 is “fast enough”. Now, solution potential, EF,0, of the iodine/iodide (1:10 concentration ratio) based redox shuttle is ~  0.11 V vs. SCE and the lowest excited state potential of the N719 dye is  0.9 V vs. SCE.18,34–36 In the scenario of VOC = 846 mV, EF,Voc can be estimated to be  0.736 vs. SCE. We measured the CB of TiO2 using a spectroelectrochemical method as described in Chapter 2. It is important to point out that our measurement was done with a mesoporous TiO2 film in contact with 0.1 M Li+ and dry acetonitrile was used as solvent. The energetics of the TiO2 film changes significantly with the type and concentration of the cations in the surrounding. Now, the electrolyte of an optimized solar cell generally contains 0.1 M Li+ in addition to tert-butylpyridine (generally 0.5 M is used). Other additives e.g. guanidinium thiocyanate are also added to improve the performance. The CB of TiO2 becomes ~200-300 mV higher in presence of tert-butylpyridine. 37– 39 Additionally, some protons are adsorbed during the dye loading process which can come from the protons of the carboxylic group of the dye and also from the solvent where the dye solution is made. The dye solutions are made both in protic (e.g. ethanol) and aprotic (e.g. acetonitrile) solvents and the dye adsorption is generally done on bench top where the photoanodes are exposed to the moisture as well as the trace amount of water impurities in the solvents. Adsorption of proton is known to make the TiO2 CB lower.38,40 Thus, predicting the CB position in the working DSSC is complicated. Our spectroelectrochemical measurements showed upward 127 movement of the CB while raising the Fermi level. In presence of 0.1M Li+/acetonitrile electrolyte and in the similar range of Fermi level, we found that the CB was 60 - 140 mV above the Fermi level. Thus, at the Fermi level calculated above ( 0.736 vs. SCE), the estimated range of the CB is  0.80 to  0.88 V vs. SCE. Now, considering the excited state position of the N719 dye ( 0.9 V vs. SCE) it can be stated that “fast enough” injection from dye to semiconductor is possible even when TiO2 CB edge is closely aligned to the dye excited state energy. We would like to point out that this value is only valid for TiO2 and N719 (or N3) dye. Bauer, Hagfeldt et al. reported electron injection from the Ru(tcterpy)(NCS)3 dye (tcterpy = 4,4′,4′′-tricarboxy2,2′:6′,2′′-terpyridine), known as black dye, to TiO2 where the excited state potential (triplet state) lies at the same potential than the TiO2 CB.41 This electron transfer rate depends on the extent of electronic coupling between the π* orbitals of the dye excited state and the empty 3d orbital of TiO2. Ruthenium based sensitizers have the advantage of efficient electron injection without significant Vloss,injection. However, recently, versatile organic dyes are being used in DSSCs having much higher excited state potentials. Their excited state potentials need to reside >300 mV above the TiO2 CB in order to achieve efficient electron injection. Kakiage and Hanaya et al. reported a DSSC where VOC is >1 V and the IPCE is >90% using [Co(phen)3]3+/2+ redox shuttles and a combination of alkoxysilylanchor dye ADEKA-1 and a carboxy-anchor organic dye LEG4.3,42 The excited states of these dyes are ~ 1.1 V vs. SCE and the solution potential, EF,0, is 0.34 vs. SCE. EF can be estimated to be 0.67 vs. SCE which implies a range of CB edge from 0.81 to 0.73 V vs. SCE. The difference between the dye excited state and the CB edge is 290-370 mV. Now, this difference is not necessarily universal and several factors can be tuned to reduce this as discussed in the early part of this section. For our discussion we can make a range of 0 to 300 mV for Vloss, injection. 128 6.3.2 Energy loss during regeneration of the oxidized dye, Vloss, regeneration The largest loss-in-potential in DSSC generally occurs during the regeneration of the oxidized dye, after the photoinduced electron injection into the semiconductor, by a redox shuttle. The regeneration efficiency is defined as the probability of an oxidized dye reduced by the redox shuttle before the recombination occurs with an electron in the semiconductor. Iodide/triiodide was used as the champion redox shuttle until recent advancements with cobalt complexes were achieved. Till date, iodide/triiodide had the major share of DSSCs research because of its slow recombination, high efficiency to regenerate the commonly used dyes e.g. N3, Z907 (reg ≈ 1) and excellent diffusion coefficient due to its small size and good solubility in commonly used electrolyte solvents.35,43–45 However, a driving force ∆G0 (= q.Vloss, regeneration, q = charge of an electron) is required for this process. Iodide/triiodide needs ~0.75 V to efficiently regenerate Ru(dcbpy)2(NCS)2 dye (N719) and ~0.6 V to efficiently regenerate Ru(tcterpy)(NCS)3, popularly known as Black dye. Vloss, regeneration for iodide/triiodide based DSSCs is thus huge and limits the overall photoconversion efficiency of a DSSC.8,46 The redox shuttle in the champion DSSC (shown in Figure 6-1), at present, is [Co(phen)3]3+/2+ and it needs 0.37 eV driving force to regenerate ADEKA-1 dye.3 Using the same TiO2-dye system, the regeneration driving forces for [Co(bpy)3]3+/2+ and [Co(Cl-phen)3]3+/2+ were found to be 0.43 eV and 0.27 eV, respectively.47 This large regeneration driving force is attributed to large reorganization energy and slow self-exchange kinetics.48 One possible way to minimize this driving force is to use a redox shuttle with much faster self exchange rate constant which can be achieved by minimizing the reorganization energy. However, the advantage of this large reorganization energy (1.73 eV associated with the electron transfer at TiO2/[Co(bpy)3]3+/2+ interface) is the slow recombination of electrons from the semiconductor to redox shuttles. 48 129 Ferrocene shows excellent dye-regeneration capability with much lower driving force, however, recombination is detrimental to the cell performance.49,50 An organic dye (carbz-PAHTDTT) has been shown to have regenerated efficiently by ferrocene and its derivatives with a driving force of 0.19 eV.49,51,52 The four orders of magnitude higher self-exchange rate constant compared to [Co(bpy)3]3+/2+ translates into approximately 100-fold faster regeneration kinetics which would allow a significant decrease in the driving force without affecting the efficient regeneration kinetics.53 Using a low spin cobalt based redox shuttle which has smaller barrier from low spin Co(II) (t2g6eg1) to low spin Co(III) (t2g6eg0) electron transfer, Xie et al. showed the potential of applying low-spin redox shuttles in DSSCs.54 They observed that a driving force of ~200 mV would be enough for excellent regeneration of MK-2 dye (2-Cyano-3-[5′′′-(9-ethyl-9H-carbazol3-yl)-3′,3′′,3′′′,4-tetra-n-hexyl-[2,2′,5′,2′′,5′′,2′′′]-quarter thiophen-5-yl] acrylic acid) by low spin cobalt bis-trithiacyclononane, [Co(ttcn)2]3+/2+.54,55 On the other hand, Ru(III)/Ru(II) based redox shuttles have also been demonstrated to have achieved good regeneration efficiency with minimal or nearly zero driving force, though, the performance is still affected by their low solubility, slow diffusion in commonly used electrolyte solvent and large light absorption behavior in the visible sprectrum.56 Yum et al. has reported a 10% efficient DSSC with an overpotential (= Vloss, regeneration) of 230 mV using an organic dye Y123.57 Based on these findings it seems plausible to push the regeneration overpotential (Vloss, regeneration) below 200 mV by utilizing dyes and redox couples with facile electron transfer kinetics. However, it also demands the blocking of recombination processes through careful surface treatments of photoanodes (e.g. deposition of Al2O3 or siloxane layer) or/and utilizing alternative photoanode materials.58–62 130 6.3.3 Recombinations losses, Vloss, recombination A very important feature of DSSC devices is the efficient transport of photoinjected electrons through the nanoparticle network in the photoanode to the conductive substrate, with minimal losses. Electrons can be lost from the nanoparticles as well as from the conductive substrate.7,63 As noted above, in order to improve the voltage by minimizing the dye-regeneration overpotential, outer-sphere redox shuttle with large self exchange rate constant needs to be utilized. Moving from I3-/I- to outer sphere redox shuttles offer a huge range of choices for dyeredox shuttle pairs. However, to realize their full potential, the origin of each of the recombination processes need to be studied and minimized. Figure 6-3 Schematic of back-electron transfer (i.e. recombination) processes from the semiconductor nanoparticles and the transparent conductive substrate (TCO) into the redox shuttle (R+/R) in the electrolyte. The photoanodes in DSSCs are generally made of TiO2 nanoparticles. The open-circuit voltage is generated by the potential difference between the redox potential of the electrolyte and the quasi-Fermi level of electrons, EF, in the semiconductor photoanode. Thus, moving the EF 131 closer to the conduction band edge (ECB) by populating the electron density will produce more voltage. These electrons come from the photoexcited dye and may be lost through the recombination processes. The processes are generally described in terms of electron lifetime, the average time an electron “survives” in the semiconductor electrode before recombining with acceptors in the electrolyte. Lowering (more positive) the redox potential of the redox shuttles also increases the driving force for recombination and facilitates trap states mediated recombination processes as depicted in Figure 6-4. Using the redox shuttle Ruthenium(II) bis(2,2’-bipyridyl)-bis(N-methylimidozole), [Ru(bpy)2(MeIm)2]3+/2+, with a very positive redox potential (> 0.8 V more positive than Co(bpy)33+2+ and I3-/I-) and small self-exchange (0.41 eV, about 25% of Co(bpy)33+/2+), Ondersma et al. observed that the back electron transfer from the trap states dominated the overall recombination processes.7 Figure 6-4 Energy diagram displaying the distribution of electrons in the TiO2 photoanode and the recombination processes from the conduction band (green) and from the trap states (red). The density of electrons in trap states in a TiO2 photoanode is about 2 orders of magnitude higher than for free electrons in the conduction band.64 Thus providing enough driving force for 132 the trap state mediated electron transfer processes by using highly positive redox shuttle would be detrimental to the cell performance. Interestingly, Ondersma et al. have observed a behavior consistent with Marcus inverted region of the recombination kinetics from conduction band to this redox shuttle.65 This demonstrates a pathway to achieve higher photovoltage by utilizing highly positive redox shuttle without making the recombination of CB electrons worse provided the trap state mediated recombination is controlled. One or two layers of Al2O3 are often used to block back-electron transfers and has been shown to improve electron diffusion length and hence charge collection efficiency in case of cobalt(II/III) complexes like [Co(bpy)3]3+/2+ and [Co(ttcn)2]3+/2+.48,54,60 However, it is not enough to circumvent the recombination processes while using redox shuttles having fast self exchange kinetics like [Co(ttcn)2]3+/2+ where the cell performance were found to be limited by the electron diffusion length. Making this insulating layer thicker is detrimental to the photoelectron injection from the dye. A better roadmap may be to find a new photoanode material which has higher interfacial charge transfer resistance and better charge transport properties. 1D structures (like nanorods) or arrays of nanowires which direct the electron transport to transparent conductive substrate (TCO) and thus improve overall charge collection efficiency may be utilized.66,67 Photoanode materials having lower trap states also need to be investigated. Additionally, Benkstein et al. showed the effect of the nanoparticle network geometry on the electron transport through mesoporous photoanode.68 They observed that transport slowed down with increasing porosity which was attributed to the decrease in the average coordination number of the particles leading to more dead ends. Thus, porosity adds one more parameter to the improvement of electron transport which leads to increasing charge collection. Using bulky dye is another way to reduce the recombination loss.69 There are several parameters controlling the recombination 133 processes and predicting the minimum “loss-in-potential” due to recombination loss is thus not feasible at present. Additionally, Ondersma et al. showed that while using a redox shuttle with very positive potential, blocking the recombination losses from the transparent conductive substrate (TCO) becomes more challenging.63 A thin compact layer of metal oxide, e.g. TiO2, is generally deposited on TCO to block recombination from the TCO to the redox shuttle which is sufficient when using traditional outer sphere redox shuttles like [Co(bpy)3]3+/2+ and similar Co3+/2+ complexes. However, when the driving force for the recombination electron transfer is larger, in case of highly positive redox shuttle, TiO2 layer is unable to stop this process. Ondersma et al. showed that in this scenario an insulating film such as poly-phenol oxide (PPO) is a good choice as the blocking layer on FTO.63 6.4 Predicting the maximum power conversion efficiency in DSSCs: Figure 6-5 Photon flux of the AM 1.5 G spectrum at 100 mW cm-2 (ASTM G173-03). 134 Solar energy is harvested by the dye adsorbed on the mesoporous semiconductor; harvesting more of the solar spectrum, shown in Figure 6-5, will result in higher photogenerated current densities. However, harvesting longer wavelengths of light, i.e. the lower energy region of the spectrum, also results in lower voltages. The solar spectrum shows strong atmospheric absorption features around 930, 1110 and 1320 nm which corresponds to 1.33, 1.12 and 0.94 eV, respectively. The photoconversion efficiency is a product of current and voltage (i x V). Thus, it is reasonable to propose that, in order to get optimum current without losing voltage considering these atmospheric absorption regions, the absorption onset (Eopt) of the sensitized semiconductor should begin around 930 or 1110 or 1320 nm as shown in Figure 6-6. Figure 6-6 Short-circuit photocurrent density as a function of absorption onset calculated by integrating the AM 1.5 solar spectrum. The maximum current attainable from the solar spectra by absorbing light up to 930, 1110 and 1320 nm are 33.9, 42.4 and 50.1 mA cm-2, respectively. The maximum incident photon to current conversion efficiency (IPCE) values in the champion DSSCs have reached as high as 91%.3 Considering ~10% loss of the incident light due to the reflection loss, this is the highest 135 limit of IPCE. Utilization of textured surface and anti-reflective coating can potentially lower down the reflection loss resulting higher IPCE. To the best of our knowledge, however, higher IPCE has not been achieved yet. Considering a IPCE = 90%, the short circuit current associated with the absorption onsets of 930, 1110 and 1320 nm are 30.5, 38.2 and 45.1 mA cm-2, respectively. The following sections will discuss how the other parameters of a working cell would be adjusted to maximize the photovoltage associated with each of these three photocurrent scenarios. A reasonable fill factor (ff) of 75% is assumed for all of the calculations.32,70–75 We point out that larger values of Voc also allow for larger ff. ff is defined by the ratio of the area under a J-V corresponding to the maximum power output to the area defined by Jsc and Voc. Increasing the VOC would increase the contribution of the smaller rectangle (J-V), provided the diode quality factor is not compromised. Thus, the ratio, i.e. the ff will increase. The ff shown in Figure 6-1 is 77% and the VOC is >1 V. However, as discussed in the following sections, the journey towards 20% photoconversion efficiency of a DSSC would likely come through the utilization of the near-infrared spectra of the sunlight which requires sensitizers of low optical band gap resulting low VOC (~ 0.7 V). Thus, it is reasonable to assume ff = 75% in our calculation. Note that, in case of Si PV, ff = 80% has been achieved with the same range of VOC. The maximum open circuit voltage, VOC can be written as VOC = Eopt/q – Vloss. The short circuit current, voltage and “loss-in-potential” corresponding to those 3 absorption onsets are shown in Table 6-1. In case of 930, 1110 and 1320 nm absorption onset scenarios, photoconversion efficiency,  = 20% is achievable with loss-in-potential as large as 450, 420 and 350 mV, respectively. As discussed in Section 2.3.1, the ruthenium based dyes like N719, N3, black dye can inject electron into TiO2 even when Vloss, injection = 0 V. Assuming Vloss, injection = 0 V and ECBEF = ~100 mV, 136 making 20% efficient DSSC looks plausible if Vloss, regeneration can be made <350, 320 and 250 mV, respectively, for the three absorption onset scenarios. It is a reasonable target as discussed in section 2.3.2. However, Vloss, injectron > 300 mV for the organic dyes used in the recent champion DSSCs. In a scenario of Vloss, injectron = 300 mV, Vloss, regeneration = 200 mV and ECBEF = ~100 mV it is not possible to go beyond 17% (see Figure 6-7). Table 6-1 Table of the photoconversion efficiencies at different absorbance onsets and the associated loss in potentials. Absorption onset / nm % JSC / mAcm-2 VOC / V Vloss / V 1.55 0.22 1.17 0.6 13 1.01 0.76 20 0.88 0.45 0.66 0.67 13 0.57 0.76 20 0.7 0.42 0.52 0.6 13 0.45 0.67 20 0.59 0.35 0.44 0.5 0.38 0.56 20 700 930 1110 1320 15 15 15 15 17.2 30.5 38.2 45.1 13 137 Figure 6-7 and Table 6-1 show that 15% efficient DSSCs can be made in all of the absorption onset scenarios mentioned here. The record efficiencies have already reached to 14.3%. However, harvesting the solar spectrum upto 930 nm would be the best choice to reach 15% efficiency because it would be achieved without much effort to decrease the loss in potential. Figure 6-7 Photoconversion efficiency of DSSC as a function of loss-in-potential at different absorption onsets of the sensitized photoanode. 138 Making a 20% efficient DSSC is, however, challenging. Using sensitizers with the absorption onset of 700 nm, it is very unlikely to make a DSSC having  = 20%. The total loss-in potential needs to be minimized to 220 mV and considering EF  ECB = ~100 mV, combination of Vloss, injection + Vloss, regeneration needs to be reduced to 120 mV which is very unlikely to happen. The near-infrared photons of the solar spectrum need to be harvested to achieve 20% photoconversion efficiency. We would also like to point out that Table 5-1 does not imply that targeting a higher energy absorption onset would offer an easier target for low loss-in-potential. The optimum photovoltage corresponding to higher energy onset is also higher which would provide some extra driving force for recombination processes. Suppressing these recombination processes to realize the maximum photovoltage output could be challenging. DSSC with an absorption onset 930 nm (i.e. 1.33 eV) needs a photovoltage of 0.88 V with a Vloss not larger than 0.45 V. However, for 1110 nm (i.e. 1.12 eV) the VOC is 0.7 V and Vloss is 0.42 V as shown in Figure 6-7. Thus, using a sensitizer having an absorbance onset of 1110 nm instead of 930 nm will face ~180 mV less driving force for recombination processes. However, Vloss, associated with both of them is almost same, ~450 mV. Thus, we believe, to achieve 20% efficiency, the scenario of 1110 nm absorption onset is likely to be a better choice than 930 nm scenario. Also, it is very unlikely that a single dye will absorb light throughout this whole wavelength region. A mixture of dyes need to be used to obtain maximum current from a working cell.4 The potential of 1320 nm onset scenario would only be realized by choosing a proper combination of dyes which is also not very straightforward. The current record efficiency is 14.3 % where Vloss, injectron = ~370 mV, Vloss, regeneration = 370 mV, ECBEF = ~60 mV which gives ~0.8 V overall potential loss. The optical band gap of the ADEKA-1 dye used there is 1.85 eV corresponding to ~670 nm. ~100 nm broadening of the 139 absorption width (IPCE, Figure 6-1) is present here which is generally observed when the sensitizer molecule is strongly coupled to the TiO2. It improves the overall light harvesting efficiency and Jsc was found to be as large as 18.36 mA cm-2. VOC was found to be 1.013 V which is the maximum attainable VOC considering ~0.8 V potential loss. Thus, with this combination of semiconductor photoanode, dye and redox shuttle, the maximum possible photoconversion efficiency has already been achieved, as shown in Figure 6-7. Vloss, injectron and Vloss, regeneration are required to be reduced in order to push the photoconversion efficiency beyond this.54,76 6.5 Conclusion In this chapter, the maximum attainable power conversion efficiency of DSSCs has been estimated. A plausible roadmap towards 20% photoconversion efficiency has been predicted. The main energy loss mechanisms have been pointed out and an outlook to minimize them has been presented. The challenges will only be met through extending our understanding of the interfacial charge transfer processes and mastering our ability to systematically engineer each of the components of the dye-sensitized solar cell. The journey of DSSC progressed keeping the initial champion components, like TiO2 photoanode, ruthenium dye, Iodide redox shuttle and platinum counter electrode, unchanged. In spite of the huge boom of DSSC research and a large effort to maximize the overall efficiency, it remained almost stagnant for two decades. The recent breakthrough in the efficiency has come through the systematic understanding and improvement of the outer sphere redox shuttles which showed pitiful performance initially.50 The absorption onset needs to be extended to near-infrared region by using proper combination of sensitizers. Additionally, great effort needs to be applied to develop a redox shuttle having very fast electron transfer kinetics which can efficiently regenerate oxidized dye with minimum driving force. However, it will also worsen the recombination mechanisms from the photoanode. Thus, 140 alternative photoanode materials need to be investigated which offer faster charge transport through nanostructured network and higher interfacial charge transfer resistance. An extended literature of various potential photoanode materials needs be developed. 141 REFERENCES 142 REFERENCES (1) Shockley, W.; Queisser, H. J. J. Appl. Phys. 1961, 32, 510–519. (2) Green, M. A.; Emery, K.; Hishikawa, Y.; Warta, W.; Dunlop, E. D. Prog. Photovoltaics Res. Appl. 2016, 24, 3–11. (3) Kakiage, K.; Aoyama, Y.; Yano, T.; Oya, K.; Fujisawa, J.; Hanaya, M. Chem. Commun. 2015, 51, 15894–15897. (4) Mathew, S.; Yella, A.; Gao, P.; Humphry-Baker, R.; Curchod, B. F. E.; Ashari-Astani, N.; Tavernelli, I.; Rothlisberger, U.; Nazeeruddin, M. K.; Grätzel, M. Nat. Chem. 2014, 6, 242–247. (5) Snaith, H. J. Adv. Funct. Mater. 2010, 20, 13–19. (6) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Chem. Rev. 2010, 110, 6595– 6663. (7) Ondersma, J. W.; Hamann, T. W. Coord. Chem. Rev. 2013, 257, 1533–1543. (8) Nazeeruddin, M. K.; De Angelis, F.; Fantacci, S.; Selloni, A.; Viscardi, G.; Liska, P.; Ito, S.; Takeru, B.; Grätzel, M. J. Am. Chem. Soc. 2005, 127, 16835–16847. (9) Snaith, H. J. Adv. Funct. Mater. 2010, 20, 13–19. (10) Koops, S. E.; Barnes, P. R. F.; O’Regan, B. C.; Durrant, J. R. J. Phys. Chem. C 2010, 114, 8054–8061. (11) Thavasi, V.; Renugopalakrishnan, V.; Jose, R.; Ramakrishna, S. Mater. Sci. Eng. R Reports 2009, 63, 81–99. (12) Koops, S. E.; O’Regan, B. C.; Barnes, P. R. F.; Durrant, J. R. J. Am. Chem. Soc. 2009, 131, 4808–4818. (13) Smeigh, A. L.; Katz, J. E.; Brunschwig, B. S.; Lewis, N. S.; McCusker, J. K. J. Phys. Chem. C 2008, 112, 12065–12068. (14) Tachibana, Y.; Haque, S. a; Mercer, I. P.; Durrant, J. R.; Klug, D. R. J. Phys. Chem. B 2000, 104, 1198–1205. (15) Tachibana, Y.; Nazeeruddin, M. K.; Grätzel, M.; Klug, D. R.; Durrant, J. R. Chem. Phys. 2002, 285, 127–132. (16) Haid, S.; Marszalek, M.; Mishra, A.; Wielopolski, M.; Teuscher, J.; Moser, J. E.; Humphry-Baker, R.; Zakeeruddin, S. M.; Grätzel, M.; Bäuerle, P. Adv. Funct. Mater. 2012, 22, 1291–1302. (17) Kuciauskas, D.; Monat, J. E.; Villahermosa, R.; Gray, H. B.; Lewis, N. S.; McCusker, J. K. 143 J. Phys. Chem. B 2002, 106, 9347–9358. (18) Haque, S. a.; Palomares, E.; Cho, B. M.; Green, A. N. M.; Hirata, N.; Klug, D. R.; Durrant, J. R. J. Am. Chem. Soc. 2005, 127, 3456–3462. (19) Staniszewski, A.; Morris, A. J.; Ito, T.; Meyer, G. J. J. Phys. Chem. B 2007, 111, 6822– 6828. (20) Knauf, R. R.; Brennaman, M. K.; Alibabaei, L.; Norris, M. R.; Dempsey, J. L. 2013, 1–8. (21) McCusker, J. K. Acc. Chem. Res. 2003, 36, 876–887. (22) Ardo, S.; Meyer, G. J. Chem. Soc. Rev. 2009, 38, 115–164. (23) Enright, B.; Redmond, G.; Fitzmaurice, D. J. Phys. Chem. 1994, 98, 6195–6200. (24) Watson, D. F.; Meyer, G. J. Coord. Chem. Rev. 2004, 248, 1391–1406. (25) She, C.; Guo, J.; Lian, T. J. Phys. Chem. B 2007, 111, 6903–6912. (26) Pollard, J. A.; Zhang, D.; Downing, J. A.; Knorr, F. J.; McHale, J. L. J. Phys. Chem. A 2005, 109, 11443–11452. (27) Konstantakou, M.; Stergiopoulos, T.; Likodimos, V.; Vougioukalakis, G. C.; Sygellou, L.; Kontos, A. G.; Tserepi, A.; Falaras, P. J. Phys. Chem. C 2014, 118, 16760–16775. (28) Asbury, J. B.; Anderson, N. A.; Hao, E.; Ai, X.; Lian, T. J. Phys. Chem. B 2003, 107, 7376–7386. (29) Kuciauskas, D.; Monat, J. E.; Villahermosa, R.; Gray, H. B.; Lewis, N. S.; McCusker, J. K. J. Phys. Chem. B 2002, 106, 9347–9358. (30) Listorti, A.; Regan, B. O.; Durrant, J. R. 2011, 3381–3399. (31) Lenzmann, F.; Krueger, J.; Burnside, S.; Brooks, K.; Grätzel, M.; Gal, D.; Rühle, S.; Cahen, D. J. Phys. Chem. B 2001, 105, 6347–6352. (32) O’Regan, B. C.; Durrant, J. R. Acc. Chem. Res. 2009, 42, 1799–1808. (33) Mandal, D.; Hamann, T. W. Phys. Chem. Chem. Phys. 2015, 17, 11156–11160. (34) Nazeeruddin, M. K.; Kay, A.; Rodicio, I.; Humphry-Baker, R.; Mueller, E.; Liska, P.; Vlachopoulos, N.; Graetzel, M. J. Am. Chem. Soc. 1993, 115, 6382–6390. (35) Boschloo, G.; Hagfeldt, A.; Spectus, C. O. N. Acc. Chem. Res. 2009, 42, 1819–1826. (36) Zhang, M.; Liu, J.; Wang, Y.; Zhou, D.; Wang, P. Chem. Sci. 2011, 2, 1401. (37) Boschloo, G.; Häggman, L.; Hagfeldt, A. J. Phys. Chem. B 2006, 110, 13144–13150. (38) Haque, S. A.; Tachibana, Y.; Willis, R. L.; Moser, J. E.; Grätzel, M.; Klug, D. R.; Durrant, J. R. J. Phys. Chem. B 2000, 104, 538–547. 144 (39) Schlichthörl, G.; Huang, S. Y.; Sprague, J.; Frank, a J. J. Phys. Chem. B 1997, 101, 8141– 8155. (40) Redmond, G.; Fitzmaurice, D. J. Phys. Chem. 1993, 97, 1426–1430. (41) Bauer, C.; Boschloo, G.; Mukhtar, E.; Hagfeldt, A. J. Phys. Chem. B 2002, 106, 12693– 12704. (42) Feldt, S. M.; Wang, G.; Boschloo, G.; Hagfeldt, A. J. Phys. Chem. C 2011, 115, 21500– 21507. (43) Clifford, J. N.; Palomares, E.; Nazeeruddin, K.; Gra, M.; Durrant, J. R. J. Phys. Chem. C 2007, 111, 6561–6567. (44) Rowley, J.; Meyer, G. J. J. Phys. Chem. Lett. 2009, 113, 18444–18447. (45) Nelson, J. J.; Amick, T. J.; Elliott, C. M. J. Phys. Chem. C 2008, 112, 18255–18263. (46) Hamann, T. W. Dalt. Trans. 2012, 41, 3111. (47) Kakiage, K.; Aoyama, Y.; Yano, T.; Otsuka, T.; Kyomen, T.; Unno, M.; Hanaya, M. Chem. Commun. 2014, 50, 6379. (48) Xie, Y.; Baillargeon, J.; Hamann, T. W. J. Phys. Chem. C 2015, acs.jpcc.5b08244. (49) Daeneke, T.; Kwon, T.-H.; Holmes, A. B.; Duffy, N. W.; Bach, U.; Spiccia, L. Nat. Chem. 2011, 3, 213–217. (50) Gregg, B. a.; Pichot, F.; Ferrere, S.; Fields, C. L. J. Phys. Chem. B 2001, 105, 1422–1429. (51) Daeneke, T.; Uemura, Y.; Duffy, N. W.; Mozer, A. J.; Koumura, N.; Bach, U.; Spiccia, L. Adv. Mater. 2012, 24, 1222–1225. (52) Daeneke, T.; Mozer, A. J.; Kwon, T.-H.; Duffy, N. W.; Holmes, A. B.; Bach, U.; Spiccia, L. Energy Environ. Sci. 2012, 5, 7090–7099. (53) Marcus, R. A.; Sutin, N. Biochim. Biophys. Acta - Rev. Bioenerg. 1985, 811, 265–322. (54) Xie, Y.; Hamann, T. W. J. Phys. Chem. Lett. 2013, 4, 328–332. (55) Soman, S.; Xie, Y.; Hamann, T. W. Polyhedron 2014, 82, 139–147. (56) Jiang, D.; Darabedian, N.; Ghazarian, S.; Hao, Y.; Zhgamadze, M.; Majaryan, N.; Shen, R.; Zhou, F. ChemPhysChem 2015, n/a-n/a. (57) Yum, J.-H.; Baranoff, E.; Kessler, F.; Moehl, T.; Ahmad, S.; Bessho, T.; Marchioro, A.; Ghadiri, E.; Moser, J.-E.; Yi, C.; Nazeeruddin, M. K.; Grätzel, M. Nat. Commun. 2012, 3, 631. (58) Klahr, B. M.; Hamann, T. W. J. Phys. Chem. C 2009, 113, 14040–14045. (59) Hamann, T. W.; Farha, O. K.; Hupp, J. T. J. Phys. Chem. C 2008, 112, 19756–19764. 145 (60) Ondersma, J. W.; Hamann, T. W. J. Phys. Chem. C 2010, 114, 638–645. (61) Palomares, E.; Clifford, J. N.; Haque, S. a; Lutz, T.; Durrant, J. R. J. Am. Chem. Soc. 2003, 125, 475–482. (62) Morris, A. J.; Meyer, G. J. J. Phys. Chem. C 2008, 112, 18224–18231. (63) Ondersma, J. W.; Hamann, T. W. 2011, 13361–13366. (64) Mandal, D.; Hamann, T. W. ACS Appl. Mater. Interfaces 2016, 8, 419–424. (65) Ondersma, J. W.; Hamann, T. W. J. Am. Chem. Soc. 2011, 133, 8264–8271. (66) Law, M.; Greene, L. E.; Johnson, J. C.; Saykally, R.; Yang, P. Nat Mater 2005, 4, 455– 459. (67) Varghese, O. K.; Paulose, M.; Grimes, C. A. Nat Nano 2009, 4, 592–597. (68) Benkstein, K. D.; Kopidakis, N.; van de Lagemaat, J.; Frank, A. J. J. Phys. Chem. B 2003, 107, 7759–7767. (69) Liu, Y.; Jennings, J. R.; Zakeeruddin, S. M.; Grätzel, M.; Wang, Q. J. Am. Chem. Soc. 2013, 135, 3939–3952. (70) Hauch, A.; Georg, A. Electrochim. Acta 2001, 46, 3457–3466. (71) Fabregat-Santiago, F.; Bisquert, J.; Garcia-Belmonte, G.; Boschloo, G.; Hagfeldt, A. Sol. Energy Mater. Sol. Cells 2005, 87, 117–131. (72) Peter, L. M. Phys. Chem. Chem. Phys. 2007, 9, 2630. (73) Zhu, K.; Jang, S.-R.; Frank, A. J. J. Phys. Chem. Lett. 2011, 2, 1070–1076. (74) Cameron, P. J.; Peter, L. M.; Hore, S. J. Phys. Chem. B 2005, 109, 930–936. (75) Ito, S.; Liska, P.; Comte, P.; Charvet, R.; Péchy, P.; Bach, U.; Schmidt-Mende, L.; Zakeeruddin, S. M.; Kay, A.; Nazeeruddin, M. K.; Grätzel, M. Chem. Commun. 2005, 4351. (76) Jiang, D.; Darabedian, N.; Ghazarian, S.; Hao, Y.; Zhgamadze, M.; Majaryan, N.; Shen, R.; Zhou, F. ChemPhysChem 2015, 16, 3385–3388. 146 Chapter 7 Conclusions and future directions We have developed a spectroelectrochemical method to determine the conduction band edge (ECB) in nanostructured photoanodes. ECB has been determined at different Fermi levels and an upward shift of the ECB position has been observed. This upward shift occurs due to the potential drop across the double layer (nanoparticle-electrolyte interface). The electron accumulation is accompanied by positive charges, i.e. the cations in the electrolyte. Accumulation of electrons in the electrically active states on the surface or within close vicinity of the surface of metal oxide nanoparticle changes the ionic distribution and hence the electric field in this interfacial double layer. The band edge unpinning has been mentioned in the nanostructured photoanode systems and has also been pointed out in other studies while studying the electron transfer processes in DSSCs.1,2 However, determination of the magnitude of this ECB shift has been missing in the literature. Hence, this phenomenon, in most of the studies related to DSSCs, has been ignored and in few cases, estimated indirectly. Thus, our work has solved a long standing problem in the context of studying the material properties of nanostructured semiconductor electrodes. Additionally, this method is versatile and can be applied to determine the band edge position in a variety of porous semiconductor materials provided a transparent film can be made. In addition to the conduction band position, trap state distribution is another important parameter required to study the various electron transfer processes related to the photoanodes. By combining spectroelectrochemical method with the electrochemical method we have determined the accurate electron distribution of nanoparticle photoanodes (TiO2 and SrTiO3). The magnitude of the band edge unpinning has been included to get a real picture of the electron distribution. The position of the trapped electrons is also important because the electrons trapped on the surface behave very differently from the electrons trapped in the bulk. However, their position 147 has been debated. Our studies have shown that the trap states are distributed on the surface as well as in the bulk and, with increasing negative bias, bulk traps dominate the overall trap state distribution. Such detailed understanding is important in studying the recombination reactions associated to the photoanode which generally limit the performance of DSSCs. A detailed study has been performed to determine the energetics and the charge distribution of mesoporous SrTiO3 electrode. Ternary oxides offer wider possibility for tuning the physical– chemical properties by altering the relative ratios of the cationic components. SrTiO3, SrSnO3, Zn2SnO4, CdSnO3 and BaSnO3 are among the ternary metal oxides utilized as photoanode in DSSCs.3–5 However, their performances have been much lower than the DSSCs constructed with TiO2 photoanode and further study needs to be performed to assess their potential as a DSSC photoanode material. TiO2 is being extensively used as photoanode material in DSSCs. Having another photoanode material, with different energetics and electron distributions from TiO2, will provide more choices for dyes and electrolytes. In comparison to TiO2, 200 mV higher conduction band edge position and a very different distribution of trapped states have been observed in SrTiO3. Overall density of trap states have been found to be lower than TiO2 which indicates a possibility of lower trap state mediated recombination processes. The charge separation in DSSC takes place at the dye-semiconductor interface. Separation between the solution potential, E(R+/R) and the Fermi level of the electron in the semiconductor, EF, is the voltage output from a cell. At the open circuit this separation is the highest and the corresponding voltage is called open circuit voltage, VOC. Thus, SrTiO3 could offer higher VOC if the recombination losses are minimized. The photoanode plays an essential role in the charge injection process from the dye, transporting these electrons and in recombination processes. The ideal photoanode should 148 demonstrate quantitative charge injection with minimal energy difference between the dye excited states and the conduction band edge. Additionally, the interfacial recombination processes also need to be minimized. To push the photoconversion efficiency of DSSC beyond 15%, a significant breakthrough is desired in the photoanode part of the device. One-electron, outer-sphere redox shuttles having ‘fast’ electron transfer kinetics can reduce the overpotential associated to the regeneration of the excited dye, which generally causes the largest voltage loss in a DSSC. Utilization of ‘fast’ redox shuttles also increases the electron recombination kinetics from nanoparticle to redox shuttle and results in a poor electron collection. In this context, photoanodes having faster electron transport and/or higher interfacial charge transfer resistance are highly desired. Exploring alternative photoanode material can lead us to the device configuration where the regeneration loss can be minimized by utilizing fast outer-sphere redox shuttles. 149 REFERENCES 150 REFERENCES (1) Liu, Y. R.; Jennings, J. R.; Zakeeruddin, S. M.; Gratzel, M.; Wang, Q. J. Am. Chem. Soc. 2013, 135, 3939–3952. (2) Liu, Y.; Jennings, J. R.; Zakeeruddin, S. M.; Grätzel, M.; Wang, Q. J. Am. Chem. Soc. 2013, 135, 3939–3952. (3) Li, Y.; Zhang, H.; Guo, B.; Wei, M. Electrochim. Acta 2012, 70, 313–317. (4) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Chem. Rev. 2010, 110, 6595– 6663. (5) Tan, B.; Toman, E.; Li, Y.; Wu, Y. J. Am. Chem. Soc. 2007, 129, 4162–4163. 151