INFLUENCE OF LIGAND EXCHANGE ON COPPER REDOX SHUTTLES IN DYE-SENSITIZED SOLAR CELLS By Eric James Firestone A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Chemistry – Doctor of Philosophy 2024 ABSTRACT Dye-sensi(cid:415)zed solar cells (DSSCs) are recognized as a promising, eco-friendly alterna(cid:415)ve to tradi(cid:415)onal photovoltaics, characterized by their unique light-harves(cid:415)ng capabili(cid:415)es and poten(cid:415)al for enhanced efficiency and stability in renewable energy applica(cid:415)ons. In this study, the focus is on the development of copper redox complexes, which exhibit varying responses to ligand exchange upon the introduc(cid:415)on of 4-tertbutylpyridine (TBP), a cri(cid:415)cal factor influencing electron transfer processes and the overall performance of DSSCs. Three copper complexes, copper(II/I) N,N(cid:3387)-Dibenzyl-N,N(cid:3387)-bis(6-methylpyridin-2-ylmethyl)ethylenediamine triflate, [Cu(dbmed)]OTf1/2, copper(II/I) 2,6-bis[1,1-bis(2-pyridyl)ethyl]pyridine triflate, [Cu(PY5)]OTf1/2 and copper(II/I) 6,6(cid:3387)- bis(1,1-di(pyridine-2-yl)ethyl)-2,2(cid:3387)-bipyridine bistriflimide, [Cu(bpyPY4)]TFSI1/2 were inves(cid:415)gated, with synthe(cid:415)c and electrochemical methodologies, including UV-Vis spectroscopy, NMR, and cyclic voltammetry, being u(cid:415)lized to examine the redox behavior and ligand exchange phenomena. The performance of DSSC devices with each redox mediator was measured to be 4.32%, 2.01%, and 1.23% respec(cid:415)vely. Methodology developed in this study, which involves using redox poten(cid:415)al as an indicator to predict ligand exchange events, represents an expansion and refinement of exis(cid:415)ng concepts found in the literature. By building on prior research, this approach not only deepens the understanding of copper complex behaviors in DSSCs but also offers a more nuanced perspec(cid:415)ve for enhancing the design and efficiency of these innova(cid:415)ve solar cells. Copyright by ERIC JAMES FIRESTONE 2024 ACKNOWLEDGMENTS Firstly I would like to thank my mentor, Thomas Hamann, for all the years he spent working to help me be(cid:425)er myself as a scien(cid:415)st and a chemist. Without his assistance I would not be the scien(cid:415)st I am today. I would also like to thank the rest of my commi(cid:425)ee, Dr. James McCusker, Dr. Aaron Odom, Dr. Mitch Smith, and Dr. Thomas O’Halloran, for their assistance in helping me understand mmy chemistry over the years. I want my fellow Hamannites, for all of the assistance both with the science and non-scien(cid:415)fic issues that have arisin over the years. I especially want to thank those that were on the DSSC project, Yujue, Aus(cid:415)n, Samhita, Abubaker, and Michael, whose constant discussions and advice greatly improved my understanding of my work. For those that were under my care these last three years thank you for being kind to my sugges(cid:415)ons and instruc(cid:415)on. I am certain I made mistakes, but you both were kind and pa(cid:415)ent through them all. I would also like to thank my friends both from before and during my (cid:415)me here at MSU, Josh, Karl, Mike, Jackie, Rachel, Julie, Adam, Connor, Nick, Ana, Courtney, Spencer, Shane, Hailey, Lizzie, Dominic, Cash, Ben, Jack, Ka(cid:415)e, Allison, Sophia, Maggie, Arzoo, Noel, Courtney, Rob, Tom, Veronica, each and every one of you will hold a special place in my heart, without you all I don’t think I would have been able to sane across these last few years. Finally I would like to thank my family, without whose support I would never be able to make this far in my pursuit of scien(cid:415)fic knowledge. From the bo(cid:425)om of my heart, thank you. iv TABLE OF CONTENTS Chapter 1 – Introduc(cid:415)on ................................................................................................................ 1 1.1 Mo(cid:415)va(cid:415)ons in solar energy ................................................................................................... 1 1.2 History of DSSCs .................................................................................................................... 3 1.3 Redox Shu(cid:425)les in DSSCs ........................................................................................................ 5 Chapter 2 – Inves(cid:415)ga(cid:415)on of [Cu(dbmed)](OTf)1/2 ....................................................................... 12 2.1 Introduc(cid:415)on ......................................................................................................................... 12 2.2 Experimental Details ........................................................................................................... 13 2.3 Results and Discussion ........................................................................................................ 18 2.4 Conclusion ........................................................................................................................... 25 Chapter 3 – Inves(cid:415)ga(cid:415)on of [Cu(bpyPY4)](TFSI)1/2 ..................................................................... 26 3.1 Introduc(cid:415)on ......................................................................................................................... 26 3.2 Experimental Details ........................................................................................................... 27 3.3 Results and Discussion ........................................................................................................ 32 3.4 Conclusions .......................................................................................................................... 47 Chapter 4 – Inves(cid:415)ga(cid:415)on of [Cu(PY5)]2+/+ ................................................................................... 49 4.1 Introduc(cid:415)on ......................................................................................................................... 49 4.2 Experimental Details ........................................................................................................... 51 4.3 Results and Discussion ........................................................................................................ 55 4.4 Conclusion ........................................................................................................................... 81 Chapter 5 – Developing a model to predict ligand exchange in DSSCs ...................................... 83 5.1 Introduc(cid:415)on ......................................................................................................................... 83 5.2 Experimental Details ........................................................................................................... 86 5.3 Results and Discussion ........................................................................................................ 89 5.4 Conclusions .......................................................................................................................... 98 WORKS CITED ............................................................................................................................... 99 APPENDIX A: CHAPTER 2 SUPPLEMENTARY DATA ..................................................................... 110 APPENDIX B: CHAPTER 3 SUPPLEMENTARY DATA ..................................................................... 114 APPENDIX C: CHAPTER 4 SUPPLEMENTARY DATA ..................................................................... 116 v Chapter 1 – Introduc(cid:415)on 1.1 Mo(cid:415)va(cid:415)ons in solar energy Energy is a cornerstone of modern society, and its consump(cid:415)on has profound implica(cid:415)ons on our environment, most notably in the form of increasing CO₂ emissions. Data from the past decade, specifically between 2010 and 2022, highlights a rise in CO₂ emissions by 1.0% annually, amoun(cid:415)ng to roughly 9.9 ± 0.5 GtC yr⁻¹. This has resulted in an atmospheric CO2 concentra(cid:415)on 51% higher than pre-industrial levels.1 An alarming 82% of these emissions can be traced back to fossil fuel consump(cid:415)on.2 This rapid surge in energy use, par(cid:415)cularly from fossil fuels, has resulted in energy-related CO₂ emissions surging to record highs, further underscoring the urgent need for carbon-neutral energy sources. If current trends in emission mi(cid:415)ga(cid:415)ons and renewable energy adop(cid:415)on persist, projec(cid:415)ons suggest a 3°C global temperature rise from pre-industrial levels by 2100.3 This poten(cid:415)al escala(cid:415)on intensifies the impera(cid:415)ve for a shi(cid:332) toward carbon-neutral energy solu(cid:415)ons. Solar energy, boas(cid:415)ng an overwhelming theore(cid:415)cal poten(cid:415)al of approximately 173,000 TW, dwarfs the current global demand of 2.9 TW and outpaces poten(cid:415)al outputs from other renewable sources like wind and hydro.4,5 Although renewable energy's contribu(cid:415)on to global power reached a commendable 12.8% in 2022,5 broader adop(cid:415)on is constrained by factors such as genera(cid:415)on costs and fabrica(cid:415)on complexity. 1 Figure 1.1: The 1.5 AM solar spectrum which represents the average amount of light that reaches the surface of the Earth across the con(cid:415)nental United States.6 Solar cells present a promising avenue in this regard. Their rising prevalence, combined with declining costs due to the SunShot Ini(cid:415)a(cid:415)ve, signifies the growing feasibility of solar energy. The SunShot Ini(cid:415)a(cid:415)ve is a DOE program with the goal to reduce the cost of solar energy by 75%, mee(cid:415)ng the 2020 goal by 2017.7 While the reduc(cid:415)on in fabrica(cid:415)on costs of tradi(cid:415)onal silicon solar cells has played a role in this price drop, further reduc(cid:415)ons face challenges chiefly due to stringent fabrica(cid:415)on requirements. The quest for more cost-effec(cid:415)ve solar solu(cid:415)ons necessitates the explora(cid:415)on of alterna(cid:415)ve photovoltaics with simpler fabrica(cid:415)on processes, reduced energy input, and the use of readily available materials. 2 Dye-Sensi(cid:415)zed Solar Cells (DSSCs) emerge as a frontrunner in this category. Their poten(cid:415)al affordability stems from their straigh(cid:414)orward fabrica(cid:415)on and the u(cid:415)liza(cid:415)on of earth-abundant materials. However, to truly posi(cid:415)on DSSCs as a game-changer in the renewable energy landscape, research must intensify around enhancing their efficiency and trimming material costs, thereby ensuring their economic viability and widespread adop(cid:415)on.8 1.2 History of DSSCs The legacy of DSSCs is based on the development of the theory of photoelectrochemistry, which can be traced back to Edmond Becquerel's groundbreaking work in 1839.9 Since the discovery of photoelectrochemistry, it has been used to develop methods of genera(cid:415)ng clean energy. Semiconductors are materials with specific bandgaps that can absorb photons of suitable energy. When they absorb these photons, they generate electron-hole pairs that can separate within the semiconductor and par(cid:415)cipate in redox reac(cid:415)ons on different surfaces. U(cid:415)lizing this idea, in 1873 and 1887 respec(cid:415)vely, Vogel and Moser independently explored "sensi(cid:415)za(cid:415)on" using silver halides mixed with visible-light-absorbing dyes.10 The theory was further elucidated almost a century later, in 1969, by Gerischer, who detailed how excited state dissolved dyes could inject electrons into ZnO semiconductor crystals.11 Further advancements were made with the anchoring of ruthenium and quinone sensi(cid:415)zers by two researchers, Weber and Matsumura in 1979 and 1980, respec(cid:415)vely.12,13 3 Figure 1.2: The composi(cid:415)on of a DSSC. Grӓtzel and his team exponen(cid:415)ally enhanced current genera(cid:415)on using polycrystalline anatase TiO2 films and the iodide/triodide redox shu(cid:425)le.14 By 1993, they achieved a 10% power conversion efficiency (PCE), thanks to refined TiO2 electrodes and expanded light-harves(cid:415)ng sensi(cid:415)zers.15 However, with minimal improvements beyond 10% PCE the progress plateaued un(cid:415)l 2010 when outer-sphere redox shu(cid:425)les became prevalent in DSSCs.16 The subsequent decade saw organic sensi(cid:415)zers paired with outer-sphere redox shu(cid:425)les, pushing PCEs to 15%.17–20 In the ensuing sec(cid:415)ons, the intricate opera(cid:415)ons of the dye-sensi(cid:415)zed solar cell will be delved into, followed by a discussion on the factors limi(cid:415)ng the PCE of these devices. DSSCs are a unique type of photovoltaic that distributes charge separa(cid:415)on and collec(cid:415)on over mul(cid:415)ple components, in contrast to conven(cid:415)onal crystalline semiconductor photovoltaics. In a 4 DSSC, a dye or sensi(cid:415)zer is anchored to semiconductor nanopar(cid:415)cles, usually ac(cid:415)ng as the primary light absorber. When this dye absorbs a photon, an electron is excited to a state where it can be injected into the semiconductor's conduc(cid:415)on band. The role of the semiconductor, o(cid:332)en materials like TiO2, is to transport these electrons to an FTO substrate (fluorine-doped (cid:415)n oxide). However, for the dye to con(cid:415)nue absorbing light, it needs another electron. This is where a redox shu(cid:425)le, dissolved in the electrolyte, comes into play. It diffuses to the dye's surface, dona(cid:415)ng an electron and ensuring the dye's readiness for subsequent photon absorp(cid:415)ons. The oxidized redox shu(cid:425)le then moves to the counter electrode to accept an electron, comple(cid:415)ng the circuit. The efficiency of a DSSC in conver(cid:415)ng sunlight to electrical power depends on several factors. The excited dye must have sufficiently nega(cid:415)ve poten(cid:415)al to inject its electron into the semiconductor. The injec(cid:415)on must be faster than excited dye decays or recombines with the semiconductor's electrons. The rate of dye regenera(cid:415)on should outpace its recombina(cid:415)on rate. Furthermore, electrons in the semiconductor need to be collected quickly to prevent their recombina(cid:415)on with the oxidized redox shu(cid:425)les. The interplay of these processes largely determines the DSSC's power conversion efficiency. 1.3 Redox Shu(cid:425)les in DSSCs DSSCs draw their essence from the ability to regenerate con(cid:415)nually. Central to this regenera(cid:415)ve ability is the redox couple. Predominantly found in a liquid electrolyte form, the redox couple— both in its reduced and oxidized states—acts as the linchpin, bridging conduc(cid:415)vity between the photoanode and the counter electrode. For the dynamic to work efficiently, the reduced form of the redox couple necessitates a rela(cid:415)ve nega(cid:415)ve reduc(cid:415)on poten(cid:415)al compared to the dye in its ground state. Once the dye molecule 5 dispatches an electron to the semiconductor, it undergoes oxida(cid:415)on. This is where the reduced redox couple, ac(cid:415)ng as an electron donor, plays its part. It regenerates the oxidized dye, priming it for the next photon absorp(cid:415)on. The pace at which this regenera(cid:415)on occurs is paramount—it contributes to the photocurrent's magnitude. Once it donates its electron, the reduced redox species becomes oxidized, primarily found near the photoanode. Through liquid electrolyte diffusion, this oxidized form dri(cid:332)s to the counter electrode to reclaim an electron from a catalyst surface, comple(cid:415)ng the cycle. Not only do redox couples play a vital kine(cid:415)c role, but their poten(cid:415)al also shapes the DSSC's photovoltage output. The open-circuit voltage (VOC) is defined by the energy difference between the Fermi level of the semiconductor and the solu(cid:415)on poten(cid:415)al of the redox shu(cid:425)le. The Fermi level of a semiconductor is defined by the equa(cid:415)on 𝐸(cid:3007) = 𝐸(cid:3030) − (cid:3038)(cid:3251)(cid:3021) (cid:3032) 𝑙𝑛 (cid:3041)(cid:3252) (cid:3015)(cid:3252) (1.1) where EC is the energy of the conduc(cid:415)on band minimum, kB is the Boltzmann constant, T is the temperature in Kelvin, e is the elementary charge of an electron, nC is the density of conduc(cid:415)on band electrons, and NC is the density of states in the conduc(cid:415)on band.21 The Fermi level is influenced by both the influx of electrons into the conduc(cid:415)on band and the rate at which electrons exit the conduc(cid:415)on band, whether through recombina(cid:415)on or charge separa(cid:415)on. By deploying diverse redox couples, one can skillfully modulate VOC, providing avenues for enhanced cell performance. 6 The equilibrium poten(cid:415)al of the liquid electrolyte at the counter electrode sets the cathode's energy level. A redox couple with a more posi(cid:415)ve redox poten(cid:415)al increases the cell's VOC, with a limit of the dye’s ground state poten(cid:415)al. In order to efficiently regenerate the dye, some energy, ac(cid:415)ng as a driving force for electron transfer, is inevitably lost. If the electron acceptor (the oxidized form of the redox couple) in the electrolyte has a more posi(cid:415)ve poten(cid:415)al than the Fermi Level, it can lead to electron recombina(cid:415)on. According to Marcus theory, faster electron transfer rates occur with larger driving forces in the normal region. However, redox couples that transfer electrons rapidly and have a higher poten(cid:415)al o(cid:332)en face electron recombina(cid:415)on, reducing the electron flow under equilibrium. Thus, adjus(cid:415)ng the redox proper(cid:415)es of the couple is crucial for enhancing solar cell performance. Triiodide/iodide redox couple's pioneering endeavors set a robust founda(cid:415)on for DSSC efficiency. These couples, due to their superior electron transfer kine(cid:415)cs, became instrumental in ensuring high photocurrent outputs. Studies, notably by Boschloo, delved deep into their mechanisms.22 Here, the iodide ion would regenerate the dye, subsequently genera(cid:415)ng the diiodide radical. This radical undergoes rapid dispropor(cid:415)ona(cid:415)on, yielding triiodide and iodide. The more posi(cid:415)ve reduc(cid:415)on poten(cid:415)al of triiodide results in a reduc(cid:415)on of electron recombina(cid:415)on, thereby increasing the cell's current output. But this comes at a price: energy loss. Despite their prowess, the intricacies of triiodide/iodide electron transfer mechanisms remain complex, promp(cid:415)ng researchers to explore alternate redox couples. Outer-sphere one electron transi(cid:415)on metal redox couples offer a more straigh(cid:414)orward electron transfer mechanism, sidestepping the complexi(cid:415)es seen in iodide/triiodide couples. An example is the cobalt(III/II) tri(2,2’-bipyridine) couple, which has gained immense trac(cid:415)on.23 Their poten(cid:415)al is exemplary, but 7 challenges persist, par(cid:415)cularly in low electron transfer rates which leads to high driving force to regenerate the dye. In summary, redox couples not only act as the heartbeat of DSSCs but also offer a rich tapestry of variables for researchers to manipulate. Balancing energy loss with efficient regenera(cid:415)on and electron transfer remains the ever-present challenge, poin(cid:415)ng towards a promising avenue of future explora(cid:415)on. Recent advancements in solar cell technology have spotlighted the promise of copper(II/I) redox couples. Following several years of research, the efficiency of solar cells featuring copper(II/I) complexes has soared, posi(cid:415)oning itself among the elite. Leading to the current highest efficiency DSSC devices to u(cid:415)lize copper(II/I) redox complexes.20 Remarkably, cells employing these couples consistently demonstrate an open-circuit voltage (VOC) exceeding 1 V, a notable leap from cells using cobalt complexes, which exhibit less posi(cid:415)ve redox poten(cid:415)als. Historically copper complexes were studied due to their role in electron transfer processes, both biologically and inorganically. Earlier studies gravitated towards understanding copper's role within metalloproteins.24 These proteins, o(cid:332)en sizable and challenging to study, necessitated the explora(cid:415)on of small molecule analogs. This eventually propelled studies into the redox chemistry of copper(II) coordina(cid:415)on complexes. In this context, the so-called "blue-copper" Type I protein garnered significant a(cid:425)en(cid:415)on due to its dis(cid:415)nc(cid:415)vely posi(cid:415)ve reduc(cid:415)on poten(cid:415)als, which hinted at the protein's u(cid:415)lity in natural electron transport systems.25 Intriguingly, these proper(cid:415)es also align perfectly with the requisites for solar energy conversion, paving the way for their adop(cid:415)on in this realm. 8 The ini(cid:415)a(cid:415)on of this paradigm shi(cid:332) can be a(cid:425)ributed to Ha(cid:425)ori in 2005, who introduced the poten(cid:415)al of copper complexes in DSSCs.26 The study delineated three diverse redox couples, all with unique ligand environments, and their compa(cid:415)bility with the N719 dye. Of par(cid:415)cular interest was the [Cu(SP)(mmt)]0/- complex, which mimicked blue-copper proteins. This seminal work laid the groundwork for subsequent researchers to unravel the intricacies of copper redox couples, ul(cid:415)mately culmina(cid:415)ng in Bai's breakthrough in 2011 that showcased a marked efficiency leap from 1.4% to 7.0%, by pairing a copper redox couple with an organic dye bound to the semiconductor.27 As researchers delved deeper into the interplay between dyes and copper redox couples, more nuanced findings emerged. Colombo's explora(cid:415)on with bulky asymmetric phenanthroline copper(II/I) complexes, for instance, highlighted the importance of careful electrolyte composi(cid:415)on.28 By 2016, a flurry of research showcased high-efficiency DSSCs using copper redox couples, with Cong's and Freitag's works standing out. Both explored different ligands and prepara(cid:415)on methods, yet achieved efficiencies hovering around 9%.29,30 Saygili's 2016 study further propelled the field, achieving over 10% efficiency with various copper bipyridyl redox couples.31 The pace of advancement in this domain did not abate, with Hu's 2018 study introducing a stable tetradentate copper(II/I) redox couple, which achieved an impressive efficiency of 9.2%.32 As of the current research landscape, the copper redox mediator copper(II/I) 4,4’,6,6’-tetramethyl-2,2’- bipyridine, [Cu(tmpy)2]2+/+, and the organic dye 3-{6-{4-[bis(2',4'-dihexyloxybiphenyl-4-yl)amino- ]phenyl}-4,4-dihexyl-cyclopenta-[2,1-b:3,4-b']dithiphene-2-yl}-2-cyanoacrylic acid, Y123, combina(cid:415)on remains at the pinnacle of DSSC performance.20 However, the quest for perfec(cid:415)on 9 con(cid:415)nues, as researchers strive to grasp the intricacies of electron transfer processes within these couples, all in a bid to op(cid:415)mize dye regenera(cid:415)on and electron recombina(cid:415)on kine(cid:415)cs. Base addi(cid:415)ves in DSSCs, such as pyridines and imidazoles, play a cri(cid:415)cal role in enhancing redox mediator performance.33,34 Lewis bases such as 4-tert-butylpyridine (TBP) are used as addi(cid:415)ves in DSSCs to increase the performance of the DSSC devices.35–37 This increase is a(cid:425)ributed to a shi(cid:332) in the (cid:415)tania conduc(cid:415)on band edge to a more nega(cid:415)ve poten(cid:415)al and a reduc(cid:415)on in recombina(cid:415)on from the (cid:415)tania surface by adsorbing to the (cid:415)tania.33,34 These addi(cid:415)ves can undergo an interac(cid:415)on with copper complexes upon oxida(cid:415)on, leading to the forma(cid:415)on of new complexes with altered proper(cid:415)es.38 It is currently being inves(cid:415)gated as to the extent of the interac(cid:415)on between the base addi(cid:415)ves and the copper complexes but there are two prevailing theories. The first is that a ligand subs(cid:415)tu(cid:415)on can occur, where, upon oxida(cid:415)on of the copper complex, the base addi(cid:415)ves will displace the ligand bound the copper complex forming a copper- base complex.38 The second hypothesis is that, upon oxida(cid:415)on, the base will coordinate to an open coordina(cid:415)on site on the copper center forming a copper-ligand-base complex.39 Through either path the presence of these addi(cid:415)ves can modify the redox poten(cid:415)al, stability, and kine(cid:415)cs of copper-based redox mediators. The interac(cid:415)on between base addi(cid:415)ves and copper redox mediators can significantly impact DSSC performance. Ligand subs(cid:415)tu(cid:415)on reac(cid:415)ons can influence the energe(cid:415)cs of electron transfer processes, affec(cid:415)ng charge injec(cid:415)on, recombina(cid:415)on, and overall cell efficiency.33,34,38 Understanding these effects is essen(cid:415)al for op(cid:415)mizing DSSC design and opera(cid:415)on. Previous research has demonstrated that ligand subs(cid:415)tu(cid:415)on reac(cid:415)ons can either enhance or hinder DSSC performance, depending on the specific copper complex and base addi(cid:415)ve 10 involved.38 Studying the factors that govern these outcomes provides valuable insights into the underlying mechanisms and informs the ra(cid:415)onal design of more efficient redox mediators. Developing a predic(cid:415)ve model for ligand subs(cid:415)tu(cid:415)on reac(cid:415)ons is essen(cid:415)al for advancing the design and op(cid:415)miza(cid:415)on of copper-based redox mediators in DSSCs. Such a model would enable researchers to predict the outcomes of different ligand subs(cid:415)tu(cid:415)on scenarios and guide the selec(cid:415)on of base addi(cid:415)ves and copper complexes to achieve desired performance enhancements. In this thesis, a comprehensive explora(cid:415)on of copper redox mediators in dye-sensi(cid:415)zed solar cells (DSSCs) is undertaken, focusing on their interac(cid:415)ons with base addi(cid:415)ves and the consequent implica(cid:415)ons for DSSC performance. The intricate interplay between copper complexes and base addi(cid:415)ves is thoroughly examined, with the goal of contribu(cid:415)ng to the fundamental understanding of ligand subs(cid:415)tu(cid:415)on processes and developing a predic(cid:415)ve model to aid in the ra(cid:415)onal design of efficient DSSCs. 11 Chapter 2 – Inves(cid:415)ga(cid:415)on of [Cu(dbmed)](OTf)1/2 2.1 Introduc(cid:415)on The demand for carbon-neutral energy genera(cid:415)on has accelerated the development of renewable technologies, par(cid:415)cularly in the domain of solar energy. Within this realm, dye- sensi(cid:415)zed solar cells (DSSCs) have been iden(cid:415)fied for their cost efficiency and photon-to-current conversion capabili(cid:415)es across diverse ligh(cid:415)ng condi(cid:415)ons.8 A standard DSSC consists of a chromophore anchored to a semiconduc(cid:415)ng mesoporous metal oxide surface. Upon photoexcita(cid:415)on, an electron is injected into the semiconductor's conduc(cid:415)on band. The oxidized chromophore is then regenerated by a redox shu(cid:425)le present in the solu(cid:415)on. This electron, a(cid:332)er traversing an external circuit, is eventually received by the counter electrode. Here, the oxidized redox shu(cid:425)le is reduced, comple(cid:415)ng the current circuit. Historically, the iodide/triiodide (I−/I3−) redox shu(cid:425)le system was predominantly used but presented challenges such as corrosiveness, non-tunability, and compe(cid:415)(cid:415)ve absorp(cid:415)on in the visible spectrum. Addi(cid:415)onally, the two-electron redox process of the I−/I3− system restricted the theore(cid:415)cal maximum photovoltage of DSSC devices.22 Co(III/II)-based redox shu(cid:425)les presented their own set of limita(cid:415)ons, primarily due to the significant inner-sphere reorganiza(cid:415)on energy during the electron transfer process, which affected dye regenera(cid:415)on kine(cid:415)cs.23 Copper-based redox shu(cid:425)les have been explored as an alterna(cid:415)ve due to their reduced inner- sphere reorganiza(cid:415)on energies for electron transfer. The geometrical adaptability of copper complexes, influenced by their oxida(cid:415)on state and ligand environment, is of significant interest. While many DSSCs with copper have primarily u(cid:415)lized bidentate ligands such as bipyridine or phenanthroline, there has been a shi(cid:332) towards tetradentate Cu(II/I) redox shu(cid:425)les, resul(cid:415)ng in 12 improved stability and reduced photovoltage losses.40,41 Innova(cid:415)ons by Freitag and associates, where a tetradentate Cu(II/I) redox shu(cid:425)le was integrated into DSSCs, have exhibited enhanced stability and reduced photovoltage losses.42 The rigidity of these tetradentate ligands can further lower reorganiza(cid:415)on energies, op(cid:415)mizing DSSC performance metrics. However, DSSC addi(cid:415)ves like TBP and N-methylbenzimidazole (NMBI), known to enhance cell performance, have demonstrated interac(cid:415)ons with copper-based redox shu(cid:425)les. Specifically, TBP has been observed to coordinate with Cu(II) species, leading to the displacement of bidentate ligands.38 This interac(cid:415)on impacts the redox poten(cid:415)al, electron transfer kine(cid:415)cs, and achievable photovoltage. Given these observa(cid:415)ons, there's an ongoing inves(cid:415)ga(cid:415)on into the poten(cid:415)al of tetradentate ligands to provide stability in the face of these challenges. In this chapter, the focus will be on the interac(cid:415)ons and stability of copper-based redox mediators in DSSCs, par(cid:415)cularly in the presence of base addi(cid:415)ves, and the poten(cid:415)al advantages offered by increased ligand den(cid:415)city. 2.2 Experimental Details All NMR spectra were taken on an Agilent DirectDrive2 500 MHz spectrometer at room temperature and referenced to residual solvent signals. All NMR spectra were evaluated using the MestReNova so(cid:332)ware package features. Cyclic voltammograms were obtained using µAutolabIII poten(cid:415)ostat using BASi glassy carbon electrode, a pla(cid:415)num mesh counter electrode, and a fabricated 0.01 M AgNO3, 0.1 M TBAPF6 in acetonitrile Ag/AgNO3 reference electrode. All measurements were internally referenced to ferrocene/ferrocenium couple via addi(cid:415)on of ferrocene to solu(cid:415)on a(cid:332)er measurements or run in a parallel solu(cid:415)on of the same solvent/electrolyte. UV-Vis spectra were taken using a PerkinElmer Lambda 35 UV-Vis 13 spectrometer using a 1 cm path length quartz cuve(cid:425)e at 480 nm/min. Elemental analysis data were obtained via Midwest Microlab. For single-crystal X-ray diffrac(cid:415)on, single crystals were mounted on a nylon loop with paratone oil using a Bruker APEX-II CCD diffractometer. Crystals were maintained at T ¼ 173(2) K during data collec(cid:415)on. Using Olex2, the structures were solved with the ShelXS structure solu(cid:415)on program using the Direct Methods solu(cid:415)on method. Photoelectrochemical measurements were performed with a poten(cid:415)ostat (Autolab PGSTAT 128N) in combina(cid:415)on with a xenon arc lamp. An AM 1.5 solar filter was used to simulate sunlight at 100 mW cm-2, and the light intensity was calibrated with a cer(cid:415)fied reference cell system (Oriel Reference Solar Cell & Meter). A black mask with an open area of 0.07 cm-2 was applied on top of the cell ac(cid:415)ve area. A monochromator (Horiba Jobin Yvon MicroHR) a(cid:425)ached to the 450 W xenon arc light source was used for monochroma(cid:415)c light for IPCE measurements. The photon flux of the light incident on the samples was measured with a laser power meter (Nova II Ophir). IPCE measurements were made at 20 nm intervals between 400 and 700 nm at short circuit current. TEC 15 FTO was cut into 1.5 cm by 2 cm pieces which were sonicated in soapy DI water for 15 minutes, followed by manual scrubbing of the FTO with Kimwipes. The FTO pieces were then sonicated in DI water for 10 minutes, rinsed with DI water, and sonicated in isopropanol for 15 minutes. The FTO pieces were dried under a stream of air then immersed in an aqueous 40 mM solu(cid:415)on TiCl4 solu(cid:415)on for 30 minutes at 70 °C. The water used for the TiCl4 treatment was preheated to 70 °C prior to adding 2 M TiCl4 to the water. The 40 mM solu(cid:415)on was immediately poured onto the samples and placed in a 70 °C oven for the 30-minute deposi(cid:415)on. The FTO pieces were immediately rinsed with 18 MΩ water followed by isopropanol and were annealed by hea(cid:415)ng from room temperature to 500 °C, holding at 500 °C for 15 minutes. A 0.36 cm2 area was 14 doctor bladed with commercial 30 nm TiO2 nanopar(cid:415)cle paste (DSL 30NRD). The transparent films were placed in a 100 °C oven for 30 minutes. The samples were annealed in an oven that was ramped to 325 °C for 5 minutes, 375 °C for 5 minutes, 450 °C for 5 minutes, and 500 °C for 15 minutes. The 30 nm nanopar(cid:415)cle film thickness was 8.2 μm. A(cid:332)er cooling to room temperature, a second TiCl4 treatment was performed as described above. When the anodes had cooled to room temperature, they were soaked in a dye solu(cid:415)on of 0.1 mM Y123 in 1:1 acetonitrile:tert- butyl alcohol for 18 hours. A(cid:332)er soaking, the anodes were rinsed with acetonitrile and were dried gently under a stream of nitrogen. The PEDOT counter electrodes were prepared by electrodeposi(cid:415)on in a solu(cid:415)on of 0.01 M EDOT and 0.1 M LiClO4 in 0.1 M SDS in 18 MΩ water. A constant current of 8.3 mA for 250 seconds was applied to a 54 cm2 piece of TEC 15 FTO with predrilled holes using an equal size piece of FTO as the counter electrode. The PEDOT electrodes were then washed with DI water and acetonitrile before being dried under a gentle stream of nitrogen and cut into 1.5 cm by 1.0 cm pieces. The working and counter electrodes were sandwiched together with 25 μm surlyn films by placing them on a 140 °C hotplate and applying pressure. The cells were then filled in a nitrogen filled glove box with electrolyte through one of the two predrilled holes and were sealed with 25 μm surlyn backed by a glass coverslip and applied heat to seal with a soldering iron. The electrolyte consisted of 0.25 M Cu(I), 0.065 M Cu(II), 0.1 M Li(Counter-ion), in acetonitrile with either no 4- tert-butylpyridine or 0.5 M 4-tert-butylpyridine added. All batches were made with at least 10 cells per batch. Contact to the TiO2 electrode was made by soldering a thin layer of indium wire onto the FTO. Synthesis of N,N(cid:3387)-Dibenzyl-N,N(cid:3387)-bis(6-methylpyridin-2-ylmethyl)ethylenediamine (dbmed): 15 2-(Chloromethyl)-6-methylpyridine hydrochloride (0.5300g, 3.0 mmol) and benzyltriethylammonium chloride (0.0121 g, 0.05 mmol) were added to 20 mL dichloromethane and s(cid:415)rred un(cid:415)l fully dissolved. N,N’-Dibenzylethylenediamine (0.35 mL, 1.48 mmol) was then added to the solu(cid:415)on, a(cid:332)er which a cloudy white solu(cid:415)on was formed. Meanwhile sodium hydroxide (20.1807 g, 500 mmol) was added to 20 mL of water which was then added to the dichloromethane solu(cid:415)on and s(cid:415)rred vigorously. A(cid:332)er 2 hours the organic solu(cid:415)on became clear and the aqueous solu(cid:415)on turned white. The solu(cid:415)on was then allowed to reflux overnight. A(cid:332)er refluxing addi(cid:415)onal water was added to the solu(cid:415)on to allow the layers to fully separate then the organic layer was then separated from the aqueous layer. The aqueous layer was then extracted with dichloromethane two (cid:415)mes and all the organic layers were combined, and dried with magnesium sulfate. Then the solvent was removed under low vacuum. The crude solid was then recrystallized in acetonitrile and dried to obtain the product (0.4270 g, 60.1% yield). 1H-NMR (CDCl3 500 MHz) δ 2.50 (s, 6H), 2.67 (s, 4H), 3.56 (s, 4H), 3.67 (s, 4H), 6.96 (d, 2H), 7.25 (m, 12H), 7.45 (t, 2H). Matching the data of the compound made by Hu et al.40 Synthesis of [Cu(dbmed)]OTf: Copper(I) tetrakisacetonitrile triflate (0.1123 g, 0.30 mmol) was dissolved in minimal dry acetonitrile. Meanwhile dbmed (0.1487 g, 0.33 mmol) was dissolved in minimal dry dichloromethane. The dbmed solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a bright yellow color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(I)dbmed product was collected (0.1876 g, 94.9% yield). 1H-NMR (CDCl3 500 MHz) δ 2.67 (s, 2H), 2.72 (s, 6H), 2.82 (s, 2H), 3.57 (s, 16 2H), 3.74 (s, 2H), 3.85 (s, 2H), 4.16 (s, 2H), 7.20 (m, 4H), 7.29 (m, 8H), 7.43 (d, 2H), 7.83 (t, 2H). Matching the data of the compound made by Hu et al.40 Synthesis of Cu(II)dbmed: Copper(II) triflate (0.808 g, 2.23 mmol) was dissolved in minimal dry acetonitrile. Meanwhile, dbmed (0.1030 g, 2.29 mmol) was dissolved in minimal dry dichloromethane. The dbmed solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning a deep blue color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(II)dbmed product was collected (0.1596 g, 87.9% yield). Matching the data of the compound made by Hu et al.40 17 2.3 Results and Discussion Scheme 2.1: Syntheses of the [Cu(dbmed)]OTf1/2 complexes with the Cu(I) species (le(cid:332)) and Cu(II) species (right). The ligand, N,N(cid:3387)-Dibenzyl-N,N(cid:3387)-bis(6-methylpyridin-2-ylmethyl)ethylenediamine (dbmed), and the copper ligand complexes, [Cu(dbmed)]OTf and [Cu(dbmed)]OTf2 were synthesized following a previously reported protocol.32 A crystal was grown of the copper(I) form, the copper center is coordinated to the two amine and two pyridyl nitrogen’s, resul(cid:415)ng in a tetrahedral geometry which is common for copper(I) complexes. This crystal matched the crystal shown in the literature.40 A(cid:425)empts at growing a crystal of the copper(II) form did not work out, forming an oil instead of crystals. However, it has been previously shown that the copper(II) form is in the 18 distorted tetrahedral geometry due to strain caused by the methyl groups on the 6 posi(cid:415)on of the pyridine.40 It is also noted that the copper(II) complex is air stable, but the copper(I) is not, it oxidizes over the course of a few hours in solu(cid:415)on or a few days in the solid state. The dbmed ligand was characterized via 1H-NMR and both copper complexes were characterized via 1H-NMR, and elemental analysis, matching previous reports, shown in Figures 2.1A, 2.2A, and 2.3A. The effect of TBP on the [Cu(dbmed)]2+/+ redox couple was evaluated in the presence of different concentra(cid:415)ons of the Lewis base. TBP was added in excess of at least 15 equivalents rela(cid:415)ve to [Cu(dbmed)](OTf)2 in solu(cid:415)on to match the condi(cid:415)ons used in the devices. e c n a b r o s b A 0.6 0.5 0.4 0.3 0.2 0.1 0 450 550 650 750 850 Wavelength / nm 950 1050 Figure 2.3: UV-Vis spectra of 2.14 mM [Cu(dbmed)]OTf2, pale green, with increasing concentra(cid:415)ons of TBP added, from 0 mM to 53.5 mM with the darker green being higher concentra(cid:415)on of TBP added, compared to a spectra of 2.14 mM [Cu(TBP)4]OTf2 in acetonitrile, purple. The effect of TBP was evaluated using UV−vis spectroscopy, containing 2 mM of [Cu(dbmed)](OTf)2, where spectra were recorded as a func(cid:415)on of added TBP. As TBP was added to a solu(cid:415)on of [Cu(dbmed)]OTf2 and monitored using UV-Vis spectroscopy. It can be seen that the absorbance of the peak at ~650 nm decreases in intensity and blueshi(cid:332)s slightly. However 19 even at 25 equivalents of TBP added to the solu(cid:415)on the spectra does not match that of [Cu(TBP)4]OTf2. A μ / t n e r r u C 30 20 10 0 -10 -20 -30 -1 -0.8 -0.6 -0.2 -0.4 0 Potential Applied vs Ferrocene / V 0.2 0.4 0.6 0.8 1 Figure 2.4: Cyclic voltammogram of 1.6 mM [Cu(dbmed)]OTf2 in acetonitrile containing 0.1 M LiOTf using a glassy carbon working electrode and a pla(cid:415)num mesh counter electrode and a scan rate of 100 mV/s. Increasing concentra(cid:415)ons of TBP were added from 0 mM to 38 mM, the darker green the higher the TBP concentra(cid:415)on. The results were compared with the cyclic voltammogram of 2.0 mM free dbmed ligand in red. As a follow up experiment, the addi(cid:415)on of TBP to a solu(cid:415)on of [Cu(dbmed)]OTf2 was monitored via cyclic voltammetry. As it can be seen in Figure 2.2, the cathodic peak of [Cu(dbmed)]OTf2, at -0.050 V vs ferrocene decreases in intensity and a second cathodic wave around -0.250 V vs ferrocene grows in as TBP is added to the solu(cid:415)on. While the cathodic peak is undergoing a dras(cid:415)c change, the anodic peak does not experience significant change nor shi(cid:332) in poten(cid:415)al. Addi(cid:415)onally, a peak appears to be growing in around 0.600 V vs ferrocene which corresponds to the redox poten(cid:415)al of the free ligand. This observa(cid:415)on suggests that dbmed is displaced by TBP upon oxida(cid:415)on, forming [Cu(TBP)4]OTf2. However, upon reduc(cid:415)on, the process reverses: TBP is displaced by the dbmed ligand, reforming [Cu(dbmed)]OTf. This demonstrates that the ligand 20 subs(cid:415)tu(cid:415)on is fully reversible. This indicates that that as TBP is added to the solu(cid:415)on some of the previously bound dbmed ligand is being released from copper and is able to be oxidized. Figure 2.5: 1H-NMR of 9.85 mM [Cu(dbmed)]OTf2, black, in deuterated acetonitrile with a known amount of dichloromethane added. Various concentra(cid:415)ons of TBP were added, 18 mM, in grey, 50 mM, in purple, and 100 mM, in green. The spectra were compared to a spectra of free dbmed ligand, red. To determine the extent of the displacement of dbmed unambiguously and quan(cid:415)ta(cid:415)vely by TBP, proton NMR spectra were taken of [Cu(dbmed)]OTf2 in deuterated acetonitrile without and with the addi(cid:415)on of TBP. Because [Cu(dbmed)]2+ is paramagne(cid:415)c, the proton signals of the dbmed ligands coordinated to the copper(II) center show very different chemical shi(cid:332)s than free dbmed ligands in the solu(cid:415)on. The sample was measured with up to 100 equivalents of TBP added, however as the concentra(cid:415)on of TBP was added to the solu(cid:415)on there were issues with the 21 shimming due to the high intensity of the methyl group on the TBP, as a result the data shown is limited to 50 equivalents of TBP added and lower. As TBP is added to the solu(cid:415)on, four new peaks appear in the NMR spectra between 2.4 and 3.8 ppm. These peaks match with the peaks of unbound dbmed ligand. By u(cid:415)lizing an internal standard of a known amount of DCM, the amount of free ligand in solu(cid:415)on can be quan(cid:415)ta(cid:415)vely determined. Using this method, it was found that the [Cu(dbmed)](OTf)2 had not fully undergone ligand subs(cid:415)tu(cid:415)on by the (cid:415)me 50 equivalents of TBP was added to the solu(cid:415)on. 2 - m c A m / t n e r r u C 10 9 8 7 6 5 4 3 2 1 0 -1 -0.8 -0.6 -0.4 -0.2 0 Voltage / V Figure 2.6: J−V curves for the [Cu(dbmed)]2+/+ DSSC devices, with TBP, dark green, and without TBP light green. Table 2.1: A summary of DSSC device performance metrics for those devices. η (%) Jsc (mA*cm- 2) Voc (V) FF Cu(dmbpy) 0.85(±0.35) 2.19(±0.77) 0.58(±0.02) 0.66(±0.05) Cu(dmbpy) with TBP 4.32(±0.64) 7.19(±1.24) 0.89(±0.02) 0.67(±0.05) 22 DSSC devices were fabricated using [Cu(dbmed)]2+/+ as the redox shu(cid:425)le in conjunc(cid:415)on with a poly(3,4-ethylenedioxythiophene) (PEDOT) counter electrode, and the commercially available dye Y123 was used as the light-harves(cid:415)ng component. The devices were constructed as described previously. The devices with TBP added to the solu(cid:415)on yielded a respectable performance with a JSC of 7.19(±1.24) mAcm-2 and an overall PCE of 4.32(±0.64) as shown in Table 2.1. However, in the absence of TBP the performance was significantly worse, with a JSC of only 2.19(±0.77) mAcm-2 and an overall PCE of 0.85(±0.35). This indicates that TBP is s(cid:415)ll performing a pivotal role in increasing the current of the devices. Table 2.2: Comparison of the solu(cid:415)on poten(cid:415)al and the predicted Nerns(cid:415)an poten(cid:415)al of the electrolyte solu(cid:415)on used in each DSSC device set. E1/2 of [Cu(dbmed)]OTf1/2 / V -0.005 Nernst equa(cid:415)on / V -0.023 Solu(cid:415)on poten(cid:415)al /V -0.053 Ef Level / V -0.633 -0.005 -0.023 -0.260 -1.150 dbmed dbmed with TBP 23 10 1 0.1 ) s / 1 ( e m i t e f i L 0.01 0 [Cu(dbmed)] without TBP [Cu(dbmed)] with TBP 0.2 0.4 0.6 0.8 1 1.2 Voltage (V) Figure 2.7: A plot of the electron life(cid:415)me vs poten(cid:415)al plot for DSSC devices using [Cu(dbmed)]OTf1/2 (green), with (pale color) and without (dark color) the addi(cid:415)on of TBP. By inves(cid:415)ga(cid:415)ng the solu(cid:415)on poten(cid:415)al of electrolyte composi(cid:415)ons, it is observed that the solu(cid:415)on poten(cid:415)al becomes about 200 mV more nega(cid:415)ve with the addi(cid:415)on of TBP to the electrolyte. Even though there was a nega(cid:415)ve shi(cid:332) in solu(cid:415)on poten(cid:415)al, there was a 300 mV increase in the open circuit voltage, indica(cid:415)ng that the fermi level of the semiconductor increased by around 500 mV with the introduc(cid:415)on of TBP to the devices. This is a further indicator of ligand subs(cid:415)tu(cid:415)on occurring and forming a more nega(cid:415)ve redox couple in solu(cid:415)on. Furthermore, the electron life(cid:415)me also significantly increases with the addi(cid:415)on of TBP, to the extent that it is hard to compare the results since there are no overlapping measured poten(cid:415)als that the (cid:415)mescale of the experiment. This dras(cid:415)c increase in poten(cid:415)al and electron life(cid:415)me can be a(cid:425)ributed to the addi(cid:415)on of TBP, more specifically both the proper(cid:415)es of Lewis bases in DSSC electrolyte (shi(cid:332)ing 24 the conduc(cid:415)on band poten(cid:415)al and reducing recombina(cid:415)on) and ligand exchange occurring and forming a [Cu(TBP)4)]OTf1/2 complex which has very poor electron kine(cid:415)cs.33,34 2.4 Conclusion In this study, the [Cu(dbmed)]OTf2 complex was found to be suscep(cid:415)ble to ligand subs(cid:415)tu(cid:415)on in the presence of TBP, without measurable full ligand subs(cid:415)tu(cid:415)on occurring when up to 50 equivalents of TBP were added to the solu(cid:415)on. Interes(cid:415)ngly, this ligand subs(cid:415)tu(cid:415)on was observed to be reversible upon reduc(cid:415)on of the complex. Despite the occurrence of ligand exchange, the addi(cid:415)on of TBP notably enhanced the efficiency of the DSSC devices. Allowing for dbmed to con(cid:415)nue to be a viable ligand for copper redox shu(cid:425)les in DSSCs in the future. The underlying cause of this ligand subs(cid:415)tu(cid:415)on was determined to be the steric strain imposed by the methyl groups on the pyridines of the complex, which effec(cid:415)vely outcompetes the thermodynamic stability usually conferred by the increased den(cid:415)city of the ligand. 25 Chapter 3 – Inves(cid:415)ga(cid:415)on of [Cu(bpyPY4)](TFSI)1/2 3.1 Introduc(cid:415)on The necessity for carbon-neutral energy genera(cid:415)on has driven the advancement of novel technologies that exploit renewable sources, notably wind and solar energy.43,44 By harnessing sunlight for solar-to-electrical energy conversion, photovoltaic cells offer a promising solu(cid:415)on.45– 47 A special subset of these, dye-sensi(cid:415)zed solar cells (DSSCs), are recognized for their cost- effec(cid:415)ve nature and their impressive photon-to-current conversion efficiencies under diverse ligh(cid:415)ng condi(cid:415)ons.20,48–52 A typical DSSC consists of a chromophore anchored to a semiconduc(cid:415)ng mesoporous metal oxide surface. When subjected to photoexcita(cid:415)on, an electron is injected into the semiconductor's conduc(cid:415)on band. Subsequently, the oxidized chromophore (or dye) is regenerated by a redox shu(cid:425)le present in the solu(cid:415)on. This injected electron, having traversed an external circuit, reaches the counter electrode. There, the oxidized redox shu(cid:425)le undergoes reduc(cid:415)on, comple(cid:415)ng the circuit.20,47,53 However, the tradi(cid:415)onally employed iodide/triiodide (I−/I3−) redox shu(cid:425)le system exhibits several drawbacks, including corrosiveness, lack of tunability, and compe(cid:415)(cid:415)ve absorp(cid:415)on in the visible spectrum by the triiodide ion.54–57 Notably, the I−/I3− redox couple's two-electron redox process o(cid:332)en leads to reduced theore(cid:415)cal maximum photovoltage output from DSSC devices, especially when compared to one-electron redox processes.57,58 Co(III/II)-based redox shu(cid:425)les, on the other hand, are accompanied by a significant inner-sphere reorganiza(cid:415)on energy during the electron transfer process.59,60 This energy hinders dye regenera(cid:415)on kine(cid:415)cs due to its limited availability for electron transfer. Copper-based redox shu(cid:425)les, however, have demonstrated parity with the classical I−/I3− and Co(III/II)-based redox shu(cid:425)les. Their success is o(cid:332)en a(cid:425)ributed to the 26 minimized inner-sphere electron transfer reorganiza(cid:415)on energies. The ability of copper complexes to adopt various geometries based on their metal oxida(cid:415)on state and ligand se(cid:427)ng is noteworthy.61–63 The majority of copper redox shu(cid:425)les employed in DSSCs thus far have relied on bidentate bipyridine or phenanthroline-based ligands.64 An approach which u(cid:415)lizes bulky groups adjacent to nitrogen donors to diminish structural shi(cid:332)s between Cu(II) and Cu(I) redox states. Recent reports also suggest that rigid tetradentate ligands can further op(cid:415)mize Cu(II/I) redox shu(cid:425)les by promo(cid:415)ng even lower inner-sphere reorganiza(cid:415)on energies, thus resul(cid:415)ng in heightened short-circuit current densi(cid:415)es and reduced VOC losses.65 Lewis bases, such as TBP and N-methylbenzimidazole (NMBI), are frequently incorporated addi(cid:415)ves in DSSC devices.35–37 These compounds have been shown to improve the overall cell performance. Such enhancements are o(cid:332)en associated with the ability of these bases to nega(cid:415)vely shi(cid:332) the (cid:415)tania conduc(cid:415)on band edge and suppress the interfacial charge recombina(cid:415)on rates through surface adsorp(cid:415)on.36,66–68 Notably, when most copper-based redox shu(cid:425)les are inves(cid:415)gated, TBP has been observed to coordinate with various Cu(II) species, poten(cid:415)ally displacing polydentate ligands.38,39,41,58,69–72 Such interac(cid:415)ons significantly influence the redox poten(cid:415)al by genera(cid:415)ng mul(cid:415)ple redox species, which in turn affects the maximum achievable photovoltage and electron transfer kine(cid:415)cs. The focal point of this research revolves around exploring the interac(cid:415)ons between a copper complex with a hexadentate ligand and TBP, with a primary goal of understanding if ligand subs(cid:415)tu(cid:415)on occurs upon oxida(cid:415)on. 3.2 Experimental Details All NMR spectra were taken on an Agilent DirectDrive2 500 MHz spectrometer at room temperature and referenced to residual solvent signals. All NMR spectra were evaluated using 27 the MestReNova so(cid:332)ware package features. Cyclic voltammograms were obtained using µAutolabIII poten(cid:415)ostat using BASi glassy carbon electrode, a pla(cid:415)num mesh counter electrode, and a fabricated 0.01 M AgNO3, 0.1 M TBAPF6 in acetonitrile Ag/AgNO3 reference electrode. All measurements were internally referenced to ferrocene/ferrocenium couple via addi(cid:415)on of ferrocene to solu(cid:415)on a(cid:332)er measurements or run in a parallel solu(cid:415)on of the same solvent/electrolyte. UV-Vis spectra were taken using a PerkinElmer Lambda 35 UV-Vis spectrometer using a 1 cm path length quartz cuve(cid:425)e at 480 nm/min. Elemental Analysis data were obtained via Midwest Microlab. For single-crystal X-ray diffrac(cid:415)on, single crystals were mounted on a nylon loop with paratone oil using a Bruker APEX-II CCD diffractometer. Crystals were maintained at T ¼ 173(2) K during data collec(cid:415)on. Using Olex2, the structures were solved with the ShelXS structure solu(cid:415)on program using the Direct Methods solu(cid:415)on method. Photoelectrochemical measurements were performed with a poten(cid:415)ostat (Autolab PGSTAT 128N) in combina(cid:415)on with a xenon arc lamp. An AM 1.5 solar filter was used to simulate sunlight at 100 mW cm-2, and the light intensity was calibrated with a cer(cid:415)fied reference cell system (Oriel Reference Solar Cell & Meter). A black mask with an open area of 0.07 cm-2 was applied on top of the cell ac(cid:415)ve area. A monochromator (Horiba Jobin Yvon MicroHR) a(cid:425)ached to the 450 W xenon arc light source was used for monochroma(cid:415)c light for IPCE measurements. The photon flux of the light incident on the samples was measured with a laser power meter (Nova II Ophir). IPCE measurements were made at 20 nm intervals between 400 and 700 nm at short circuit current. TEC 15 FTO was cut into 1.5 cm by 2 cm pieces which were sonicated in soapy DI water for 15 minutes, followed by manual scrubbing of the FTO with Kimwipes. The FTO pieces were then sonicated in DI water for 10 minutes, rinsed with acetone, and sonicated in isopropanol for 10 28 minutes. The FTO pieces were dried in room temperature air then immersed in an aqueous 40 mM TiCl4 solu(cid:415)on for 60 minutes at 70 °C. The water for the TiCl4 treatment was preheated to 70 °C prior to adding 2 M TiCl4 to the water. The 40 mM solu(cid:415)on was immediately poured onto the samples and placed in a 70 °C oven for the 60-minute deposi(cid:415)on. The FTO pieces were immediately rinsed with 18 MΩ water followed by isopropanol and were annealed by hea(cid:415)ng from room temperature to 500 °C, holding at 500 °C for 30 minutes. A 0.36 cm2 area was doctor bladed with commercial 30 nm TiO2 nanopar(cid:415)cle paste (DSL 30NRD). The transparent films were le(cid:332) to rest for 10 minutes and were then placed in a 125 °C oven for 30 minutes. The samples were annealed in an oven that was ramped to 325 °C for 5 minutes, 375 °C for 5 minutes, 450 °C for 5 minutes, and 500 °C for 15 minutes. The 30 nm nanopar(cid:415)cle film thickness was approximately 8 μm. A(cid:332)er cooling to room temperature, a second TiCl4 treatment was performed as described above. When the anodes had cooled to 80 °C, they were soaked in a dye solu(cid:415)on of 0.1 mM Y123 in 1 : 1 acetonitrile : tert-butyl alcohol for 18 hours. A(cid:332)er soaking, the anodes were rinsed with acetonitrile and were dried gently under a stream of nitrogen. The PEDOT counter electrodes were prepared by electropolymeriza(cid:415)on in a solu(cid:415)on of 0.01 M EDOT and 0.1 M LiClO4 in 0.1 M SDS in 18 MΩ water. A constant current of 8.3 mA for 250 seconds was applied to a 54 cm2 piece of TEC 8 FTO with predrilled holes using an equal size piece of FTO as the counter electrode. The PEDOT electrodes were then washed with DI water and acetonitrile before being dried under a gentle stream of nitrogen and cut into 1.5 cm by 1.0 cm pieces. The working and counter electrodes were sandwiched together with 25 μm surlyn films by placing them on a 140 °C hotplate and applying pressure. The cells were then filled in a nitrogen filled glove box with electrolyte through one of the two predrilled holes and were sealed with 25 μm 29 surlyn backed by a glass coverslip and applied heat to seal with a soldering iron. The electrolyte consisted of 0.10 M Cu(I), 0.05 M Cu(II), 0.1 M Li(Counter-ion), and 0.5 M 4-tert-butylpyridine in acetonitrile. All batches were made with at least 10 cells per batch. Contact to the TiO2 electrode was made by soldering a thin layer of indium wire onto the FTO. Prepara(cid:415)on of copper(I) 6,6(cid:3387)-bis(1,1-di(pyridine-2-yl)ethyl)-2,2(cid:3387)-bipyridine Bistriflimide ([Cu(bpyPY4)]TFSI: Copper(I) tetrakisacetonitrile bistriflimide (0.1647 g, 0.43 mmol) was dissolved in minimal dry acetonitrile. Meanwhile 6,6(cid:3387)-bis(1,1-di(pyridin-2-yl)ethyl)-2,2(cid:3387)-bipyridine (bpyPY4) (0.2545 g, 0.49 mmol) was dissolved in minimal dry dichloromethane. The bpyPY4 solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a dark red color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(I)bpyPY4 product was collected (0.2829 g, 76.1% yield). 1H-NMR (CDCl3 500 MHz) δ 2.25 (s, 6H), 7.219 (s, 4H), 7.76 (s, 4H), 8.10 (d, 2H), 8.17 (t, 2H), 8.35 (d, 2H) Elem. Anal. calc. for C36H28N7F7O4S2Cu: C, 50.03; H, 3.27; N, 11.34. Found: C, 50.05; H, 3.73; N, 11.08. Copper(II) bistriflimide (0.4215 g, 1.16 mmol) was dissolved in minimal dry acetonitrile. Meanwhile BPYPY4 (0.6942 g, 1.33 mmol) was dissolved in minimal dry dichloromethane. The BPYPY4 solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a pale blue color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(II)bpyPY4 product was collected (0.7912 g, 63.9% yield). Elem. Anal. calc. for C38H28N8F12O8S4Cu + ether: C, 41.40; H, 3.14; N, 9.20. Found: C, 41.48; H, 3.10; N, 9.13. 30 All the electrochemical measurements were performed by using μAutoLabII poten(cid:415)ostat or Autolab PGSTAT128N in a home-made electrochemical cell under N2 atmosphere. Copper complexes were dissolved in dry acetonitrile, the specific concentra(cid:415)ons will be specified in the result and discussion sec(cid:415)on. Also dissolved in the solu(cid:415)on was 0.1 M suppor(cid:415)ng electrolyte, which will be specified in the main text. For the cyclic voltammetry measurements, glassy carbon disk was used as the working electrode. The reference electrode was a home-made 0.1 M Ag/AgNO3 reference electrode. A salt bridge was used to prevent the leakage of the silver containing electrolyte in the reference electrode. The CV of ferrocene was measured in the same solvent before and a(cid:332)er measuring the sample. The redox poten(cid:415)al of ferrocene/ferrocenium was used to correct the poten(cid:415)al of the home- made reference electrode. The counter electrode for CV measurement was a pla(cid:415)num mesh. Scan rate was varied and will be specified in the main text. For the (cid:415)tra(cid:415)on measurement, different concentra(cid:415)ons of TBP were added into the solu(cid:415)on containing copper complexes and suppor(cid:415)ng electrolyte. The solu(cid:415)on was s(cid:415)rred and then stabilized before measuring. The absorp(cid:415)on spectra were measured with Lambda 35 (PerkinElmer) spectrometer. Solu(cid:415)ons containing copper complexes were prepared in the glovebox under an inert atmosphere with dry acetonitrile. Screw-cap quartz cuve(cid:425)e with 1 cm path length was used for all the absorp(cid:415)on measurements in solu(cid:415)on. Blank acetonitrile was used as the reference for the absorp(cid:415)on. For the (cid:415)tra(cid:415)on measurement, a stock solu(cid:415)on containing a constant concentra(cid:415)on of copper complex was prepared in the glove box. Each sample contains 4 mL of the stock solu(cid:415)on and a stoichiometric amount of TBP. The absorp(cid:415)on spectra of different samples were measured. 31 Proton NMR of copper complexes were measured using a J-Young NMR tube under an inert atmosphere with Agilent DDR2 500 MHz NMR spectrometer at room temperature. For the (cid:415)tra(cid:415)on measurement, 0.5 mL solu(cid:415)on containing a constant concentra(cid:415)on of copper complex was prepared in the glove box with deuterated acetonitrile. Dichloromethane was used as a non- coordina(cid:415)ng reference proton source to calibrate the proton peak area. A(cid:332)er running NMR of the pure copper complex, the tube was brought back into the glove box and another aliquot of TBP was added into the same sample. For each run, the parameters, number of scans, temperature, decay (cid:415)me, and scan range, used for the NMR measurement were kept the same. 3.3 Results and Discussion The synthesis of the ligand, bpyPY4, has been previously reported, wherein 1,1- bis(2- pyridyl)ethane underwent deprotona(cid:415)on with n-BuLi before the addi(cid:415)on of the electrophile, 6,6(cid:3387)- dibromo-2,2(cid:3387)-bipyridine. The metala(cid:415)on was done u(cid:415)lizing two different counter ions, triflate and bistriflimide, in an a(cid:425)empt to overcome solubility limita(cid:415)ons. The copper(I) triflate version exhibits a solubility of less than 0.1 M. In contrast, the solubility of the copper(I) bistriflimide version exceeds 1.25 M, leading to the use of bistriflimide versions for all studies on this copper complex system. As depicted in Scheme 3.1, the metala(cid:415)on is carried out using an equimolar ra(cid:415)o of the appropriate copper salt, either [Cu(acetonitrile)4](TFSI) or Cu(TFSI)2 where TFSI = Bistriflimide, with bpyPY4 in acetonitrile to afford [Cu(bpyPY4)](TFSI) or [Cu(bpyPY4)](TFSI)2, respec(cid:415)vely. The complexes were purified by slow diffusion of diethyl ether into concentrated acetonitrile solu(cid:415)ons, and their characteriza(cid:415)on was accomplished with 1H-NMR spectroscopy and elemental analysis. 32 Scheme 3.1: Syntheses of the [Cu(bpyPY4)]TFSI1/2 complexes and crystal structures of the ca(cid:415)ons of the Cu(I) species (le(cid:332)) and Cu(II) species (right). Depicted ellipsoids are at the 50% probability level. The anions and hydrogens were omi(cid:425)ed for clarity. Single crystal X-ray diffrac(cid:415)on ascertained the solid-state structures of the copper complexes, with details documented in Tables 3.1A and 3.2A. It was observed that the hexadentate ligand formed a five-coordinate geometry around the copper(II) center, binding five out of the six pyridine donors of the bpyPY4. Alterna(cid:415)vely, the copper(I) center revealed a four-coordinate geometry with two pyridine donors from bpyPY4 remaining unbound. Geometric indices, τ5 and τ4, are used for five-coordinate and four-coordinate complexes, respec(cid:415)vely, to determine how much distor(cid:415)on there is between ideal geometries.73,74 The τ value in each case will have a range from 0 to 1. For five-coordinate complexes, a τ5 value of 0 represents an ideal square pyramidal geometry, and a value of 1 represents an ideal trigonal bipyramidal geometry. For the five- coordinate copper(II) complex, the τ5 value was 0.61, indica(cid:415)ve of a distorted trigonal bipyramidal geometry. Within this configura(cid:415)on, the bipyridine fragment of the bpyPY4 ligand 33 posi(cid:415)ons itself both axially and equatorially around the copper(II) ion. Notably, the copper(II) ion exhibited a displacement distance of 0.225 Å from the plane defined by the equatorial donors, poin(cid:415)ng towards the axial pyridine donor. As a result, the angles involving the axial pyridine nitrogen donor, the copper(II) ion, and the equatorial donors slightly exceed 90°. Moreover, the bipyridine fragment showcases a near-planar orienta(cid:415)on, with its pyridine rings forming a torsion angle of 5.0(4)°. In the case of the copper(I) complex, a four-coordinate complex, an ideal square planar geometry is represented by a τ4 value of 0, and an ideal tetrahedral geometry is represented by a τ4 value of 1.74 The copper(I) counterpart showcased a τ4 value of 0.63, deno(cid:415)ng a highly distorted tetrahedral geometry. In terms of coordina(cid:415)on, the bpyPY4 ligand coordinates the copper(I) ion in a manner where the bipyridine unit occupies two of the coordina(cid:415)on sites, and the remaining two sites have a pyridine from each of the 1,1-bis(2-pyridyl)ethane fragments bound. The average bond distance between the copper-bipyridine nitrogen stands at 2.082(3) Å, aligning with data from other tetrahedral copper(I) complexes with bipyridine ligands.75,76 Yet, the copper-pyridine nitrogen bond is slightly shorter, averaging 2.015(2) Å. The bipyridyl segment, in this instance, demonstrates a pronounced distor(cid:415)on, marked by a torsion angle of -18.0(4)° between its pyridine rings. The 1H-NMR spectrum of [Cu(bpyPY4)]TFSI, recorded in deuterated acetonitrile at room temperature, integrates to 20 protons. Two broad peaks of equal intensity at δ = 7.17 and 7.73 ppm, integra(cid:415)ng only to four protons each, combine for eight protons in total. Sharp resonances at δ = 810, 8.17, and 8.35 ppm, comprising two doublets and a triplet further downfield in the aroma(cid:415)c spectrum, account for six protons, two protons per peak, and align with the bipyridine 34 unit. A sharp singlet at 2.25 ppm accounts for the remaining six protons, represen(cid:415)ng the two methyl peaks. However, the bpyPY4 framework has 28 protons. Intriguingly, while the bpyPY4 framework boasts 28 aroma(cid:415)c protons, the spectrum under-represents this number by eight protons. Insights from the solid-state structure of the copper(I) complex, depicted in Scheme 3.1, reveal two non-coordina(cid:415)ng pyridines, which could account for the eight proton discrepancy. The observable broad peaks at 7.17 ppm and 7.73 ppm suggest a rapid exchange on the NMR (cid:415)mescale between non-coordinated and coordinated pyridine donors. Figure 3.1: 1H-NMR spectrum (500 MHz, CD3CN) of [Cu(bpyPY4)](TFSI). To delve deeper into this phenomenon, variable-temperature 1H-NMR spectroscopy was executed, with spectra obtained from −40 to 25 °C consolidated in Figure 3.2. This allows for an interroga(cid:415)on of the solu(cid:415)on's dynamic coordina(cid:415)on behavior. As the temperature decreased, 35 new broad peaks emerged by 0°C, yielding eight pronounced peaks by −30°C. Integra(cid:415)ng to 16 protons, these peaks originate from the four terminal pyridines, illustra(cid:415)ng two dis(cid:415)nc(cid:415)ve chemical environments. The presump(cid:415)on is that two of these fluctua(cid:415)ng pyridines, one from each dipyridylethane “arm”, remain uncoordinated, each possessing four dis(cid:415)nct aroma(cid:415)c protons. In contrast, the other two are coordinated, accoun(cid:415)ng for the remaining four peaks. Notably, chemical shi(cid:332)s for non-coordinated pyridines diverge slightly from the free ligand, but such a downfield shi(cid:332) has been documented in prior studies for dissociated heterocycles within a coordinated ligand matrix.77,78 Figure 3.2: Variable temperature 1H-NMR spectra of [Cu(bpyPY4)]TFSI in CD3CN. In anhydrous acetonitrile, the UV-visible spectra of the copper complexes reveal dis(cid:415)nct absorp(cid:415)on characteris(cid:415)cs. The [Cu(bpyPY4)]TFSI complex shows absorp(cid:415)on peaks of moderate intensity spanning from 249 to 324 nm, with molar absorp(cid:415)vi(cid:415)es ranging from 51,500 to 33,400 M−1cm−1. The copper(I) complex also displays lower intensity shoulders at 329, 391, and 450 nm, characterized by molar absorp(cid:415)vi(cid:415)es of 10000, 6600, and 4100 M−1cm−1, respec(cid:415)vely. A subtle shoulder peak at 538 nm is observed, signifying a metal-to-ligand charge transfer with an ex(cid:415)nc(cid:415)on coefficient of 2500 M−1cm−1. The copper(II) complex demonstrates d−d transi(cid:415)on bands at 616 and 845 nm, each with ex(cid:415)nc(cid:415)on coefficients of 270 and 194 M−1cm−1, respec(cid:415)vely. 36 These characteris(cid:415)cs are emblema(cid:415)c of a structure predisposed to a trigonal bipyramidal geometry in solu(cid:415)on. Figure 3.3 UV−visible spectra of the[Cu(bpyPY4)]TFSI complex (yellow) and the [Cu(bpyPY4)]TFSI2 complex (blue) in anhydrous CH3CN. The cyclic voltammogram of the copper(II) complex revealed a copper(II/I) redox couple at −0.454 V versus the ferrocenium/ferrocene couple as shown in Figure 3.4, with a peak spli(cid:427)ng value of 93 mV, indica(cid:415)ng that it is quasi-reversible. This redox poten(cid:415)al is found to be surprisingly nega(cid:415)ve rela(cid:415)ve to most previously reported copper(II/I) redox couples. 79–84 37 A µ / t n e r r u C 15.00 10.00 5.00 0.00 -5.00 -10.00 -15.00 -1.2 -1 -0.8 -0.6 -0.4 Potential vs Ferrocene / V -0.2 0 0.2 0.4 Figure 3.4: A cyclic voltammogram of 2.00 mM [Cu(bpyPY4)](TFSI)2 in anhydrous CH3CN with 0.1 M TBAPF6 as a suppor(cid:415)ng electrolyte, measured at a scan rate of 0.1 V/s on a glassy carbon working electrode. The interac(cid:415)on of TBP with the [Cu(bpyPY4)]2+/+ redox couple was probed via cyclic voltammetry at varying concentra(cid:415)ons of the Lewis base, with TBP present in an excess of at least 10 equivalents to align with device condi(cid:415)ons delineated later. The addi(cid:415)on of TBP to the copper complex shows no significant influence on either the cathodic or anodic wave of the copper(II/I) redox couple, indica(cid:415)ng there is no reac(cid:415)on between the [Cu(bpyPY4)]2+ complex and TBP. This observa(cid:415)on was further substan(cid:415)ated by UV-Vis spectroscopic analysis, showing that, despite increasing TBP concentra(cid:415)ons, the UV-visible absorp(cid:415)on spectra remained consistent a(cid:332)er being corrected for dilu(cid:415)on. This reinforces the finding that TBP remains unbound to the metal center of the redox shu(cid:425)le. If there was subs(cid:415)tu(cid:415)on occurring the peak at 845 nm would show a rapid decrease in intensity due to a lack of contribu(cid:415)on from the [Cu(TBP)4]2+ complex. Further insight was sought using 1H-NMR spectroscopy. Titra(cid:415)on of the [Cu(bpyPY4)]2+ sample with TBP unveiled 38 no signals corresponding to free bpyPY4, sugges(cid:415)ng an absence of ligand subs(cid:415)tu(cid:415)on. This behavior is a(cid:425)ributed to the transient, non-coordinated pyridines within the hexadentate ligand, which likely occlude any open coordina(cid:415)on sites on the metal, precluding the detrimental binding of external Lewis bases. Figure 3.5: A) Cyclic Voltammogram of 2 mM [Cu(bpyPY4)](TFSI)2 with 0.1 M TBAPF6 in anhydrous CH3CN with increasing concentra(cid:415)on of TBP at a glassy carbon working electrode; scan rate = 100 mVs-1. B) UV-Vis absorp(cid:415)on spectra of 2 mM [Cu(bpyPY4)](TFSI)2 in anhydrous CH3CN with increasing equivalents of TBP. A spectra of 2 mM [Cu(TBP)4]TFSI is shown as a control. C) 1H-NMR spectra of [Cu(bpyPY4)](TFSI)2 in CH3CN with increasing equivalents of TBP. Unbound bpyPY4 and TBP are shown as controls. Stopped-flow spectroscopy was u(cid:415)lized to measure the cross-exchange electron transfer rate constant, k12, between [Cu(bpyPY4)]2+ and decamethylferrocene (Fe(Cp*)2). 39 Scheme 2.2: Reac(cid:415)on followed in stopped-flow experiments. Decamethylferrocene was chosen for the cross-exchange to op(cid:415)mize the driving force and because it has a well-known self-exchange rate constant for electron transfer. Due to the large poten(cid:415)al difference between [Cu(bpyPY4)]2+/+ and [Fe(Cp*)2]+/0, it was assumed that the reac(cid:415)on went to comple(cid:415)on, with no significant back reac(cid:415)on. Figure 3.6A shows a fit of the absorbance at 450 nm vs (cid:415)me plot, which represents the growth of the [Cu(bpyPY4)]+ species in solu(cid:415)on due to the reduc(cid:415)on of [Cu(bpyPY4)]2+ by Fe(Cp*)2, using the following equa(cid:415)on: 𝐴 = 𝐴(cid:2998) + (𝐴(cid:2868) − 𝐴(cid:2998))𝑒(cid:2879)(cid:3038)(cid:3290)(cid:3277)(cid:3294)(cid:3047) (3.1) For all reac(cid:415)ons, Fe(Cp*)2 was held in excess to maintain pseudo-first order condi(cid:415)ons, allowing kobs to be represented by: 𝑘𝑜𝑏𝑠 = 𝑘12[𝐹𝑒(𝐶𝑝∗ )2 ] (3.2) Figure 3.6: A) An absorbance vs (cid:415)me plot at 450 nm, showing the increase of [Cu(bpyPY4)]TFSI species (black dot) and fi(cid:427)ng (red line) for the reduc(cid:415)on of [Cu(bpyPY4)]TFSI2 by decamethylferrocene (Fe(Cp*)2). B) The pseudo-first order rate constant, kobs, vs the concentra(cid:415)on of [Fe(Cp* )2] for the reac(cid:415)ons of [Cu(bpyPY4)]TFSI2 and [Fe(Cp* )2] in CH3CN containing 0.1 M LiTFSI. 40 A linear fit of the previously fi(cid:425)ed kobs values vs concentra(cid:415)on of Fe(Cp*)2 provided the value of the cross-exchange rate constant, k12, from the slope of the line (Figure 3.6B). The ini(cid:415)al reac(cid:415)on mixtures and the pseudo-first order observed rate constants can be found in Table 3.1. Table 3.3: The ini(cid:415)al reac(cid:415)on mixture and the observed rate constant, kobs, for the cross exchange between [Cu(bpyPY4)]TFSI2 and [Fe(Cp* )2]. Using the experimentally determined cross-exchange rate constant from Equa(cid:415)on 3.2, k12, and the previously determined self-exchange constant for [Fe(Cp*)2]+/0 , k11, the Marcus cross-rela(cid:415)on 𝑘12 = (𝑘11𝑘22𝐾12𝑓12)1/2𝑊12 (3.3) was used to calculate the self-exchange rate constant for [Cu(bpyPY4)]2+/+ , k22, where K12 is the equilibrium constant, f12 is a nonlinear correc(cid:415)on term, and W12 is an electrosta(cid:415)c work term for bringing the reactants into contact. The nonlinear correc(cid:415)on term and the work term were calculated (see below for details on the calcula(cid:415)on) and determined to be 2.60 and 0.99, respec(cid:415)vely (Table 3.3). However, it should be noted that there will be some error in the calcula(cid:415)on of the work term since the Debye-Huckel model is not expected to have accurate results when the solu(cid:415)on has a high ionic strength, as was the case for this experiment with 0.1 M suppor(cid:415)ng electrolyte. The equilibrium constant, K12, was determined using Equa(cid:415)on 3.4: 𝑛𝐹𝛥𝐸 = 𝑅𝑇𝑙𝑛𝐾12 (3.4) where n is the number of electrons transferred, F is Faraday’s constant, ΔE is the difference in formal poten(cid:415)al between the oxidant and reductant in solu(cid:415)on, R is the ideal gas constant, and 41 T is the temperature in Kelvin. The redox poten(cid:415)al for [Cu(bpyPY4)]2+/+ was found to be -0.43 V vs ferrocene and the redox poten(cid:415)al of [Fe(Cp*)2]+/0 has been previously determined to be -0.52 V2 vs ferrocene85, using cyclic voltammetry which gave a calculated K12 of 18.4. The self-exchange rate constant for [Fe(Cp*)2]+/0 was previously iden(cid:415)fied to be 3.8 × 107 M-1s-1 using NMR techniques.86 Using these values, the self-exchange rate for the [Cu(bpyPY4)]2+/+ couple was determined to be 8.78 M-1s-1. The measurement was run mul(cid:415)ple (cid:415)mes to confirm reproducibility and all samples were within error (see Table 3.2). Table 3.4: The self-exchange rate of [Cu(bpyPY4)]TFSI1/2 complex for each trial. The work required to move the reactant complexes to a distance, r, for the electron transfer reac(cid:415)on was calculated using the Equa(cid:415)on 3.5. 𝑊(cid:2869)(cid:2870) = exp [ (cid:2879)(cid:3050)(cid:3117)(cid:3118)(cid:2878)(cid:3050)(cid:3118)(cid:3117)(cid:2879)(cid:3050)(cid:3117)(cid:3117)(cid:2879)(cid:3050)(cid:3118)(cid:3118) (cid:2870)(cid:3019) ] (3.5) 𝑤(cid:3036)(cid:3037)(𝑟) = (cid:3053)(cid:3284)(cid:3053)(cid:3285)(cid:3044)(cid:3118)(cid:3015)(cid:3250) (cid:2872)(cid:3095)(cid:3084)(cid:3116)(cid:3084)(cid:3045)((cid:2869)(cid:2878)(cid:3081)(cid:3045)) (3.6) Equa(cid:415)on 3.6 is used to determine the work associated with the forward and reverse cross- exchange reac(cid:415)on, w12 and w21 respec(cid:415)vely, and the self-exchange reac(cid:415)ons, w11 and w22. In this equa(cid:415)on zi and zj are the charges of the interac(cid:415)ng complexes, q is the charge of an electron, NA is Avogadro’s constant, ε0 is the permi(cid:427)vity of free space, ε is the sta(cid:415)c dielectric of the medium, 𝛽 = ( (cid:2870)(cid:3044)(cid:3118)(cid:3015)(cid:3250)(cid:3010) (cid:2869)(cid:2868)(cid:2868)(cid:2868) (cid:3116)(cid:3084)(cid:3038)(cid:3251)(cid:3021) )(cid:2869)/(cid:2870), I is the ionic strength of the solu(cid:415)on, and kB is the Boltzmann’s constant. The calcula(cid:415)on has the following assump(cid:415)ons: the work is assumed to be primarily Coulombic, the 42 distance, r, is assumed to be the center-to-center distance between the complexes, and the reactants are assumed to be spherical. The non-linear correc(cid:415)on term, f12, was calculated using the equa(cid:415)on 𝑙𝑛𝑓(cid:2869)(cid:2870) = (cid:2869) (cid:2872) (cid:4672)(cid:3039)(cid:3041) (cid:3117)(cid:3118)(cid:2878) (cid:3286)(cid:3117)(cid:3117)(cid:3286)(cid:3118)(cid:3118) (cid:3275)(cid:3118) (cid:2922)(cid:2924)(cid:4672) (cid:3118) (cid:4673) (cid:3298)(cid:3117)(cid:3118)(cid:3127)(cid:3298)(cid:3118)(cid:3117) (cid:3267)(cid:3269) (cid:3298)(cid:3117)(cid:3117)(cid:3126)(cid:3298)(cid:3118)(cid:3118) (cid:3267)(cid:3269) (cid:4673)(cid:2878) (3.7) where K12 is the equilibrium constant, w12 and w21 are the work associated with the forward and reverse cross-exchange reac(cid:415)on respec(cid:415)vely, R is the ideal gas constant, T is temperature in Kelvin, and Z is the frequency factor, which is assumed to be 1013 M-1s-1 due to the larger inner- sphere contribu(cid:415)ons to the total reorganiza(cid:415)on energy. The total reorganiza(cid:415)on energy of the self-exchange reac(cid:415)on was calculated following the equa(cid:415)on 𝑘(cid:2869)(cid:2869) = 𝐾(cid:2869)(cid:2870)𝑍𝛤𝑒(cid:2879)((cid:3090)(cid:3117)(cid:3117) (cid:2872)(cid:3038)(cid:3251)(cid:3021)) ⁄ (3.8) where k11 is the previously calculated self-exchange rate, K12 is the equilibrium constant, Z is the frequency factor, same as in Equa(cid:415)on 3.7, Γ is a correc(cid:415)on for nuclear tunneling, which is assumed to be ~1, λ11 is the total reorganiza(cid:415)on energy of the self-exchange reac(cid:415)on, kB is the Boltzmann constant, and T is the temperature of solu(cid:415)on. The resul(cid:415)ng λ11 was found to be 3.15 eV. The outer-sphere reorganiza(cid:415)on energy for two spherical reactants can be determined using 𝜆(cid:2869)(cid:2869),(cid:3042) = ((cid:3057)(cid:3053)(cid:3044))(cid:3118) (cid:2872)(cid:3095)(cid:3084)(cid:3116) (cid:4672) (cid:2869) (cid:2870)(cid:3028)(cid:3117) + (cid:2869) (cid:2870)(cid:3028)(cid:3118) − (cid:2869) (cid:3045)(cid:3117)(cid:3117) (cid:4673) (cid:3436) (cid:2869) (cid:3005)(cid:3290)(cid:3291),(cid:3294)(cid:3290)(cid:3287) − (cid:2869) (cid:3005)(cid:3294),(cid:3294)(cid:3290)(cid:3287) (cid:3440) (3.9) where λ11,o is the outer-sphere reorganiza(cid:415)on energy, Δz is the change in charge of the complex, q is the charge of an electron, ε0 is the permi(cid:427)vity of free space, a1 and a2 are the atomic radii of the reactants, r11 is the center−center distance between the reactants, Dop,sol is the op(cid:415)cal dielectric constant of the medium, which is equal to the square of the refrac(cid:415)ve index of the 43 medium, and Ds,sol is the sta(cid:415)c dielectric constant of the medium. The outer-sphere reorganiza(cid:415)on energy for the self-exchange reac(cid:415)on was calculated to be 0.99 eV. The inner-sphere reorganiza(cid:415)on energy was determined using 𝜆11,𝑖 = 𝜆11 − 𝜆11,𝑂 yielding a result of 2.16 eV for the inner-sphere reorganiza(cid:415)on energy of the self-exchange reac(cid:415)on. Table 3.5: Summary of the kine(cid:415)c data used to calculate the self-exchange rate constant of the [Cu(bpyPY4)]TFSI1/2 complex by monitoring the reac(cid:415)on of [Cu(bpyPY4)]TFSI2 and [Fe(Cp* )2] in CH3CN containing 0.1 M LiTFSI at 25°C. DSSC devices were fabricated with [Cu(bpyPY4)]TFSI1/2 as the redox mediator, where poly(3,4- ethylenedioxythiophene) (PEDOT) was u(cid:415)lized at the counter electrode and the commercial dye Y123 was used as the light-harvester. The electrolyte consisted of 0.10 M [Cu(bpyPY4)]TFSI, 0.05 M [Cu(bpyPY4)]TFSI2, 0.1 M LiTFSI, and 0.5 M 4-tert-butylpyridine, 10 equivalents vs the Cu(II) in the electrolyte, in acetonitrile. The constructed devices gave respectable performance with a JSC of 4.98(±0.01) mAcm-2 and an overall power conversion efficiency of 1.23% for the devices with TBP and 4.73(±0.36) mAcm-2 and an overall power conversion efficiency of 0.21% for the devices without TBP, as shown in Table 3.6. 44 2 - m c A m / t n e r r u C 6 5 4 3 2 1 0 -0.5 -0.4 -0.3 -0.2 Voltage / V -0.1 0 Figure 3.7: J−V curves for the [Cu(bpyPY4)]TFSI1/2 based DSSC devices. The devices represented by the black curve contains TBP while the gray curve is without TBP. Table 3.6: Summary of device performance, parameters represent the average of three devices. η (%) Jsc (mA*cm-2) Voc (V) FF Cu(PY6) 0.21(±0.07) 4.73(±0.36) 0.14(±0.01) 0.32(±0.05) Cu(PY6) with TBP 1.23(±0.08) 4.98(±0.01) 0.44(±0.01) 0.56(±0.02) Both sets of devices were affected by mass transport issues which limited their current. Looking at the incident photon-to-current conversion efficiency (IPCE) spectra, which measures the current at various wavelengths to determine how much of the photons absorbed at each wavelength are converted to current. The IPCE in Figure 3.8 shows that as TBP is added there is a slight increase in the photon-to-current conversion between 450 and 650 nm. This increase could be caused by decreased recombina(cid:415)on to the redox shu(cid:425)le as TBP is being added to the solu(cid:415)on. The IPCE spectra can also be integrated to determine what the theore(cid:415)cal current of the device, resul(cid:415)ng with an integrated current of 7.87 mAcm-2 for the devices with TBP and 7.77 mAcm-2 for 45 the devices without TBP, we can see that roughly 3 mAcm-2 were lost due to the mass transport limita(cid:415)ons. % / E C P I 60 50 40 30 20 10 0 9.00 8.00 7.00 6.00 5.00 4.00 3.00 2.00 1.00 0.00 2 - m c A m / t n e r r u C 400 450 500 550 600 650 700 Wavelength / nm Figure 3.8: IPCE curves for the [Cu(bpyPY4)]TFSI1/2 based DSSC devices. The devices represented by the black points and curve contains TBP while the gray points and curve is without TBP. For the IPCE, the points represent IPCE responses while the solid lines represent the integrated photocurrent. The effect of TBP on the fermi level of the devices was inves(cid:415)gated. By knowing the solu(cid:415)on poten(cid:415)al and the VOC of the devices the fermi level could be calculated, which can be determined by adjus(cid:415)ng the open circuit voltage (VOC) by the solu(cid:415)on poten(cid:415)al. For both devices with and without TBP the solu(cid:415)on poten(cid:415)al was roughly the same, at -0.443 V vs ferrocene for the devices without TBP and -0.466 V vs ferrocene for the devices with TBP. However, there was a stark difference in the VOC between both sets of devices, devices without TBP having a VOC of only 0.14 V vs solu(cid:415)on poten(cid:415)al where devices with TBP had a VOC of 0.44 V vs solu(cid:415)on poten(cid:415)al. This indicates that the addi(cid:415)on of TBP caused the fermi level to undergo a nega(cid:415)ve shi(cid:332) of roughly 300 mV. Although the [Cu(bpyPY4)]TFSI1/2 complex does not undergo ligand subs(cid:415)tu(cid:415)on the 46 addi(cid:415)on of TBP to the electrolyte s(cid:415)ll dras(cid:415)cally increases the electron life(cid:415)me, by almost 1000 fold. This indicates that the addi(cid:415)on of TBP is s(cid:415)ll able to reduce recombina(cid:415)on in the system, although not as significantly as the [Cu(dbmed)]OTf1/2 system. This shi(cid:332) in poten(cid:415)al could be a(cid:425)ributed to the addi(cid:415)on of TBP in solu(cid:415)on, which can shi(cid:332) the poten(cid:415)al of the conduc(cid:415)on band of the (cid:415)tania to more nega(cid:415)ve poten(cid:415)als and reducing recombina(cid:415)on by adsorbing to the (cid:415)tania.33,34 10 1 0.1 ) s / 1 ( e m i t e f i L 0.01 0 [Cu(dbmed)] without TBP [Cu(bpyPY4)] without TBP [Cu(dbmed)] with TBP [Cu(bpyPY4)] with TBP 0.2 0.4 0.6 0.8 1 1.2 Voltage (V) Figure 3.9: A plot of the electron life(cid:415)me vs poten(cid:415)al plot for DSSC devices using [Cu(dbmed)]OTf1/2 (green), [Cu(bpyPY4)]TFSI1/2 (red), with (pale color) and without (dark color) the addi(cid:415)on of TBP. 3.4 Conclusions This study has demonstrated the complex rela(cid:415)onship between ligand design and metal coordina(cid:415)on in copper complexes using the bpyPY4 scaffold. The crystallographic analysis revealed that the bpyPY4 ligand can adapt to both five-coordinate and four-coordinate structures 47 around copper(II) and copper(I) centers. These structures showed significant devia(cid:415)ons from ideal geometries, indica(cid:415)ng the influence of the bpyPY4 scaffold on how the ligand and metal interact. The behavior of these copper-ligand interac(cid:415)ons was further explored using 1H-NMR studies, which showed a rapid interchange between pyridine donors that are coordinated and those that are not. The results from variable-temperature NMR spectroscopy highlighted the unstable nature of the copper-pyridine bonds, providing a detailed view of how these bonds behave in solu(cid:415)on. The UV-visible spectroscopic analysis helped improve understanding on the electronic transi(cid:415)ons in these complexes, revealing specific absorp(cid:415)on pa(cid:425)erns that correspond to their structural preferences in solu(cid:415)on. The consistent behavior of the [Cu(bpyPY4)]2+/+ redox couple, even with added TBP, suggests a strong metal-ligand bond, which is promising for applica(cid:415)ons where interac(cid:415)ons with Lewis bases are expected. The high theore(cid:415)cal current and the nega(cid:415)ve redox poten(cid:415)al make this redox shu(cid:425)le a candidate for u(cid:415)liza(cid:415)on with narrow band gap dyes for high photocurrent DSSC devices. Future studies will focus on u(cid:415)lizing this redox complex with near IR absorbing dyes and developing new ligand designs to vary the energe(cid:415)cs and stability of the copper species. Overall, this comprehensive study not only deepens our understanding of copper complexes with hexadentate ligands but also lays the groundwork for designing new ligands that can control and alter the shapes and reac(cid:415)vi(cid:415)es of metal centers. The poten(cid:415)al uses of these systems in areas like catalysis and electronic devices are significant, highligh(cid:415)ng the value of detailed studies like this one. 48 Chapter 4 – Inves(cid:415)ga(cid:415)on of [Cu(PY5)]2+/+ 4.1 Introduc(cid:415)on Following Grätzel's seminal 1993 report on dye-sensi(cid:415)zed solar cells (DSSCs) achieving 10% efficiency, advancements were hindered by the reliance on the I3-/I- redox shu(cid:425)le.15 The high overpoten(cid:415)al required for efficient dye regenera(cid:415)on was a major limita(cid:415)on. Recent progress with outer sphere redox shu(cid:425)les offers a solu(cid:415)on to mi(cid:415)gate this limita(cid:415)on and enhance DSSC performance.23 Conversely, Co(III/II) based redox shu(cid:425)les introduce an inner-sphere reorganiza(cid:415)on energy challenge due to electron transfer transi(cid:415)ons between d7 Co(II) and d6 Co(III) states. This affects dye regenera(cid:415)on kine(cid:415)cs by limi(cid:415)ng the available electron transfer driving force.23 The robust performance of copper-based redox shu(cid:425)les in DSSCs has posi(cid:415)oned them as a compelling avenue for overcoming conven(cid:415)onal limita(cid:415)ons. Their success stems from their remarkable ability to a(cid:425)ain efficiencies that surpass previous benchmarks. With an excep(cid:415)onal 15.2% efficiency under standard solar irradiance and an astonishing 34.5% efficiency under indoor fluorescent ligh(cid:415)ng at 1000 lux intensity, copper-based redox shu(cid:425)les have firmly established their creden(cid:415)als.20,87 This achievement can be a(cid:425)ributed to their intrinsic advantage of lower inner-sphere electron transfer reorganiza(cid:415)on energies compared to tradi(cid:415)onal op(cid:415)ons. Furthermore, their versa(cid:415)le coordina(cid:415)on geometries, witnessed in copper(II) d9 complexes which frequently adopt six-coordinate, octahedral or tetragonal, five-coordinate, square pyramidal or trigonal bipyramidal, or four-coordinate, tetrahedral or square planar geometries, and copper(I) d10 complexes which predominantly adopt four-coordinate, tetrahedral or square planar geometries, underscore their adaptability.60–63 Employing ligands based on bidentate bipyridine 49 or phenanthroline mo(cid:415)fs, augmented with strategically posi(cid:415)oned bulky groups, these shu(cid:425)les can reduce the reorganiza(cid:415)on energy needed for the transi(cid:415)on between Cu(II) and Cu(I) redox states, facilita(cid:415)ng more efficient electron transfer processes within the DSSC system.64 Recent observa(cid:415)ons have illuminated a cri(cid:415)cal facet of copper-based redox shu(cid:425)le performance within DSSCs. Notably, prevalent Lewis base addi(cid:415)ves, including 4-tert-butylpyridine (TBP), have been revealed to coordinate with Cu(II) species.38 This coordina(cid:415)on has poten(cid:415)ally adverse impacts on copper redox mediator electrochemistry and, consequently, overall device efficacy. Studies into this phenomenon elucidate that TBP-induced coordina(cid:415)on or ligand subs(cid:415)tu(cid:415)on— o(cid:332)en involving bipyridyl and phenanthroline ligands—leads to intricate electrochemical changes within the DSSC devices. When the ligand subs(cid:415)tu(cid:415)on occurs and [Cu(TBP)4]2+ is present in the devices it is accompanied by a substan(cid:415)al nega(cid:415)ve shi(cid:332) in formal redox poten(cid:415)als, by hundreds of millivolts.38 This phenomenon limits the a(cid:425)ainable photovoltage, exer(cid:415)ng a significant constraint on device performance. Recent strides in copper-based redox shu(cid:425)le design have been underscored by the efforts of Sun and colleagues, who introduced copper based redox shu(cid:425)les incorpora(cid:415)ng pentadentate ligands within DSSC devices.41 The pentadentate Cu(II) complex was shown to have a resistance to subs(cid:415)tu(cid:415)on, even when exposed to commonly employed Lewis base addi(cid:415)ves such as 4-tert- butylpyridine (TBP). This resistance to ligand subs(cid:415)tu(cid:415)on is a(cid:425)ributed to two factors. First, the increased den(cid:415)city of the ligand translates to a heightened overall stability of the metal complex, owing to the chela(cid:415)ng effect. Second, they specifically designed the coordina(cid:415)on sphere's steric constraints to shield the copper complexes—par(cid:415)cularly their oxidized forms—from TBP 50 coordina(cid:415)on. Building upon this concept, a copper-based redox shu(cid:425)le harnessing the potency of a pentadentate ligand, 2,6-bis[1,1-bis(2-pyridyl)ethyl]pyridine (PY5), is presented. 4.2 Experimental Details All NMR spectra were taken on an Agilent DirectDrive2 500 MHz spectrometer at room temperature and referenced to residual solvent signals. All NMR spectra were evaluated using the MestReNova so(cid:332)ware package features. Cyclic voltammograms were obtained using µAutolabIII poten(cid:415)ostat using BASi glassy carbon electrode, a pla(cid:415)num mesh counter electrode, and a fabricated 0.01 M AgNO3, 0.1 M TBAPF6 in acetonitrile Ag/AgNO3 reference electrode. All measurements were internally referenced to ferrocene/ferrocenium couple via addi(cid:415)on of ferrocene to solu(cid:415)on a(cid:332)er measurements or run in a parallel solu(cid:415)on of the same solvent/electrolyte. UV-Vis spectra were taken using a PerkinElmer Lambda 35 UV-Vis spectrometer using a 1 cm path length quartz cuve(cid:425)e at 480 nm/min. Elemental Analysis data were obtained via Midwest Microlab. Con(cid:415)nuous-wave (Cw) EPR measurements were performed using a Bruker E-680X spectrometer opera(cid:415)ng at X-band frequencies and equipped with an SHQ- E cavity. EPR samples comprising 0.01 M Cu(II) complexes in acetonitrile or dichloromethane. The liquid solu(cid:415)ons were studied in a quartz flat cell. An Oxford ESR-900 cryostat and an ITC-503 temperature controller were used for measurements at 298 K. Cw-EPR simula(cid:415)ons were performed with EasySpin 5.2.28, running in MATLAB 2020a. We simulated the solu(cid:415)on spectra using the “chili” module from EasySpin. Our simula(cid:415)ons varied the coefficients for copper hyperfine coupling, isotropic diffusion correla(cid:415)on (cid:415)me, and intrinsic linewidth. The results are shown as dashed curves offset from the spectra (solid lines).For single-crystal X-ray diffrac(cid:415)on, single crystals were mounted on a nylon loop with paratone oil using a Bruker APEX-II CCD 51 diffractometer. Crystals were maintained at T ¼ 173(2) K during data collec(cid:415)on. Using Olex2, the structures were solved with the ShelXS structure solu(cid:415)on program using the Direct Methods solu(cid:415)on method. Photoelectrochemical measurements were performed with a poten(cid:415)ostat (Autolab PGSTAT 128N) in combina(cid:415)on with a xenon arc lamp. An AM 1.5 solar filter was used to simulate sunlight at 100 mW cm-2, and the light intensity was calibrated with a cer(cid:415)fied reference cell system (Oriel Reference Solar Cell & Meter). A black mask with an open area of 0.07 cm-2 was applied on top of the cell ac(cid:415)ve area. A monochromator (Horiba Jobin Yvon MicroHR) a(cid:425)ached to the 450 W xenon arc light source was used for monochroma(cid:415)c light for IPCE measurements. The photon flux of the light incident on the samples was measured with a laser power meter (Nova II Ophir). IPCE measurements were made at 20 nm intervals between 400 and 700 nm at short circuit current. TEC 15 FTO was cut into 1.5 cm by 2 cm pieces which were sonicated in soapy DI water for 15 minutes, followed by manual scrubbing of the FTO with Kimwipes. The FTO pieces were then sonicated in DI water for 10 minutes, rinsed with acetone, and sonicated in isopropanol for 10 minutes. The FTO pieces were dried in room temperature air then immersed in an aqueous 40 mM TiCl4 solu(cid:415)on for 60 minutes at 70 °C. The water the for the TiCl4 treatment was preheated to 70 °C prior to adding 2 M TiCl4 to the water. The 40 mM solu(cid:415)on was immediately poured onto the samples and placed in a 70 °C oven for the 60-minute deposi(cid:415)on. The FTO pieces were immediately rinsed with 18 MΩ water followed by isopropanol and were annealed by hea(cid:415)ng from room temperature to 500 °C, holding at 500 °C for 30 minutes. A 0.36 cm2 area was doctor bladed with commercial 30 nm TiO2 nanopar(cid:415)cle paste (DSL 30NRD). The transparent films were le(cid:332) to rest for 10 minutes and were then placed in a 125 °C oven for 30 minutes. The samples 52 were annealed in an oven that was ramped to 325 °C for 5 minutes, 375 °C for 5 minutes, 450 °C for 5 minutes, and 500 °C for 15 minutes. The 30 nm nanopar(cid:415)cle film thickness was ~8 μm. A(cid:332)er cooling to room temperature, a second TiCl4 treatment was performed as described above. When the anodes had cooled to 80 °C, they were soaked in a dye solu(cid:415)on of 0.1 mM Y123 in 1:1 acetonitrile : tert-butyl alcohol for 18 hours. A(cid:332)er soaking, the anodes were rinsed with acetonitrile and were dried gently under a stream of nitrogen. The PEDOT counter electrodes were prepared by electropolymeriza(cid:415)on in a solu(cid:415)on of 0.01 M EDOT and 0.1 M LiClO4 in 0.1 M SDS in 18 MΩ water. A constant current of 8.3 mA for 250 seconds was applied to a 54 cm2 piece of TEC 8 FTO with predrilled holes using an equal size piece of FTO as the counter electrode. The PEDOT electrodes were then washed with DI water and acetonitrile before being dried under a gentle stream of nitrogen and cut into 1.5 cm by 1.0 cm pieces. The working and counter electrodes were sandwiched together with 25 μm surlyn films by placing them on a 140 °C hotplate and applying pressure. The cells were then filled in a nitrogen filled glove box with electrolyte through one of the two predrilled holes and were sealed with 25 μm surlyn backed by a glass coverslip and applied heat to seal with a soldering iron. The electrolyte consisted of 0.10 M Cu(I), 0.05 M Cu(II), 0.1 M Li(Counter-ion), and 0.5 M 4-tert-butylpyridine in acetonitrile. All batches were made with at least 5 cells per batch. Contact to the TiO2 electrode was made by soldering a thin layer of indium wire onto the FTO. Acetonitrile, deuterated acetonitrile, dichloromethane, methanol, ethanol, diethyl-ether, deionized water, 1,1-bis(2-pyridyl)ethane, 2,6-difluoropyridine, 2.5 M n-butyl lithium in hexanes, tetrakisacetonitrile copper(I) triflate, copper(II) triflate , silver nitrate, tetrabutylammonium 53 hexafluorophosphate, lithium triflate, lithium bistriflimide, silver bistriflimide, copper (I) chloride, copper (I) bistriflimide, and isopropyl alcohol were purchased from Sigma-Aldrich. Star(cid:415)ng material 1,1-Bis(2-pyridyl)ethane, and ligand PY5 were synthesized according to published procedures.88 [Cu(PY5)](OTf) A mixture of PY5 (74.5 mg, 0.168 mmol) and [Cu(ACN)4](OTf) (57.5 mg, 0.153 mmol) in anhydrous acetonitrile was s(cid:415)rred for 30 minutes at room temperature. The solu(cid:415)on was precipitated with anhydrous diethyl ether, forming a yellow solid, and the solid was collected. The solid was dried under vacuum. (97.8 mg 97.4% yield) 1H NMR (500 MHz, acetonitrile-d3): δ = 8.53 (d, 4H); 8.04 (t, 1H); 7.92 (d, 2H); 7.77 (t, 4H); 7.37 (d, 4H); 7.26 (t, 4H); 2.20 (s, 6H); 2.18 (1.5 H). Elem. Anal. Calc. for C30H25CuF3N6O3S C, 54.42; H, 3.98; N, 12.28. Found: C, 54.79; H, 3.91; N, 12.05. TOF-MS- ES+ m/z calc. for [Cu(PY5)] C29H25CuN5 506.14; Found, 506.1407. [Cu(PY5)](TFSI) Silver Bistriflimide (47.0 mg, 0.121 mmol) and copper chloride (12.0 mg, 0.121 mmol) was dissolved in minimal anhydrous ACN and was s(cid:415)rred for 30 minutes at room temperature. A(cid:332)er mixing, the solid silver chloride was removed from the solu(cid:415)on, and the solu(cid:415)on was then added to PY5 (52.5 mg, 0.118 mmol) and s(cid:415)rred overnight. The solu(cid:415)on was precipitated with anhydrous diethyl ether, forming a yellow solid, and the solid was collected. The solid was dried under vacuum. (90.2 mg 96.9% yield). 1H NMR (500 MHz, acetonitrile-d3): δ = 8.53 (d, 4H); 8.04 (t, 1H); 7.92 (d, 2H); 7.77 (t, 4H); 7.37 (d, 4H); 7.26 (t, 4H); 2.20 (s, 6H); 2.18 (1.5 H). Elem. Anal. Calc. for C31H25CuF6N6O4S2 C, 54.42; H, 3.98; N, 12.28. Found: C, 54.79; H, 3.91; N, 12.05. TOF-MS-ES+ m/z calc. for [Cu(PY5)], 506.14; Found, 506.14. 54 [Cu(PY5)](OTf)2 A mixture of PY5 (0.2924 g, 0.66 mmol) and Cu(OTf)2 (0.2210 g, 0.61 mmol) in anhydrous DCM was s(cid:415)rred for 30 minutes at room temperature. The solu(cid:415)on was precipitated with anhydrous diethyl ether, forming a blue solid, and the solid was collected. The solid was dried under vacuum. (0.4776 mg 92.5% yield) Elem. Anal. Calc. for C31H25CuF6N6O6S2 C, 46.24; H, 3.13; N, 7.89. Found: C, 45.93; H, 3.32; N, 8.38. [Cu(PY5)](TFSI)2 A mixture of PY5 (47.8 mg, 0.108 mmol) and Cu(TFSI)2 (67.2 mg, 0.108 mmol) in anhydrous DCM was s(cid:415)rred for 30 minutes at room temperature. The solu(cid:415)on was precipitated with anhydrous diethyl ether, forming a blue solid, and the solid was collected. The solid was dried under vacuum. Any further purifica(cid:415)on was done via recrystalliza(cid:415)on from ACN with diffused ether.(46.3 mg 40.1% yield) Elem. Anal. Calc. for C35H28CuF12N8O8S4 C, 37.93; H, 2.55; N, 10.11. Found: C, 37.38; H, 2.31; N, 9.41. 4.3 Results and Discussion The synthesis of the ligand, PY5, has been previously documented, involving the deprotona(cid:415)on of 1,1-bis(2-pyridyl)ethane by n-BuLi followed by the addi(cid:415)on of 2,6-difluoropyridine as the electrophile.89 The copper complexes were synthesized by reac(cid:415)ng equimolar ra(cid:415)os of the PY5 ligand with copper precursors in different solvents, leading to the forma(cid:415)on of [Cu(PY5)]OTf, [Cu(PY5)]OTf2, [Cu(PY5)]TFSI, and [Cu(PY5)]TFSI2. The copper(I) bistriflimide was synthesized by combining equimolar amounts of silver bistriflimide and copper chloride. The complexes were purified via recrystalliza(cid:415)on from acetonitrile for copper(I) and dichloromethane for copper(II) solu(cid:415)ons and characterized using 1H-NMR and elemental analysis. 55 Scheme 4.3: Syntheses of the [Cu(PY5)]2+/+ complexes, where is the counter ion used for each batch, either triflate (OTf) or bistriflimide (TFSI). Single crystal X-ray diffrac(cid:415)on revealed the solid-state structures of the complexes. A geometric index, τ4, can be used for four-coordinate complexes to determine how much distor(cid:415)on there is from ideal tetrahedral or square planer geometries.74 Where a τ4 of 1 represents an ideal tetrahedral geometry and 0 represents an ideal square planer geometry. The [Cu(PY5)]OTf complex exhibited a highly distorted tetrahedral geometry with a calculated τ4 value of 0.65. The constrained nature of the ligand led to a significant por(cid:415)on of the copper center being solvent- exposed. The choice of solvent during synthesis played a role in preven(cid:415)ng a dispropor(cid:415)ona(cid:415)on reac(cid:415)on in which the copper(I) form dispropor(cid:415)onated to copper metal and a copper(II) complex when the complex was synthesized in dichloromethane. Interes(cid:415)ngly when an NMR was taken of the dispropor(cid:415)onated copper(II) product, it did not match the synthe(cid:415)c [Cu(PY5)]OTf2 product. 56 Acetonitrile was used as the solvent of choice when synthesizing the copper(I) form since it did not cause the copper(I) form to undergo dispropor(cid:415)ona(cid:415)on. Figure 4.1: Crystal structures of the ca(cid:415)ons of the [Cu(PY5)]OTf species (A), and [Cu(PY5)]OTf2 species (B), and the [Cu(PY5)]TFSI2 species which was grown in dichloromethane, a non- coordina(cid:415)ng solvent (C). Depicted ellipsoids are at the 50% probability level. The non- interac(cid:415)ng anions and hydrogens were omi(cid:425)ed for clarity. The solid-state structure of [Cu(PY5)]OTf2 showed pseudo-octahedral geometry, with the PY5 ligand coordinated at five sites and the acetonitrile, the solvent used for synthesis, coordinated at the sixth site. The average bond length for the PY5 ligand is 2.081 Å, and the bond length between the copper and the nitrogen on the acetonitrile is 2.369 Å, a significantly longer distance. 57 To inves(cid:415)gate changes in a system with a vacant axial site, the crystals were regrown using an innocent solvent, dichloromethane. This process revealed an interac(cid:415)on between one of the oxygens on the trifluoromethanesulfonate counter-ion and the copper center in the axial posi(cid:415)on. The bond length was measured at 2.453 Å. This copper-counter-ion bond length increases further when the bistriflimide counter-ion is studied, yielding a copper-oxygen bond length of 2.728 Å. Challenges arose in purifying acetonitrile-bound copper(II) complexes due to the lability of acetonitrile, leading to a mixture of copper complexes with and without acetonitrile bound. To ensure accurate measurements, the copper(II) form was synthesized in dichloromethane. To assess the structural nuances of the copper complex with different counter ions, 1H-NMR analysis was employed. The 1H-NMR spectra for both [Cu(PY5)]OTf and [Cu(PY5)]TFSI, shown in Figures 4.1A and 4.4A respec(cid:415)vely, were recorded in deuterated acetonitrile at room temperature. Both spectra display an integra(cid:415)on for 25 protons. Four peaks represen(cid:415)ng the pyridine “arms” are observed: one at approximately 8.5 ppm and three others between 7.35 and 7.95 ppm, with each integra(cid:415)ng to four protons. Peaks at around 8.00 ppm, integrated to one proton, and 7.25 ppm, integrated to two protons, correspond to the central pyridine. The peak at approximately 2.20 ppm, integra(cid:415)ng to six protons, is a(cid:425)ributed to the ligand's methyl groups. The copper(II) complex is a paramagne(cid:415)c compound making it hard to pull any quan(cid:415)fiable informa(cid:415)on from the spectra, however, it should be noted that the NMR of both of the OTf and TFSI counter ion, irrespec(cid:415)ve of it they are of the copper(I) or copper(II) complexes, yield similar spectra. 58 Figure 4.3: A) Cyclic voltammograms of 2 mM [Cu(PY5)]OTf2, blue, [Cu(PY5)]TFSI2, red, in anhydrous CH3CN. B) [Cu(PY5)]OTf2 in dichloromethane, blue, with 5 equivalents of acetonitrile added to the solu(cid:415)on. All samples contained 0.1 M TBAPF6 as a suppor(cid:415)ng electrolyte, measured at a scan rate of 0.1 V/s on a glassy carbon working electrode. To further understand how the electrochemical behavior of the complex was affected by the different counter ions, cyclic voltammetry was conducted. In the [Cu(PY5)]OTf2 complex, two peaks were observed in the cyclic voltammogram. One peak appeared at -0.372 V vs ferrocene, while the other was observed at -0.662 V vs ferrocene. When the complex was measured in the absence of acetonitrile, using dichloromethane as the solvent, the peak at -0.662 V went away, and when acetonitrile was (cid:415)trated into the solu(cid:415)on, the peak at -0.622 V returned. Which 59 indicates that the peak at -0.662 V was the acetonitrile bound complex, and the peak at -0.372 V was either triflate bound or a 5-coordinate [Cu(PY5)]OTf2 complex. When the [Cu(PY5)]TFSI2 complex was inves(cid:415)gated in anhydrous acetonitrile it exhibited a single peak at -0.382 V vs ferrocene, indica(cid:415)ng that the acetonitrile was not axially bound in the TFSI version of the complex. This could be due to the increased bulk of the TFSI counter-ion when compared to the OTf counter-ion, the TFSI may be physically blocking the acetonitrile from binding with the copper center. The minimal change in redox poten(cid:415)al of the peak around -0.375 V vs ferrocene with various counter ions indicates that the redox couple being measured does not require the presence of acetonitrile, and it is indifferent to the counter-ion present, even though the triflate counter-ion has much stronger interac(cid:415)ons with the copper center. This suggests that the peak at around -0.375 mV corresponds to a 5-coordinate species where the axial posi(cid:415)on is vacant. Figure 4.2: The UV-Vis spectra of the copper(I), A, and copper(II), B, complexes in anhydrous acetonitrile. The triflate complexes are in blue while the bistriflimide complexes are in red. Shi(cid:332)ing our focus to how the different counter ions affect the op(cid:415)cal proper(cid:415)es of the copper complexes, we examined their UV-visible absorp(cid:415)on spectra. The UV-Vis spectra of the copper(I) complexes in acetonitrile displayed peaks below 450 nm which were a(cid:425)ributed to π-π* absorp(cid:415)ons from pyridine units. Metal to ligand charge transfer bands were observed between 60 450-500 nm in copper(I) complexes, with no significant difference between the triflate and bistriflimide variants. The [Cu(PY5)]OTf2 complex exhibits two d-d transi(cid:415)ons, the first at 597 nm, with an ex(cid:415)nc(cid:415)on coefficient of 79.8 M-1s-1, and the second at 909 nm, with an ex(cid:415)nc(cid:415)on coefficient of 13.2 M-1s-1. The [Cu(PY5)]TFSI2 complex only exhibits one peak at 598 nm, with an ex(cid:415)nc(cid:415)on coefficient of 53.0 M-1s-1, although the tailing on the peak could indicate the overlap of two peaks. The observa(cid:415)on of two d-d transi(cid:415)ons is indica(cid:415)ve of either a Jahn-Teller distorted octahedral complex, or a square pyramidal complex. However, the large difference in wavelengths between the two transi(cid:415)ons indicates that this is a square pyramidal geometry. The difference in absorbance of spectra when the counter-ion is changed is most likely due the lack of acetonitrile bound to the axial posi(cid:415)on in the TFSI version of the complex. 61 Figure 4.4: Simulated, red, and experimental, black, EPR spectra of 2 mM [Cu(PY5)]OTf2 in either anhydrous acetonitrile, A, or dichloromethane, B. Spectra C shows the difference between the dichloromethane spectra, red, and the acetonitrile spectra, black, when corrected for solvent affects. Electron Paramagne(cid:415)c Resonance spectra were taken in an a(cid:425)empt to elucidate the solu(cid:415)on geometry of the copper(II) form of the redox couple. Two spectra were taken of the [Cu(PY5)]OTf2 complex, one of the complex synthesized and measured in acetonitrile and the other in dichloromethane. The EPR showed that in both systems the complex is axially elongated which is indica(cid:415)ve of the complex having an elongated octahedral, square pyramidal, or square planer geometry. The two systems have similar g-value but the hyperfine coupling is quite different 62 between the two, indica(cid:415)ng that the complexes are similar in geometry but not iden(cid:415)cal. When the spectra were adjusted to account for the effects of using two different solvents, it becomes easy to see that the species in solu(cid:415)on are not iden(cid:415)cal. However, it remains unclear what the complex’s exact makeup in solu(cid:415)on is. Table 4.1: EPR Parameters of [Cu(PY5)]OTf2 complexes with liquid solu(cid:415)ons in acetonitrile (ACN) or dichloromethane (DCM) at room temperature. All parameters were simulated using MATLAB. g parallel g perpendicular hyperfine parallel (G) hyperfine perpendicular (G) In ACN 2.28 In DCM 2.30 2.04 2.04 437 372 42.6 39.8 To inves(cid:415)gate poten(cid:415)al interac(cid:415)ons between the triflate counterion and the copper center in solu(cid:415)on, a variable temperature 19F-NMR experiment was conducted. The experiment involved cooling the solu(cid:415)on from room temperature to -40°C. Upon lowering the temperature, the fluorine peak corresponding to the triflate counterion exhibited a decrease in intensity coupled with an increase in sharpness. This observa(cid:415)on suggests that the counter-ion does not undergo exchange with the axial site on the copper center, implying that the counter-ion doesn't interact with the copper center in solu(cid:415)on. Such an exchange would typically result in broadening of the peak as the temperature decreases. Notably, this trend remained consistent irrespec(cid:415)ve of whether acetonitrile or dichloromethane was used as the solvent. When this evidence is compiled with the previous results of the UV-Vis, cyclic voltammetry, and EPR experiments it becomes clear that the solu(cid:415)on geometry of the [Cu(PY5)]2+/+ complexes a square pyramidal where the axial site is vacant. 63 2100000 2000000 1900000 1800000 1700000 1600000 1500000 1400000 1300000 1200000 1100000 1000000 900000 800000 700000 600000 500000 400000 300000 200000 100000 0 -100000 -200000 10 0 -10 -20 -30 -40 -50 -60 -70 -80 -90 -100 f1 (ppm ) -110 -120 -130 -140 -150 -160 -170 -180 -190 -200 -210 Figure 4.5: 19F-NMR spectra of [Cu(PY5)]OTf2 in anhydrous deuterated acetonitrile at various temperatures, 25°C-red, 0°C-green, -20°C-blue, and -40°C-purple. 64 150000 100000 50000 0 -50000 -100000 10 0 -10 -20 -30 -40 -50 -60 -70 -80 -90 -100 f1 (ppm ) -110 -120 -130 -140 -150 -160 -170 -180 -190 -200 -210 Figure 48.6: 19F-NMR spectra of [Cu(PY5)]OTf2 in deuterated dichloromethane at various temperatures, 25°C-red, 0°C-green, -20°C-blue, and -40°C-purple. In order to understand how the dynamic nature of the interac(cid:415)on with the open axial site the self- exchange kine(cid:415)cs of both systems were inves(cid:415)gated. First the [Cu(PY5)]OTf complex was inves(cid:415)gated u(cid:415)lizing stopped-flow spectroscopy. Stopped-flow spectroscopy was u(cid:415)lized to measure the cross-exchange electron transfer rate constant, k12, between [Cu(PY5)]OTf2 and octamethylferrocene (FcMe8). [𝐶𝑢(𝑃𝑌5)](cid:2870)(cid:2878) + 𝐹𝑐𝑀𝑒8 (cid:3038)(cid:3117)(cid:3118) (cid:4653)⎯⎯⎯(cid:4654) [𝐶𝑢(𝑃𝑌5)](cid:2878) + 𝐹𝑐𝑀𝑒8(cid:2878) (4.1) 65 Figure 4.7: An absorbance vs (cid:415)me plot at 450 nm, showing the increase of [Cu(PY5)]OTf2 species (black dot) and fi(cid:427)ng (red line) for the reduc(cid:415)on of [Cu(PY5)]OTf2 by octamethylferrocene (FcMe8). Octamethylferrocene was chosen for the cross-exchange to op(cid:415)mize the driving force and because it has a well-known self-exchange rate constant for electron transfer.90 Due to the large poten(cid:415)al difference between [Cu(PY5)]2+/+ and [FcMe8]+/0, it was assumed that the reac(cid:415)on went to comple(cid:415)on, with no significant back reac(cid:415)on. Figure 4.7 shows a fit of the absorbance at 450 nm vs (cid:415)me plot, which represents the growth of the [Cu(PY5)]+ species in solu(cid:415)on due to the reduc(cid:415)on of [Cu(PY5)]2+ by FcMe8, using the following equa(cid:415)on: 𝐴 = 𝐴(cid:2998) + (𝐴(cid:3016) − 𝐴(cid:2998))𝑒(cid:2879)(cid:3038)(cid:3290)(cid:3277)(cid:3294)(cid:3047) (4.2) 66 For all reac(cid:415)ons, FcMe8 was held in excess to maintain pseudo-first order condi(cid:415)ons, allowing kobs to be represented by: 𝑘(cid:3042)(cid:3029)(cid:3046) = 𝑘(cid:2869)(cid:2870)[𝐹𝑐𝑀𝑒8] (4.3) A linear fit of the fi(cid:425)ed kobs values vs concentra(cid:415)on of FcMe8 provided the value of the cross- exchange rate constant, k12, from the slope of the line. The ini(cid:415)al reac(cid:415)on mixtures and the pseudo-first order observed rate constants can be found in Table 4.2. Table 4.2: The ini(cid:415)al reac(cid:415)on mixture and the observed rate constant, kobs, for the cross exchange between [Cu(PY5)]OTf2 and FcMe8. [Cu(PY5)(OTf)2] / M [FcMe8] / M 2.74E-05 2.58E-04 3.10E-04 3.61E-04 4.13E-04 4.65E-04 Using the experimentally determined cross-exchange rate constant from Equa(cid:415)on 4.3, k12, and the previously determined self-exchange constant for [FcMe8]+/0, k11, the Marcus cross-rela(cid:415)on 𝑘(cid:2869)(cid:2870) = (𝑘(cid:2869)(cid:2869)𝑘(cid:2870)(cid:2870)𝐾(cid:2869)(cid:2870)𝑓(cid:2869)(cid:2870))(cid:2869)/(cid:2870)𝑊(cid:2869)(cid:2870) (4.4) was used to calculate the self-exchange rate constant for [Cu(PY5)]2+/+, k22. Where K12 is the equilibrium constant, f12 is a nonlinear correc(cid:415)on term, and W12 is an electrosta(cid:415)c work term for bringing the reactants into contact. The nonlinear correc(cid:415)on term and the work term were calculated (see below for details on the calcula(cid:415)on) and determined to be 2.60 and 0.99, respec(cid:415)vely as seen in Table 4.3. However, it should be noted that there will be some error in the calcula(cid:415)on of the work term since the Debye-Huckel model is not expected to have accurate 67 results when the solu(cid:415)on has a high ionic strength, as was the case for this experiment (0.1 M suppor(cid:415)ng electrolyte). The equilibrium constant, K12, can be determined using Equa(cid:415)on 4.5: 𝑛𝐹𝛥𝐸 = 𝑅𝑇𝑙𝑛𝐾(cid:2869)(cid:2870) (4.5) where n is the number of electrons transferred, F is Faraday’s constant, ΔE is the difference in formal poten(cid:415)al between the oxidant and reductant in solu(cid:415)on, R is the ideal gas constant, and T is the temperature in Kelvin. The redox poten(cid:415)al for [Cu(PY5)]2+/+ was found to be -0.372 V vs ferrocene91 and the redox poten(cid:415)al of [FcMe8]+/0 has been previously determined to be -0.377 V vs ferrocene, using cyclic voltammetry which gave a calculated K12 of 0.62. The self-exchange rate constant for [FcMe8]+/0 was previously iden(cid:415)fied to be 2.01 × 107 M-1s-1 using NMR techniques.92 Using these values, the self-exchange rate for the [Cu(PY5)]OTf1/2 couple was determined to be 88.1 (±7.3) M-1s-1. The work required to move the reactant complexes to a distance, r, for the electron transfer reac(cid:415)on was calculated using the Equa(cid:415)on 4.6. 𝑊(cid:2869)(cid:2870) = exp [− (cid:3050)(cid:3117)(cid:3118)(cid:2878)(cid:3050)(cid:3118)(cid:3117)(cid:2879)(cid:3050)(cid:3117)(cid:3117)(cid:2879)(cid:3050)(cid:3118)(cid:3118) (cid:2870)(cid:3019) ] (4.6) 𝑤(cid:3036)(cid:3037)(𝑟) = (cid:3053)(cid:3284)(cid:3053)(cid:3285)(cid:3044)(cid:3118)(cid:3015)(cid:3250) (cid:2872)(cid:3095)(cid:3084)(cid:3116)(cid:3084)(cid:3045)((cid:2869)(cid:2878)(cid:3081)(cid:3045)) (4.7) Equa(cid:415)on 4.7 was used to determine the work associated with the forward and reverse cross- exchange reac(cid:415)on, w12 and w21 respec(cid:415)vely, and the self-exchange reac(cid:415)ons, w11 and w22. In this equa(cid:415)on zi and zj are the charges of the interac(cid:415)ng complexes, q is the charge of an electron, NA is Avogadro’s constant, ε0 is the permi(cid:427)vity of free space, ε is the sta(cid:415)c dielectric of the medium, 𝛽 = (cid:4672) (cid:2870)(cid:3044)(cid:3118)(cid:3015)(cid:3250)(cid:3010) (cid:2869)(cid:2868)(cid:2868)(cid:2868)(cid:3084)(cid:3116)(cid:3084)(cid:3038)(cid:3251)(cid:3021) (cid:2869)/(cid:2870) (cid:4673) , I is the ionic strength of the solu(cid:415)on and kB is the Boltzmann’s constant. The calcula(cid:415)on has the following assump(cid:415)ons: the work is assumed to be primarily Coulombic, the 68 distance, r, is assumed to be the center-to-center distance between the complexes, and the reactants are assumed to be spherical. The non-linear correc(cid:415)on term, f12, was calculated using the equa(cid:415)on 𝑙𝑛𝑓(cid:2869)(cid:2870) = (cid:2869) (cid:2872) (cid:4672)(cid:3039)(cid:3041)(cid:3012)(cid:3117)(cid:3118)(cid:2878) (cid:3286)(cid:3117)(cid:3117)(cid:3286)(cid:3118)(cid:3118) (cid:3275)(cid:3118) (cid:2922)(cid:2924)(cid:4672) (cid:3118) (cid:4673) (cid:3298)(cid:3117)(cid:3118)(cid:3127)(cid:3298)(cid:3118)(cid:3117) (cid:3267)(cid:3269) (cid:3298)(cid:3117)(cid:3117)(cid:3126)(cid:3298)(cid:3118)(cid:3118) (cid:3267)(cid:3269) (cid:4673)(cid:2878) (4.8) where K12 is the equilibrium constant, w12 and w21 are the work associated with the forward and reverse cross-exchange reac(cid:415)on respec(cid:415)vely, R is the ideal gas constant, T is temperature in Kelvin, and Z is the frequency factor, which is assumed to be 1013 M-1s-1 due to the larger inner- sphere contribu(cid:415)ons to the total reorganiza(cid:415)on energy. Table 4.3: Summary of the kine(cid:415)c data used to calculate the self-exchange rate constant of the [Cu(PY5)]OTf1/2 complex by monitoring the reac(cid:415)on of [Cu(PY5)]OTf2 and FcMe8 in acetonitrile containing 0.1 M LiOTf at 25°C. Kine(cid:415)c Parameter Values K12 0.62 k12 / M-1s-1 4.53E+04 k22 / M-1s-1 2.01E+07 f12 W12 k11 / M-1s-1 λse / eV λi / eV λo / eV 1 2.6 88.1 2.57 1.55 1.02 The self-exchange of fast exchanging redox complexes can be found by inves(cid:415)ga(cid:415)ng the methyl peaks in the 1H-NMR spectra of the pure diamagne(cid:415)c [Cu(PY5)]TFSI complex, the pure 69 paramagne(cid:415)c [Cu(PY5)]TFSI2 complex, and mixtures of the diamagne(cid:415)c and paramagne(cid:415)c species. All peaks were fi(cid:425)ed to a Lorentzian/Gaussian lineshape. The self-exchange rate constant, kse, was calculated from 𝑘(cid:3046)(cid:3032) = (cid:2872)(cid:3095)(cid:3076)(cid:3279)(cid:3076)(cid:3291)((cid:3057)(cid:3092))(cid:3118) (cid:3435)(cid:3024)(cid:3279)(cid:3291)(cid:2879)(cid:3076)(cid:3291)(cid:3024)(cid:3291)(cid:2879)(cid:3076)(cid:3279)(cid:3024)(cid:3279)(cid:3439)(cid:3004) (4.9) where Wdp is the line width (full width at half-maximum) of the mixed species methyl resonance peak, Wp and Wd are the line widths of the paramagne(cid:415)c and diamagne(cid:415)c peaks, respec(cid:415)vely, Xp and Xd are the mole frac(cid:415)ons of the paramagne(cid:415)c and diamagne(cid:415)c species, respec(cid:415)vely, C is the total concentra(cid:415)on of the exchanging species, and Δν is the observed frequency shi(cid:332) rela(cid:415)ve to the posi(cid:415)on of the resonance for the diamagne(cid:415)c species. The diamagne(cid:415)c and paramagne(cid:415)c line widths and the frequency shi(cid:332) are listed in Table 4.4. This analysis produced a kse of 389 M- 1s-1 for [Cu(PY5)]TFSI1/2 in deuterated acetonitrile at 22°C. Table 4.4: List of parameters used to calculate the self-exchange of [Cu(PY5)]TFSI1/2. Wp Wd Xd Xp ΔV Wdp Kex 102.27 11.3 0.404249 0.595751 22.562 204.66 389.0535 The total reorganiza(cid:415)on energy for each of the self-exchange reac(cid:415)ons was calculated following the equa(cid:415)on 𝑘(cid:2869)(cid:2869) = 𝐾(cid:2869)(cid:2870)𝑍𝛤𝑒(cid:2879)((cid:3090)(cid:3117)(cid:3117) (cid:2872)(cid:3038)(cid:3251)(cid:3021)) ⁄ (4.10) where k11 is the previously calculated self-exchange rate, K12 is the equilibrium constant, Z is the frequency factor, which is assumed to be 1013 M-1s-1 due to the larger inner-sphere contribu(cid:415)ons 70 to the total reorganiza(cid:415)on energy, Γ is a correc(cid:415)on for nuclear tunneling, which is assumed to be ~1, λ11 is the total reorganiza(cid:415)on energy of the self-exchange reac(cid:415)on, kB is the Boltzmann constant, and T is the temperature of solu(cid:415)on in Kelvin. The resul(cid:415)ng λ11 was found to be 2.57 eV for the OTf complex and 2.41 for the TFSI complex. The outer-sphere reorganiza(cid:415)on energy for two spherical reactants can be determined using 𝜆(cid:2869)(cid:2869),(cid:3042) = ((cid:3057)(cid:3053)(cid:3044))(cid:3118) (cid:2872)(cid:3095)(cid:3084)(cid:3116) (cid:4672) (cid:2869) (cid:2870)(cid:3028)(cid:3117) + (cid:2869) (cid:2870)(cid:3028)(cid:3118) − (cid:2869) (cid:3045)(cid:3117)(cid:3117) (cid:4673) (cid:3436) (cid:2869) (cid:3005)(cid:3290)(cid:3291),(cid:3294)(cid:3290)(cid:3287) − (cid:2869) (cid:3005)(cid:3294),(cid:3294)(cid:3290)(cid:3287) (cid:3440) (4.11) where λ11,o is the outer-sphere reorganiza(cid:415)on energy, Δz is the change in charge of the complex, q is the charge of an electron, ε0 is the permi(cid:427)vity of free space, a1 and a2 are the atomic radii of the reactants, r11 is the center−center distance between the reactants, Dop,sol is the op(cid:415)cal dielectric constant of the medium, which is equal to the square of the refrac(cid:415)ve index of the medium, and Ds,sol is the sta(cid:415)c dielectric constant of the medium. The outer-sphere reorganiza(cid:415)on energy for the self-exchange reac(cid:415)on was calculated to be 1.02 eV for the OTf complex and 1.01 eV for the TFSI complex. The inner-sphere reorganiza(cid:415)on energy was determined using 𝜆(cid:2869)(cid:2869),(cid:3036) = 𝜆(cid:2869)(cid:2869) − 𝜆(cid:2869)(cid:2869),(cid:3016) yielding a result of 1.55 eV for the OTf complex and 1.40 eV for the TFSI complex for the inner-sphere reorganiza(cid:415)on energy of the self-exchange reac(cid:415)on. 71 Figure 4.8: A) Cyclic Voltammogram of 2 mM [Cu(PY5)](OTf)2 with 0.1 M TBAPF6 in anhydrous acetonitrile with increasing concentra(cid:415)on of 4-tert-butylpyridine (TBP) at a glassy carbon working electrode; scan rate = 100 mVs-1. B) UV-Vis absorp(cid:415)on spectra of 2 mM [Cu(PY5)](OTf)2 in anhydrous acetonitrile with increasing equivalents of TBP. C) 1H-NMR spectra of [Cu(PY5)](OTf)2 in anhydrous acetonitrile with increasing equivalents of TBP. Unbound PY5 and TBP are shown as controls. Lewis bases such as TBP are used as addi(cid:415)ves in DSSCs to increase the performance of the DSSC devices.35–37 Several copper based redox shu(cid:425)les have been shown to undergo a ligand subs(cid:415)tu(cid:415)on with the TBP, even polydentate ligands.38,39 This can have a significant impact on the 72 redox poten(cid:415)al and kine(cid:415)cs of the system as mul(cid:415)ple redox systems are present at any given (cid:415)me. Looking at the cyclic voltammetry data, Upon the addi(cid:415)on of TBP to the solu(cid:415)on containing [Cu(PY5)]OTf2, both of the redox waves go away and a new wave grows in at -0.512 V vs ferrocene, indica(cid:415)ng that the [Cu(PY5)]OTf2 complex is reac(cid:415)ng with TBP. The new peak con(cid:415)nues to grow in as TBP is added un(cid:415)l about 10 equivalents rela(cid:415)ve to the amount of copper(II) in the solu(cid:415)on, indica(cid:415)ng that whatever the change that occurred has fully stabilized by 10 equivalents added. As such the amount of TBP used in the devices in this study was 10 equivalents, in order to minimize mixed species in solu(cid:415)on. When the addi(cid:415)on of TBP is monitored via UV-Vis spectroscopy, there is a no(cid:415)ceable blue shi(cid:332) of 23 nm for the peak at ~600 nm, which also comes with an almost 50% increase in absorbance, and a blue shi(cid:332) of 58 nm for the peak at ~900 nm, with minimal change in absorbance. Once 10 equivalents of TBP is added, rela(cid:415)ve to the copper(II) in solu(cid:415)on, the reac(cid:415)on seems to have reached equilibrium and the peaks no longer shi(cid:332). This agrees with the previous cyclic voltammetry study indica(cid:415)ng that there is a reac(cid:415)on occurring which has reached equilibrium by 10 equivalents of TBP added to the solu(cid:415)on. Although it should be noted that when the reac(cid:415)on has stabilized the resul(cid:415)ng spectra does not match the spectra of [Cu(TBP)4]OTf2, meaning that whatever is formed it is not the ligand subs(cid:415)tuted product. In an a(cid:425)empt to understand what the interac(cid:415)on of the [Cu(PY5)]OTf2 complex and TBP was producing, the reac(cid:415)on was monitored via 1H-NMR. Upon the addi(cid:415)on of TBP to the [Cu(PY5)]OTf2 solu(cid:415)on a change in the spectra is seen as soon as a single equivalent of TBP is added. However, a(cid:332)er that ini(cid:415)al change in the spectra, the only change in the spectra is the broad peaks that seem to correspond to TBP. The broadness of the peaks seems to indicate that 73 there is a rapid exchange on the NMR (cid:415)mescale of the TBP and the copper center. However, the spectra show no signs of unbound ligand in the solu(cid:415)on as TBP added, sugges(cid:415)ng the PY5 ligand remains bound to the copper center. This indicates that the interac(cid:415)on being seen likely stems from TBP displacing the labile axial acetonitrile, resul(cid:415)ng in a [Cu(PY5)(TBP)]2+ copper complex, rather than the displacement of the ligand. It is also worth no(cid:415)ng that it has been posited that the interac(cid:415)on of TBP with copper complexes may not necessarily be undergoing a full subs(cid:415)tu(cid:415)on, instead it may be just a coordina(cid:415)on of the TBP to the copper center causing the change in proper(cid:415)es, similar to what is being seen with [Cu(PY5)]OTf2.39,93 CuPY5(OTf) without TBP CuPY5(OTf) with TBP CuPY5(TFSI) without TBP CuPY5(TFSI) with TBP 2 - m c A m / t n e r r u C 10 9 8 7 6 5 4 3 2 1 0 -0.5 -0.4 -0.3 -0.2 -0.1 0 Voltage / V Figure 4.9: J−V curves for the [Cu(PY5)]OTf1/2, blue, and [Cu(PY5)]TFSI1/2, red, based DSSC devices. The devices represented by the pale colored curves contain TBP while the dark colored curves are without TBP. 74 Table 4.5: Summary of device performance of devices comparing [Cu(PY5)]OTf and [Cu(PY5)]TFSI, parameters represent the average of five devices. η / % JSC / mAcm-2 VOC / V FF CuPY5(OTf) 1.30(±0.25) 8.22(±1.16) 0.26(±0.02) 0.60(±0.01) CuPY5(OTf) with TBP 1.40(±0.30) 6.22(±1.97) 0.35(±0.05) 0.66(±0.04) CuPY5(TFSI) 1.01(±0.35) 6.57(±2.18) 0.26(±0.01) 0.59(±0.02) CuPY5(TFSI) with TBP 1.33(±0.09) 5.84(±0.49) 0.36(±0.03) 0.64(±0.01) DSSC devices were fabricated with the commercial dye, Y123, being used as the light-harvester, and where poly(3,4-ethylenedioxythiophene) (PEDOT) was u(cid:415)lized at the counter electrode. The devices u(cid:415)lized either [Cu(PY5)]OTf1/2 or [Cu(PY5)]TFSI1/2 as the redox shu(cid:425)le used, and for each redox shu(cid:425)le a batch was fabricated with 10 equivalents of TBP rela(cid:415)ve to the amount of copper(II) in solu(cid:415)on, and another batch without the addi(cid:415)on of TBP. The devices yielded sta(cid:415)s(cid:415)cally similar results for the JSC, around 6.71 mAcm-2, and efficiencies, around 1.26% for all the batches. The difference between the cells with and without TBP is no(cid:415)ceable in the open circuit voltage (VOC) where the devices with TBP have an increase of around 100 mV compared to their counterparts without TBP. While the devices with and without TBP show differences from one another the devices with differing redox mediator showed sta(cid:415)s(cid:415)cally iden(cid:415)cal performance when compared, as shown in Table 4.5. 75 This similarity in performance across the different redox shu(cid:425)les indicates that even though there was a difference in both the redox and kine(cid:415)c proper(cid:415)es of the two versions of the redox shu(cid:425)le, the choice of counter-ion did not significantly affect device characteris(cid:415)cs. When looking at the solu(cid:415)on poten(cid:415)al of the devices it can be seen that the solu(cid:415)on poten(cid:415)al is similar between the OTf and TFSI versions of the complex, -0.372 V and -0.398 V vs ferrocene respec(cid:415)vely, which for both complexes is slightly nega(cid:415)ve of the Nerns(cid:415)an poten(cid:415)al of -0.357 and -0.365 respec(cid:415)vely. The predominate form of the redox shu(cid:425)le that seems to be controlling the solu(cid:415)on poten(cid:415)al of the devices is the five-coordinate redox couple represented by the wave at ~-0.375 V vs ferrocene, with minimal contribu(cid:415)on from the acetonitrile bound wave at -0.662 V. When TBP is added to the devices the solu(cid:415)on poten(cid:415)al nega(cid:415)vely shi(cid:332)s to -0.496 V and -0.501 V vs ferrocene for the OTf and TFSI complexes respec(cid:415)vely, which matches well with the predicted -0.503 V vs ferrocene for the [Cu(PY5)(TBP)]OTf1/2 complex. This was expected due to the binding of TBP to the copper center, so both versions of the complex would form the [Cu(PY5)(TBP)]2+ complex. Due to the similarity of the device results between the varia(cid:415)ons of the complex only [Cu(PY5)]OTf1/2 was used for subsequent experiments. 76 12 10 8 6 4 2 0 2 - m c A m / t n e r r u C -2 -0.5 0 M TBP 0.055 M TBP 0.526 M TBP -0.4 -0.3 -0.2 -0.1 0 Voltage / V Figure 4.10: J-V curves of best performing devices containing [Cu(PY5)]OTf1/2 in dry acetonitrile and 0 M TBP (green), 0.055 M TBP (blue), 0.211 M TBP (gray), 0.526 M TBP (yellow). Table 4.6: Summary of device performance of devices containing [Cu(PY5)]OTf with various equivalents of TBP added, parameters represent the average of five devices. η / % Jsc / mA*cm-2 Voc / V FF 0 M TBP 2.00(±0.06) 8.40(±0.09) 0.39(±0.00) 0.61(±0.02) 0.055 M TBP 0.526 M TBP 2.12(±0.10) 8.91(±0.39) 0.41(±0.02) 0.58(±0.04) 2.01(±0.24) 7.81(±0.63) 0.45(±0.01) 0.57(±0.03) The effect of the addi(cid:415)on of TBP in the DSSC devices was inves(cid:415)gated via fabrica(cid:415)ng devices with various concentra(cid:415)ons of TBP used. All of the devices had an improved power conversion efficiency of around 2.04%, which can be a(cid:425)ributed to both an increased short circuit current (JSC), around 8.38 mAcm-2, and a significantly increased open circuit voltage (VOC), which is over 100 mV improved over the previous a(cid:425)empt. However, it can be noted that the current of the devices were hampered by mass transport issues. The incident photon-to-current conversion efficiency (IPCE) spectra was measured, which measures the current at various wavelengths to 77 determine how much of the photons absorbed at each wavelength are converted to current. The IPCE in figure 4.12 shows that as TBP is added there is a general trend towards an increased photon-to-current conversion between 400 and 650 nm. This increase could be caused by decreased recombina(cid:415)on to the redox shu(cid:425)le as TBP is being added to the solu(cid:415)on. The IPCE spectra can also be integrated to determine what the theore(cid:415)cal current of the device is, with the results listed in table 4.7, and all of the devices show higher theore(cid:415)cal current than the measure short circuit current, ge(cid:427)ng as high as 10 mAcm-2 when 0.526 M of TBP is added to the solu(cid:415)on. The high theore(cid:415)cal current and the nega(cid:415)ve redox poten(cid:415)al make this redox shu(cid:425)le a candidate for u(cid:415)liza(cid:415)on with narrow band gap dyes for high photocurrent DSSC devices. 78 % / E C P I 90 80 70 60 50 40 30 20 10 0 0 M TBP 0 M TBP 0.055 M TBP 0.526 M TBP 0.055 M TBP 0.526 M TBP 12 10 2 - m c A m / t n e r r u C 8 6 4 2 0 400 450 500 550 Wavelength / nm 600 650 700 Figure 4.11: IPCE curves for the best performing [Cu(PY5)]OTf1/2 based DSSC devices with various amounts of TBP added. The devices represented by the green diamonds and curve contain 0 M TBP added, the blue square and curve represent 0.055 M TBP added, the gray triangles and curve represent 0.211 M TBP added, and the yellow circles and curve represent 0.526 M TBP added. The points represent IPCE responses while the solid lines represent the integrated photocurrent. 79 Table 4.7: The measured current and the integrated current found via IPCE for devices containing [Cu(PY5)]OTf1/2 and various equivalents of TBP added. Jsc / mA*cm-2 Integrated Current / 0 M TBP 8.40(±0.09) 0.055 M TBP 8.91(±0.39) 0.526 M TBP 7.81(±0.63) mA*cm-2 8.92 9.82 10.64 The solu(cid:415)on poten(cid:415)al of the [Cu(PY5)]OTf1/2 solu(cid:415)on without TBP was -0.374 V vs ferrocene matching well with the calculated Nerns(cid:415)an poten(cid:415)al of -0.357 V vs ferrocene. Upon the addi(cid:415)on of 0.526 M TBP the solu(cid:415)on poten(cid:415)al shi(cid:332)ed to -0.490 V vs ferrocene as the TBP bound to the copper center, the calculated poten(cid:415)al was -0.503 V vs ferrocene. This shi(cid:332) in redox poten(cid:415)al and the change in the open circuit poten(cid:415)al from 0.39 V to 0.45 V as TBP is added to the solu(cid:415)on indicates that the fermi level of the devices increased from -0.764 V vs ferrocene to -0.940 V vs ferrocene. The addi(cid:415)on of TBP seems to have shi(cid:332)ed the fermi level ~200 mV which is a small shi(cid:332) in the fermi level when compared to the [Cu(bpyPY4)]TFSI1/2 complex and the [Cu(dbmed)]OTf1/2 complex. When the electron life(cid:415)me is compared to the previous complexes it can be seen that the life(cid:415)me of the [Cu(PY5)]OTf1/2 complex is similar to that of the [Cu(bpyPY4)]TFSI1/2 complex. The addi(cid:415)on of TBP shows an increase in the electron life(cid:415)me by a factor of 10. This is likely due to a combina(cid:415)on of low driving force for recombina(cid:415)on for both the complexes due to their nega(cid:415)ve redox poten(cid:415)al, and their rela(cid:415)vely slow electron transfer kine(cid:415)cs. 80 10 1 0.1 s / 1 / e m i t e f i L 0.01 0 [Cu(dbmed)] without TBP [Cu(bpyPY4)] without TBP [Cu(PY5)] without TBP [Cu(dbmed)] with TBP [Cu(bpyPY4)] with TBP [Cu(PY5)] with TBP 0.2 0.4 0.6 0.8 1 1.2 Voltage / V Figure 4.12: A plot of the electron life(cid:415)me vs poten(cid:415)al plot for DSSC devices using [Cu(dbmed)]OTf1/2 (green), [Cu(bpyPY4)]TFSI1/2 (red), [Cu(PY5)]OTf1/2 (blue), with (pale color) and without (dark color) the addi(cid:415)on of TBP. 4.4 Conclusion A copper-based redox shu(cid:425)le featuring a pentadentate polypyridyl ligand with a labile axial posi(cid:415)on was developed for use in DSSC devices. This complex was studied using two different counter ions which provide different redox and kine(cid:415)c proper(cid:415)es for the copper complex, including a 4-fold increase in self-exchange rate. Despite these differences in kine(cid:415)cs and redox behavior, both variants of the copper complex showed iden(cid:415)cal performance in DSSC applica(cid:415)ons. In solu(cid:415)on, the copper(II) complexes predominantly adopted a five-coordinate geometry with a vacant axial site. Further analysis revealed that while there is an interac(cid:415)on between the copper complex and TBP, this interac(cid:415)on is limited to the subs(cid:415)tu(cid:415)on at the labile axial posi(cid:415)on by the base, rather than involving the ligand itself. Notably, the addi(cid:415)on of TBP to 81 the DSSC’s electrolyte led to an increase in device performance. This suggests that the presence of a Lewis base, even when reac(cid:415)ng with the copper complexes, is beneficial for the overall efficiency of DSSCs. 82 Chapter 5 – Developing a model to predict ligand exchange in DSSCs 5.1 Introduc(cid:415)on In the wake of Grätzel's pioneering 1993 report on achieving a 10% efficient dye-sensi(cid:415)zed solar cell (DSSC) through the use of dye-sensi(cid:415)zed photoelectrodes, further advancements in efficiency were hampered by the reliance on the triiodide/iodide (I3-/I-) redox shu(cid:425)le.15 A primary constraint of I3-/I- u(cid:415)liza(cid:415)on arises from the considerable overpoten(cid:415)al required for efficient dye regenera(cid:415)on. Over the last decade, the adop(cid:415)on of outer sphere redox shu(cid:425)les has enabled the design of systems that mi(cid:415)gate this overpoten(cid:415)al, resul(cid:415)ng in performance enhancement.23 However, the incorpora(cid:415)on of Co(III/II) based redox shu(cid:425)les has encountered challenges due to substan(cid:415)al inner-sphere reorganiza(cid:415)on energy accompanying electron transfer. This complica(cid:415)on, arising from the transi(cid:415)on of a high-spin d7 Co(II) system to a low-spin d6 Co(III) system, directly affects dye regenera(cid:415)on kine(cid:415)cs by limi(cid:415)ng the available electron transfer driving force.23 In contrast, there has been a growing interest in copper-based redox shu(cid:425)les for DSSCs. These shu(cid:425)les have achieved record efficiencies of 15.2% under 1 sun illumina(cid:415)on and an impressive 34.5% under 1000 lux intensity fluorescent ligh(cid:415)ng.20,87 The success of copper-based redox shu(cid:425)les can be a(cid:425)ributed to their diminished inner-sphere electron transfer reorganiza(cid:415)on energies, placing them on par with conven(cid:415)onal I3−/I− and Co(III/II)-based counterparts. Copper complexes are versa(cid:415)le, adop(cid:415)ng diverse geometries influenced by oxida(cid:415)on state and ligand environment. Cu(II) d9 complexes typically exhibit six-coordinate, octahedral or tetragonal, five-coordinate, square pyramidal or trigonal bipyramidal, or four-coordinate, tetrahedral or 83 square planar geometries. Similarly, Cu(I) d10 complexes predominantly assume four-coordinate, tetrahedral or square planar geometries.60–63 Most copper redox shu(cid:425)les in DSSCs u(cid:415)lize bidentate bipyridine or phenanthroline ligands with bulky flanking groups to minimize structural varia(cid:415)on between Cu(II) and Cu(I) states. Notably, recent experimental findings indicate that commonly employed Lewis base addi(cid:415)ves, like 4-tert- butylpyridine, can coordinate to Cu(II) species, poten(cid:415)ally having a detrimental impact on the electrochemical behavior and device performance.38 This coordina(cid:415)on-induced complexity influences redox species and shi(cid:332)s formal redox poten(cid:415)als, limi(cid:415)ng photovoltage. However, not all copper redox couples undergo ligand subs(cid:415)tu(cid:415)on.39,41,94,95 The focus of this work lies in harnessing a profound insight contributed by Dr. Rorabacher. Dr. Rorabacher demonstrated a strong correla(cid:415)on between the poten(cid:415)al of a copper redox couple and the stability constant of its Cu(II) form.96 It was found that the Cu(I) form of copper redox shu(cid:425)les tend to have a similar stability constant, where the stability constant is determined by the steric strain. Dr. Rorabacher was also able to show that, in aqueous solu(cid:415)ons, the stability constant of the Cu(II) form was found to be inversely propor(cid:415)onal to the redox poten(cid:415)al of the Cu(II/I) complex. This rela(cid:415)onship was found to be linear, shown in Figure 5.1, with a slope of - 0.059, matching the electrochemical constant of the Nernst equa(cid:415)on, which indicates that this is a fundamental property. The trend was tested against a variety of ligand types from cyclic ligands to tripodal ligands, and the trend held true across all the different types. This means that ligands that bind with copper to form complexes with more nega(cid:415)ve redox poten(cid:415)als should be able to subs(cid:415)tute ligands bound to copper complexes with more posi(cid:415)ve redox poten(cid:415)als. This rela(cid:415)onship between redox poten(cid:415)al and stability constant presents a novel pathway for 84 predic(cid:415)ng ligand subs(cid:415)tu(cid:415)on in copper complexes, specifically in the presence of 4-tert- butylpyridine (TBP). Figure 5.1: A plot of copper redox poten(cid:415)als versus the log of the stability constant of the copper(II) form of the redox complex in aqueous solu(cid:415)ons. The ligands represent samples of various ligand types. The solid line has the Nerns(cid:415)an slope of -0.059, indica(cid:415)ng that this rela(cid:415)onship follows the fundamental trend.97 Building upon Dr. Rorabacher's work, our study delves into the interplay between redox poten(cid:415)al and stability constant. The ability to predict ligand subs(cid:415)tu(cid:415)on by TBP, a crucial factor in the 85 electrochemical behavior of copper complexes, opens the door to engineering redox mediators with enhanced performance in DSSCs. 5.2 Experimental Details All NMR spectra were taken on an Agilent DirectDrive2 500 MHz spectrometer at room temperature and referenced to residual solvent signals. All NMR spectra were evaluated using the MestReNova so(cid:332)ware package features. Cyclic voltammograms were obtained using µAutolabIII poten(cid:415)ostat using BASi glassy carbon electrode, a pla(cid:415)num mesh counter electrode, and a fabricated 0.01 M AgNO3, 0.1 M TBAPF6 in acetonitrile Ag/AgNO3 reference electrode. All measurements were internally referenced to ferrocene/ferrocenium couple via addi(cid:415)on of ferrocene to solu(cid:415)on a(cid:332)er measurements or run in a parallel solu(cid:415)on of the same solvent/electrolyte. UV-Vis spectra were taken using a PerkinElmer Lambda 35 UV-Vis spectrometer using a 1 cm path length quartz cuve(cid:425)e at 480 nm/min. Elemental Analysis data were obtained via Midwest Microlab. Synthesis of [Cu(dmbpy)2]OTf: Copper(I) tetrakisacetonitrile triflate (1.0284 g, 2.73 mmol) was dissolved in minimal dry acetonitrile. Meanwhile dmbpy (1.0605 g, 5.76 mmol) was dissolved in minimal dry dichloromethane. The dmbpy solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a red color. Reac(cid:415)on was s(cid:415)rred for 30 minutes then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(I)(dmbpy)2OTf product was collected (1.4716 g, 96.1% yield). 1H NMR (500 MHz, CDCl3) δ 8.26 (d, 2H), 8.02 (t, 2H), 7.44 (d, 2H), 2.22 (s, 6H). Synthesis of [Cu(dmbpy)2]OTf2: 86 Copper(II) triflate (1.0325 g, 2.85 mmol) was dissolved in minimal dry acetonitrile. Meanwhile dmbpy (1.1255 g, 6.11 mmol) was dissolved in minimal dry dichloromethane. The dmbpy solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a dark green color. Reac(cid:415)on was s(cid:415)rred for 30 minutes then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(II)(dmbpy)2(OTf)2product was collected (1.7771 g, 85.4% yield). Synthesis of N,N(cid:3387)-Dibenzyl-N,N(cid:3387)-bis(6-methylpyridin-2-ylmethyl)ethylenediamine (dbmed): 2-(Chloromethyl)-6-methylpyridine hydrochloride (0.5300g, 3.0 mmol) and benzyltriethylammonium chloride (0.0121 g, 0.05 mmol) were added to 20 mL dichloromethane and s(cid:415)rred un(cid:415)l fully dissolved. N,N’-Dibenzylethylenediamine (0.35 mL, 1.48 mmol) was then added to the solu(cid:415)on, a(cid:332)er which a cloudy white solu(cid:415)on was formed. Meanwhile sodium hydroxide (20.1807 g, 500 mmol) was added to 20 mL of water which was then added to the dichloromethane solu(cid:415)on and s(cid:415)rred vigorously. A(cid:332)er 2 hours the organic solu(cid:415)on became clear and the aqueous solu(cid:415)on turned white. The solu(cid:415)on was then allowed to reflux overnight. A(cid:332)er refluxing addi(cid:415)onal water was added to the solu(cid:415)on to allow the layers to fully separate then the organic layer was then separated from the aqueous layer. The aqueous layer was then extracted with dichloromethane two (cid:415)mes, and all the organic layers were combined and dried with magnesium sulfate. Then the solvent was removed under low vacuum. The crude solid was then recrystallized in acetonitrile and dried to obtain the product (0.4270 g, 60.1% yield). 1H-NMR (CDCl3 500 MHz) δ 2.50 (s, 6H), 2.67 (s, 4H), 3.56 (s, 4H), 3.67 (s, 4H), 6.96 (d, 2H), 7.25 (m, 12H), 7.45 (t, 2H). Synthesis of [Cu(dbmed)]OTf: 87 Copper(I) tetrakisacetonitrile triflate (0.1123 g, 0.30 mmol) was dissolved in minimal dry acetonitrile. Meanwhile dbmed (0.1487 g, 0.33 mmol) was dissolved in minimal dry dichloromethane. The dbmed solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a bright yellow color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(I)dbmed product was collected (0.1876 g, 94.9% yield). 1H-NMR (CDCl3 500 MHz) δ 2.67 (s, 2H), 2.72 (s, 6H), 2.82 (s, 2H), 3.57 (s, 2H), 3.74 (s, 2H), 3.85 (s, 2H), 4.16 (s, 2H), 7.20 (m, 4H), 7.29 (m, 8H), 7.43 (d, 2H), 7.83 (t, 2H). Synthesis of [Cu(dbmed)]OTf2: Copper(II) triflate (0.808 g, 2.23 mmol) was dissolved in minimal dry acetonitrile. Meanwhile dbmed (0.1030 g, 2.29 mmol) was dissolved in minimal dry dichloromethane. The dbmed solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a deep blue color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether and filtered. The solvent was then removed via vacuum, and the Cu(II)dbmed product was collected (0.1596 g, 87.9% yield). Synthesis of [Cu(TBP)4]OTf: Copper(I) tetrakisacetonitrile triflate (0.2513 g, 0.67 mmol) was dissolved in minimal dry dichloromethane. 4-tert-butylpyridine (0.972 mL, 6.63 mmol) was then added to the solu(cid:415)on, turning the solu(cid:415)on a bright yellow color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether, forming pale yellow crystals, and the solid was filtered. The product was then dried via vacuum and was collected (0.4351 g, 86.% yield). 1H NMR (500 MHz, acetonitrile-d3) δ 8.47 (s, 2H), 7.44 (s, 2H) 1.33 (s, 9H). Synthesis of [Cu(TBP)4]OTf2: 88 Copper(II) triflate (0.4135 g, 1.14 mmol) was dissolved in minimal dry acetonitrile. TBP (1.14 mL, 8.00 mmol) was then added to the solu(cid:415)on, turning the solu(cid:415)on a deep purple color. Reac(cid:415)on was s(cid:415)rred for 2 hours then precipitated with dry ether and filtered. The reac(cid:415)on was then recrystallized using acetonitrile and ether. The product was then dried via vacuum and was collected (0.9043 g, 87.5% yield). Synthesis of [Cu(bpyPY4)]OTf: Copper(I) tetrakisacetonitrile triflate (0.1647 g, 0.43 mmol) was dissolved in minimal dry acetonitrile. Meanwhile 6,6(cid:3387)-bis(1,1-di(pyridin-2-yl)ethyl)-2,2(cid:3387)-bipyridine (bpyPY4) (0.2545 g, 0.49 mmol) was dissolved in minimal dry dichloromethane. The bpyPY4 solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a dark red color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(I)bpyPY4 product was collected (0.2440 g, 76.1% yield). 1H-NMR (CDCl3 500 MHz) δ 2.25 (s, 6H), 7.20 (s, 4H), 7.76 (s, 4H), 8.10 (d, 2H), 8.18 (t, 2H), 8.37 (d, 2H) Synthesis of [Cu(bpyPY4)]OTf2: Copper(II) triflate (0.4215 g, 1.16 mmol) was dissolved in minimal dry acetonitrile. Meanwhile BPYPY4 (0.6942 g, 1.33 mmol) was dissolved in minimal dry dichloromethane. The BPYPY4 solu(cid:415)on was added dropwise to the copper solu(cid:415)on, turning the solu(cid:415)on a pale blue color. Reac(cid:415)on was s(cid:415)rred overnight then precipitated with dry ether and filtered. The solvent was then removed via vacuum and the Cu(II)bpyPY4 product was collected (0.6571 g, 63.9% yield). 5.3 Results and Discussion While Dr. Rorabacher already confirmed the Nerns(cid:415)an rela(cid:415)onship between copper(II) stability constants and poten(cid:415)al, the rela(cid:415)onship was only inves(cid:415)gated in aqueous solu(cid:415)ons, whereas 89 most DSSC electrolytes use acetonitrile as the solvent. To discern the rela(cid:415)onship between the redox poten(cid:415)al and stability constants of copper complexes in acetonitrile, data from previous studies was analyzed, specifically concentra(cid:415)ng on the stability constants of these complexes in acetonitrile.98,99 A(cid:332)er all the collated data was plo(cid:425)ed, Figure 5.2A, a linear regression, with a fixed slope of -0.059, was applied, resul(cid:415)ng in the equa(cid:415)on y = -0.059x + 0.9002. This equa(cid:415)on was then used to simulate the stability constants of redox shu(cid:425)led u(cid:415)lized in DSSC by using their previous determined redox poten(cid:415)als, Figure 5.2B.38,39,41,94,95 Figure 5.2: A) a plot of published copper redox poten(cid:415)als versus the log of the stability constant of the copper(II) form of the redox complex found in acetonitrile.98,99 The solid line was fit while holding the slope at the Nerns(cid:415)an slope of -0.059, resul(cid:415)ng in an equa(cid:415)ons of y = -0.059x + 0.9002. B) A simulated plot where the stability constant of the copper(II) form of various copper complexes used in DSSCs was simulated using the equa(cid:415)on from plot A.38,39,41,94,95 The simulated plots reveal a pa(cid:425)ern: copper complexes that undergo ligand subs(cid:415)tu(cid:415)on display more posi(cid:415)ve poten(cid:415)als and subsequently exhibit lower stability constants compared to the [Cu(TBP)4]2+/+ redox shu(cid:425)le, while those complexes that do not undergo ligand subs(cid:415)tu(cid:415)on have more nega(cid:415)ve poten(cid:415)als and thus higher stability constants. This correla(cid:415)on underscores the role of redox poten(cid:415)al as a poten(cid:415)al tool for predic(cid:415)ng the suscep(cid:415)bility of copper complexes to ligand subs(cid:415)tu(cid:415)on. 90 The valida(cid:415)on of the simulated data was pursued by examining four copper complexes u(cid:415)lized in DSSCs, based on their relevance to the study's goals and their dis(cid:415)nc(cid:415)ve redox poten(cid:415)als. The selected complexes were [Cu(TBP)4]2+/+, [Cu(dmbpy)2]2+/+, [Cu(dbmed)]2+/+, and [Cu(bpyPY4)]2+/+. The [Cu(TBP)4]2+/+ was chosen as a reference point for compara(cid:415)ve evalua(cid:415)on, forming the baseline against which other complexes were assessed. The selec(cid:415)on of [Cu(dmbpy)2]2+/+ was mo(cid:415)vated by its posi(cid:415)ve redox poten(cid:415)al, placing it as one of the copper redox shu(cid:425)les with the highest predicted stability constant among those commonly employed in DSSCs. Conversely, [Cu(bpyPY4)]2+/+ was singled out for its nega(cid:415)ve redox poten(cid:415)al, reflec(cid:415)ng its status among copper redox shu(cid:425)les with the lowest poten(cid:415)als integrated into DSSCs. The selec(cid:415)on of [Cu(dbmed)]2+/+ was due to its redox poten(cid:415)al closely resembling that of [Cu(TBP)4]2+/+, which enabled a targeted explora(cid:415)on of ligand subs(cid:415)tu(cid:415)on behavior while upholding a uniform poten(cid:415)al. For [Cu(dmbpy)2]2+/+, [Cu(dbmed)]2+/+, and [Cu(bpyPY4)]2+/+ the stability constants were determined via the applica(cid:415)on of square wave voltammetry. Due to the high stability of copper(II) complexes in acetonitrile, making them difficult to measure, the stability constants of the copper(I) form were experimentally determined. Copper(I) complexes have been shown to have a significantly lower stability constant in acetonitrile than in water. This has been a(cid:425)ributed to the significantly higher binding energy of copper(I) to acetonitrile than that of water, causing about a 106 decrease in stability constants in acetonitrile rela(cid:415)ve to their aqueous values.99,100 Square wave voltammetry was u(cid:415)lized to determine the stability constant of the copper(I) form of the complexes. When using square wave voltammetry, assuming the sweep rate is held constant, the copper(I) oxida(cid:415)on peak can be directly related to the [CuIL] following the equa(cid:415)on 91 (cid:3117) 𝑖 = 𝑛𝐹𝐴𝐶(𝜋𝐷𝜎) (cid:3118)𝛸(𝜎𝑡) (5.1) where i is the peak current, n is the number of electrons transferred, F is the Faraday constant, C is the bulk concentra(cid:415)on of CuIL, D is the diffusion coefficient, σ = (nFν/RT), ν is the sweep rate, R is the gas constant, T is the absolute temperature, and Χ(σt) is the normalized current. The concentra(cid:415)on at each measurement was determined using the propor(cid:415)onality of the peak current, i, for the measurement compared to the peak current when all copper is bound to ligand, i∞, usually at 120% ligand added, following the equa(cid:415)on [CuL] = CCu(i/i∞). From the combina(cid:415)on of defini(cid:415)on of equilibrium constant between copper and a ligand 𝐾(cid:3004)(cid:3048)(cid:3013) = [(cid:3004)(cid:3048)(cid:3013)] [(cid:3004)(cid:3048)][(cid:3013)] (5.2) and the known concentra(cid:415)ons of copper, CCu, and ligand, CL, 𝐶(cid:3004)(cid:3048) = [𝐶𝑢] + [𝐶𝑢𝐿] (5.3) 𝐶(cid:3013) = [𝐿] + [𝐶𝑢𝐿] (5.4) a new equa(cid:415)on can be derived [𝐶𝑢𝐿] = (cid:3012)(cid:3252)(cid:3296)(cid:3261)(cid:3004)(cid:3261)((cid:3004)(cid:3252)(cid:3296)(cid:2879)[(cid:3004)(cid:3048)(cid:3013)]) (cid:2869)(cid:2878)(cid:3012)(cid:3252)(cid:3296)(cid:3261)((cid:3004)(cid:3252)(cid:3296)(cid:2879)[(cid:3004)(cid:3048)(cid:3013)]) (5.5) then subs(cid:415)tu(cid:415)ng equa(cid:415)on 5.5 into equa(cid:415)on 5.1 yields the equa(cid:415)on (cid:3004)(cid:3252)(cid:3296) (cid:3036) = (cid:2869) (cid:3026) + (cid:2869) (cid:3026)(cid:3012)(cid:3252)(cid:3296)(cid:3261)((cid:3004)(cid:3261)(cid:2879)[(cid:3004)(cid:3048)(cid:3013)]) (5.6) where CCu is the total concentra(cid:415)on of copper in solu(cid:415)on, CL is the total concentra(cid:415)on of ligand in solu(cid:415)on, i is the peak current, KCuL is the stability constant of the copper(I) form of the complex, and Y is the product of several electrochemical constants. By plo(cid:427)ng CCu/i by 1/(CL – [CuL]) and taking a linear regression of the plot, you obtain a slope of 1/YKCuL and an intercept of 1/Y. By dividing the intercept by the slope, one is le(cid:332) with KCuL. 92 To get the stability constant of the copper(II) form, the Nerst equa(cid:415)on was used 𝐸(cid:3004)(cid:3048)(cid:3258)(cid:3258)(cid:3013) (cid:3033) = 𝐸(cid:3004)(cid:3048)((cid:3010)/(cid:3010)(cid:3010))(cid:3046)(cid:3042)(cid:3039)(cid:3049) (cid:3033) − (cid:2870).(cid:2871)(cid:2868)(cid:2871)(cid:3019)(cid:3021) (cid:3007) 𝑙𝑜𝑔 (cid:3012) (cid:3252)(cid:3296)(cid:3258)(cid:3258)(cid:3261) (cid:3012)(cid:3252)(cid:3296)(cid:3258)(cid:3261) (5.7) where ECu(I/II)solv f is the formal poten(cid:415)al of the solvated copper(I/II), [Cu(ACN)4]2+/+, redox couple, which was found to be 0.601 V vs ferrocene99. The log of stability constants of the copper(II) for of the complexes were found to be 9.05, 15.02, and 21.53 respec(cid:415)vely for the [Cu(dmbpy)2]2+/+, [Cu(dbmed)]2+/+, and [Cu(bpyPY4)]2+/+ complexes. These results match well with the log of the predicted stability constants found previously, 9.49, 15.34, and 22.55 respec(cid:415)vely. However, when the [Cu(TBP)4]2+/+ complex was a(cid:425)empted to be measured via square wave voltammetry it was found that the electron kine(cid:415)cs of the complex were too slow at the sweep rate used for a redox wave to show, so a different technique was used to determine the copper(II) stability constant. 93 Figure 5.13: The square wave voltammetry plots for 2.0 mM of A) [Cu(dmbpy)2]OTf2, blue, B) [Cu(dbmed)]OTf2, orange, C) [Cu(bpyPY4)]OTf2, green, and [Cu(ACN)4]OTf2, in anhydrous acetonitrile containing 0.1 M LiOTf. D) The UV-Vis spectra of 2.0 mM [Cu(dbmed)]OTf2 (cid:415)trated increasing equivalents of TBP. For all spectra as the concentra(cid:415)on of ligand increases the darker the spectra. For [Cu(TBP)4]2+/+ the redox couple was electrochemically irreversible at the sweep rates used for the square wave measurement, so instead UV-Vis was used to determine the rela(cid:415)ve stability constant between [Cu(TBP)4]2+/+ and [Cu(dbmed)]2+/+. Since the difference of the predicted stability constants for the [Cu(dbmed)]OTf2 and the [Cu(TBP)4]OTf2 complexes is small, only 10.6, UV-Vis spectroscopy can be used to determine the equillibrium constant for the reac(cid:415)on [Cu(dbmed)](OTf)2 + 4 TBP [Cu(TBP)4](OTf)2 + dbmed (Scheme 5.1) The wavelength of 1100 nm was used due to the minimal contribu(cid:415)on from the [Cu(TBP)4] complex, free TBP, and free dbmed to the absorbance at this wavelength. Due to this we can use the Beer-Lambert law 𝐴 = 𝜀(cid:3004)(cid:3048)(cid:3013)𝑏[𝐶𝑢𝐿] (5.8) 94 where εCuL is the ex(cid:415)nc(cid:415)on coefficient of the [Cu(dbmed)]OTf2 complex, b is the path length in cm, and [CuL] is the concentra(cid:415)on of the [Cu(dbmed)]OTf2 complex in solu(cid:415)on. By using the defini(cid:415)on of an equilibrium constant 𝐾(cid:3004)(cid:3048)(cid:3013) = [(cid:3004)(cid:3048)((cid:3021)(cid:3003)(cid:3017))(cid:3120)][(cid:3031)(cid:3029)(cid:3040)(cid:3032)(cid:3031)] [(cid:3004)(cid:3048)((cid:3031)(cid:3029)(cid:3040)(cid:3032)(cid:3031))][(cid:3021)(cid:3003)(cid:3017)](cid:3120) (5.9) and the defini(cid:415)ons of the analy(cid:415)cal concentra(cid:415)ons of copper(II), CCu, and dbmed, Cdbmed, and TBP, CTBP, in solu(cid:415)on 𝐶(cid:3004)(cid:3048) = [𝐶𝑢(𝑇𝐵𝑃)(cid:2872)] + [𝐶𝑢(𝑑𝑏𝑚𝑒𝑑)] (5.10) 𝐶(cid:3031)(cid:3029)(cid:3040)(cid:3032)(cid:3031) = [𝑑𝑏𝑚𝑒𝑑] + [𝐶𝑢(𝑑𝑏𝑚𝑒𝑑)] (5.11) 𝐶(cid:3021)(cid:3003)(cid:3017) = [𝑇𝐵𝑃](cid:2872) + [𝐶𝑢(𝑇𝐵𝑃)(cid:2872)] (5.12) under the assump(cid:415)on that all the copper(II) is complexed with a ligand, the equilibrium coefficient calcula(cid:415)on can be rearranged to be 𝐾(cid:3032)(cid:3044) = ((cid:3004)(cid:3252)(cid:3296)(cid:2879)[(cid:3004)(cid:3048)((cid:3031)(cid:3029)(cid:3040)(cid:3032)(cid:3031))])((cid:3004)(cid:3279)(cid:3277)(cid:3288)(cid:3280)(cid:3279)(cid:2879)[(cid:3004)(cid:3048)((cid:3031)(cid:3029)(cid:3040)(cid:3032)(cid:3031))]) [(cid:3004)(cid:3048)((cid:3031)(cid:3029)(cid:3040)(cid:3032)(cid:3031))]((cid:3004)(cid:3269)(cid:3251)(cid:3265)(cid:2879)(cid:3004)(cid:3252)(cid:3296)(cid:2879)[(cid:3004)(cid:3048)((cid:3031)(cid:3029)(cid:3040)(cid:3032)(cid:3031))]) (5.13) Using this rela(cid:415)onship the equilibrium constant for the rela(cid:415)onship between [Cu(dbmed)]OTf2 and [Cu(TBP)4]OTf2 was determined to be 1.02 (± 0.27) which, when referenced against the stability constant for [Cu(dbmed)]OTf2, gave a stability constant of 16.74 for [Cu(TBP)4]OTf2, which is close to the predicted stability constant of 16.21. 95 Table 5.8: Simulated and experimental copper(II) stability constants and the percent error in between the measurements. Experimental Cu(II) stability constant Simulated Cu(II) stability constant Percent Error (%) Cu(dmpby)2 9.05 9.49 Cu(dbmed) 15.02 Cu(bpyPY4) 21.53 Cu(TBP)4 16.68 15.34 22.55 16.21 4.92 2.01 4.73 2.80 The stability constants determined for all the copper complexes are listed in Table 5.1 along with the predicted stability constant and the percent error for each copper complex. The comparison between these experimental results and the simulated data revealed that all the complexes displaying devia(cid:415)ons of less than 5% from the simulated outcomes. This agreement to the predic(cid:415)ve model denotes the poten(cid:415)al of the redox poten(cid:415)al-stability constant rela(cid:415)onship to be used as a predic(cid:415)ve tool for an(cid:415)cipa(cid:415)ng ligand subs(cid:415)tu(cid:415)on behavior. 96 V / e n e c o r r e F s v l a i t n e t o P -0.6 -0.4 -0.2 0 0.2 0.4 0.6 Cu Complexes potential vs Stability constants in ACN y = -0.059x + 0.900 y = -0.060x + 0.900 R² = 0.987 Cu(bpyPY4) Cu(bpyPY4) Cu(PY5) Cu(TBP)4 Cu(dbmed) Cu(TBP)4 Cu(dbmed) Cu(dmbppy)2 Cu(tmbpy)2 Cu(dmbpy)2 Cu(ACN)4 5 10 15 log KCu(II)L 20 25 Figure 5.4: A plot of copper redox poten(cid:415)als versus the log of the simulated, orange, and experimental, blue, stability constant of the copper(II) form of the redox complex found in acetonitrile. The linear fit equa(cid:415)on for the experimental results was found to be y = -0.060x + 0.900. Addi(cid:415)onally, to deepen insights, the poten(cid:415)al of each redox shu(cid:425)le was plo(cid:425)ed against the measured stability constant, enabling a subsequent linear regression analysis. This analysis yielded a slope of -0.060, closely mirroring the an(cid:415)cipated fundamental value of -0.059. Although it is important to note that this rela(cid:415)onship should only hold true as long as the stability constants of the copper(I) form of the redox couple have similar values. This adherence to theore(cid:415)cal expecta(cid:415)ons further underscores the legi(cid:415)macy of the correla(cid:415)on, offering a compelling avenue for leveraging redox poten(cid:415)al to predict ligand subs(cid:415)tu(cid:415)on behavior in copper complexes. 97 5.4 Conclusions This study capitalizes on Dr. Rorabacher's work, revealing a compelling connec(cid:415)on between redox poten(cid:415)al and stability constants in copper complexes. This correla(cid:415)on enables a predic(cid:415)ve model of poten(cid:415)al-stability constant rela(cid:415)onships. Assuming the stability constants of the copper(I) form are consistent, this model facilitates the accurate predic(cid:415)on of copper complexes. Specifically, it helps dis(cid:415)nguish between those that undergo ligand subs(cid:415)tu(cid:415)on and those that do not. Valida(cid:415)on with four relevant copper complexes shows the predic(cid:415)ve model's accuracy. Given the knowledge that ligand subs(cid:415)tu(cid:415)on can be beneficial for DSSCs, enhancing efficiencies could be achieved by enabling ligand subs(cid:415)tu(cid:415)on with a base that induces a smaller nega(cid:415)ve shi(cid:332) in the solu(cid:415)on poten(cid:415)al.38 This predic(cid:415)ve method could allow for the selec(cid:415)on of targeted redox shu(cid:425)le/base combina(cid:415)ons, aiming to maximize the efficiency of DSSC devices. This work paves the way for informed design and op(cid:415)miza(cid:415)on of copper-based redox shu(cid:425)les in dye-sensi(cid:415)zed solar cells, offering new dimensions for enhancing their performance. 98 WORKS CITED (1) Friedlingstein, P.; O’Sullivan, M.; Jones, M. W.; Andrew, R. M.; Gregor, L.; Hauck, J.; Le Quéré, C.; Luijkx, I. T.; Olsen, A.; Peters, G. P.; Peters, W.; Pongratz, J.; Schwingshackl, C.; Sitch, S.; Canadell, J. G.; Ciais, P.; Jackson, R. B.; Alin, S. R.; Alkama, R.; Arneth, A.; Arora, V. K.; Bates, N. R.; Becker, M.; Bellouin, N.; Bi(cid:427)g, H. C.; Bopp, L.; Chevallier, F.; Chini, L. P.; Cronin, M.; Evans, W.; Falk, S.; Feely, R. A.; Gasser, T.; Gehlen, M.; Gkritzalis, T.; Gloege, L.; Grassi, G.; Gruber, N.; Gürses, Ö.; Harris, I.; Hefner, M.; Houghton, R. A.; Hur(cid:425), G. C.; Iida, Y.; Ilyina, T.; Jain, A. K.; Jersild, A.; Kadono, K.; Kato, E.; Kennedy, D.; Klein Goldewijk, K.; Knauer, J.; Korsbakken, J. I.; Landschützer, P.; Lefèvre, N.; Lindsay, K.; Liu, J.; Liu, Z.; Marland, G.; Mayot, N.; McGrath, M. J.; Metzl, N.; Monacci, N. M.; Munro, D. R.; Nakaoka, S.-I.; Niwa, Y.; O’Brien, K.; Ono, T.; Palmer, P. I.; Pan, N.; Pierrot, D.; Pocock, K.; Poulter, B.; Resplandy, L.; Robertson, E.; Rödenbeck, C.; Rodriguez, C.; Rosan, T. M.; Schwinger, J.; Séférian, R.; Shutler, J. D.; Skjelvan, I.; Steinhoff, T.; Sun, Q.; Su(cid:425)on, A. J.; Sweeney, C.; Takao, S.; Tanhua, T.; Tans, P. P.; Tian, X.; Tian, H.; Tilbrook, B.; Tsujino, H.; Tubiello, F.; van der Werf, G. R.; Walker, A. P.; Wanninkhof, R.; Whitehead, C.; Willstrand Wranne, A.; Wright, R.; Yuan, W.; Yue, C.; Yue, X.; Zaehle, S.; Zeng, J.; Zheng, B. Global Carbon Budget 2022. Earth Syst Sci Data 2022, 14 (11), 4811–4900. h(cid:425)ps://doi.org/10.5194/essd-14-4811-2022. (2) Interna(cid:415)onal Energy Agency. CO2 Emissions in 2022; 2023. www.iea.org. (3) Hausfather, Z.; Peters, G. P. Emissions – the ‘Business as Usual’ Story Is Misleading. Nature 2020, 577 (7792), 618–620. h(cid:425)ps://doi.org/10.1038/d41586-020-00177-3. (4) Center for Sustainable Systems, U. of M. Photovoltaic Energy Factsheet; 2023. (5) Interna(cid:415)onal Energy Agency. Solar PV; 2023. (6) Gueymard, C. A.; Myers, D.; Emery, K. Proposed Reference Irradiance Spectra for Solar Energy Systems Tes(cid:415)ng. Solar Energy 2002, 73 (6), 443–467. h(cid:425)ps://doi.org/10.1016/S0038-092X(03)00005-7. (7) U.S. DOE. The SunShot Ini(cid:415)a(cid:415)ve. (8) (9) Jung, H. S.; Lee, J.-K. Dye Sensi(cid:415)zed Solar Cells for Economically Viable Photovoltaic Systems. J Phys Chem Le(cid:425) 2013, 4 (10), 1682–1693. h(cid:425)ps://doi.org/10.1021/jz400112n. Becquerel, E. Mémoire Sur Les Effets Électriques Produits Sous l’influence Des Rayons Solaires. Comptes Rendus 1839, 9, 561–567. (10) Moser, J. No(cid:415)z �ber Verst�rkung Photoelektrischer Str�me Durch Op(cid:415)sche Sensibilisirung. Monatshe(cid:332)e f�r Chemie 1887, 8 (1), 373–373. h(cid:425)ps://doi.org/10.1007/BF01510059. 99 (11) Gerischer, H.; Tributsch, H. Elektrochemische Untersuchungen Zur Spektralen Sensibilisierung von ZnO-Einkristallen. Berichte der Bunsengesellscha(cid:332) für physikalische Chemie 1968, 72 (3), 437–445. h(cid:425)ps://doi.org/10.1002/bbpc.196800013. (12) ANDERSON, S.; CONSTABLE, E. C.; DARE-EDWARDS, M. P.; GOODENOUGH, J. B.; HAMNETT, A.; SEDDON, K. R.; WRIGHT, R. D. Chemical Modifica(cid:415)on of a Titanium (IV) Oxide Electrode to Give Stable Dye Sensi(cid:415)sa(cid:415)on without a Supersensi(cid:415)ser. Nature 1979, 280 (5723), 571–573. h(cid:425)ps://doi.org/10.1038/280571a0. (13) Matsumura, M.; Matsudaira, S.; Tsubomura, H.; Takata, M.; Yanagida, H. Dye Sensi(cid:415)za(cid:415)on and Surface Structures of Semiconductor Electrodes. Industrial & Engineering Chemistry Product Research and Development 1980, 19 (3), 415–421. h(cid:425)ps://doi.org/10.1021/i360075a025. (14) Vlachopoulos, Nick.; Liska, Paul.; Augustynski, Jan.; Graetzel, Michael. Very Efficient Visible Light Energy Harves(cid:415)ng and Conversion by Spectral Sensi(cid:415)za(cid:415)on of High Surface Area Polycrystalline Titanium Dioxide Films. J Am Chem Soc 1988, 110 (4), 1216–1220. h(cid:425)ps://doi.org/10.1021/ja00212a033. (15) Nazeeruddin, M. K.; Kay, A.; Rodicio, I.; Humphry-Baker, R.; Mueller, E.; Liska, P.; Vlachopoulos, N.; Graetzel, M. Conversion of Light to Electricity by Cis-X2bis(2,2’- Bipyridyl-4,4’-Dicarboxylate)Ruthenium(II) Charge-Transfer Sensi(cid:415)zers (X = Cl-, Br-, I-, CN-, and SCN-) on Nanocrystalline Titanium Dioxide Electrodes. J Am Chem Soc 1993, 115 (14), 6382–6390. h(cid:425)ps://doi.org/10.1021/ja00067a063. (16) Feldt, S. M.; Gibson, E. A.; Gabrielsson, E.; Sun, L.; Boschloo, G.; Hagfeldt, A. Design of Organic Dyes and Cobalt Polypyridine Redox Mediators for High-Efficiency Dye-Sensi(cid:415)zed Solar Cells. J Am Chem Soc 2010, 132 (46), 16714–16724. h(cid:425)ps://doi.org/10.1021/ja1088869. (17) Daeneke, T.; Kwon, T.-H.; Holmes, A. B.; Duffy, N. W.; Bach, U.; Spiccia, L. High-Efficiency Dye-Sensi(cid:415)zed Solar Cells with Ferrocene-Based Electrolytes. Nat Chem 2011, 3 (3), 211– 215. h(cid:425)ps://doi.org/10.1038/nchem.966. (18) Bai, Y.; Yu, Q.; Cai, N.; Wang, Y.; Zhang, M.; Wang, P. High-Efficiency Organic Dye- Sensi(cid:415)zed Mesoscopic Solar Cells with a Copper Redox Shu(cid:425)le. Chemical Communica(cid:415)ons 2011, 47 (15), 4376. h(cid:425)ps://doi.org/10.1039/c1cc10454c. (19) Cao, Y.; Liu, Y.; Zakeeruddin, S. M.; Hagfeldt, A.; Grätzel, M. Direct Contact of Selec(cid:415)ve Charge Extrac(cid:415)on Layers Enables High-Efficiency Molecular Photovoltaics. Joule 2018, 2 (6), 1108–1117. h(cid:425)ps://doi.org/10.1016/j.joule.2018.03.017. (20) Ren, Y.; Zhang, D.; Suo, J.; Cao, Y.; Eickemeyer, F. T.; Vlachopoulos, N.; Zakeeruddin, S. M.; Hagfeldt, A.; Grätzel, M. Hydroxamic Acid Pre-Adsorp(cid:415)on Raises the Efficiency of 100 Cosensi(cid:415)zed Solar Cells. Nature 2023, 613 (7942), 60–65. h(cid:425)ps://doi.org/10.1038/s41586-022-05460-z. (21) Ondersma, J. W.; Hamann, T. W. Conduc(cid:415)on Band Energy Determina(cid:415)on by Variable Temperature Spectroelectrochemistry. Energy Environ Sci 2012, 5 (11), 9476. h(cid:425)ps://doi.org/10.1039/c2ee22926a. (22) Boschloo, G.; Hagfeldt, A. Characteris(cid:415)cs of the Iodide/Triiodide Redox Mediator in Dye- Sensi(cid:415)zed Solar Cells. Acc Chem Res 2009, 42 (11), 1819–1826. h(cid:425)ps://doi.org/10.1021/ar900138m. (23) Mathew, S.; Yella, A.; Gao, P.; Humphry-Baker, R.; Curchod, B. F. E.; Ashari-Astani, N.; Tavernelli, I.; Rothlisberger, U.; Nazeeruddin, Md. K.; Grätzel, M. Dye-Sensi(cid:415)zed Solar Cells with 13% Efficiency Achieved through the Molecular Engineering of Porphyrin Sensi(cid:415)zers. Nat Chem 2014, 6 (3), 242–247. h(cid:425)ps://doi.org/10.1038/nchem.1861. (24) Shi, J.; Zhang, Y.; Luo, B.; Chen, L. Synthesis and Structural Characteriza(cid:415)on of the Copper(II) Complex with N, N(cid:3387)-Bis-(1-Benzimidazolylethyl)Ethylenediamine. Inorganica Chim Acta 1989, 162 (1), 29–31. h(cid:425)ps://doi.org/10.1016/S0020-1693(00)83115-1. (25) Randall, D. W.; Gamelin, D. R.; LaCroix, L. B.; Solomon, E. I. Electronic Structure Contribu(cid:415)ons to Electron Transfer in Blue Cu and CuA. JBIC Journal of Biological Inorganic Chemistry 2000, 5 (1), 16–29. h(cid:425)ps://doi.org/10.1007/s007750050003. (26) Ha(cid:425)ori, S.; Wada, Y.; Yanagida, S.; Fukuzumi, S. Blue Copper Model Complexes with Distorted Tetragonal Geometry Ac(cid:415)ng as Effec(cid:415)ve Electron-Transfer Mediators in Dye- Sensi(cid:415)zed Solar Cells. J Am Chem Soc 2005, 127 (26), 9648–9654. h(cid:425)ps://doi.org/10.1021/ja0506814. (27) Bai, Y.; Yu, Q.; Cai, N.; Wang, Y.; Zhang, M.; Wang, P. High-Efficiency Organic Dye- Sensi(cid:415)zed Mesoscopic Solar Cells with a Copper Redox Shu(cid:425)le. Chemical Communica(cid:415)ons 2011, 47 (15), 4376. h(cid:425)ps://doi.org/10.1039/c1cc10454c. (28) Colombo, A.; Dragone(cid:427), C.; Magni, M.; Roberto, D.; Demar(cid:415)n, F.; Caramori, S.; Bignozzi, C. A. Efficient Copper Mediators Based on Bulky Asymmetric Phenanthrolines for DSSCs. ACS Appl Mater Interfaces 2014, 6 (16), 13945–13955. h(cid:425)ps://doi.org/10.1021/am503306f. (29) Cong, J.; Kinschel, D.; Daniel, Q.; Safdari, M.; Gabrielsson, E.; Chen, H.; Svensson, P. H.; Sun, L.; Kloo, L. Bis(1,1-Bis(2-Pyridyl)Ethane)Copper( i / ii ) as an Efficient Redox Couple for Liquid Dye-Sensi(cid:415)zed Solar Cells. J Mater Chem A Mater 2016, 4 (38), 14550–14554. h(cid:425)ps://doi.org/10.1039/C6TA06782D. 101 (30) Freitag, M.; Giordano, F.; Yang, W.; Pazoki, M.; Hao, Y.; Zietz, B.; Grätzel, M.; Hagfeldt, A.; Boschloo, G. Copper Phenanthroline as a Fast and High-Performance Redox Mediator for Dye-Sensi(cid:415)zed Solar Cells. The Journal of Physical Chemistry C 2016, 120 (18), 9595–9603. h(cid:425)ps://doi.org/10.1021/acs.jpcc.6b01658. (31) Saygili, Y.; Söderberg, M.; Pellet, N.; Giordano, F.; Cao, Y.; Muñoz-García, A. B.; Zakeeruddin, S. M.; Vlachopoulos, N.; Pavone, M.; Boschloo, G.; Kavan, L.; Moser, J.-E.; Grätzel, M.; Hagfeldt, A.; Freitag, M. Copper Bipyridyl Redox Mediators for Dye-Sensi(cid:415)zed Solar Cells with High Photovoltage. J Am Chem Soc 2016, 138 (45), 15087–15096. h(cid:425)ps://doi.org/10.1021/jacs.6b10721. (32) Hu, M.; Shen, J.; Yu, Z.; Liao, R.-Z.; Gurzadyan, G. G.; Yang, X.; Hagfeldt, A.; Wang, M.; Sun, L. Efficient and Stable Dye-Sensi(cid:415)zed Solar Cells Based on a Tetradentate Copper(II/I) Redox Mediator. ACS Appl Mater Interfaces 2018, 10 (36), 30409–30416. h(cid:425)ps://doi.org/10.1021/acsami.8b10182. (33) Long, H.; Zhou, D.; Zhang, M.; Peng, C.; Uchida, S.; Wang, P. Probing Dye-Correlated Interplay of Energe(cid:415)cs and Kine(cid:415)cs in Mesoscopic Titania Solar Cells with 4- Tert - Butylpyridine. The Journal of Physical Chemistry C 2011, 115 (29), 14408–14414. h(cid:425)ps://doi.org/10.1021/jp202826m. (34) XIONG, B.; ZHOU, B.; ZHU, Z.; GAO, T.; LI, L.; CAI, J.; CAI, W. Adsorp(cid:415)on of 4 -tert- Butylpyridine on TiO 2 Surface in Dye-Sensi(cid:415)zed Solar Cells. Chin J Chem 2008, 26 (1), 70– 76. h(cid:425)ps://doi.org/10.1002/cjoc.200890040. (35) Kashif, M. K.; Axelson, J. C.; Duffy, N. W.; Forsyth, C. M.; Chang, C. J.; Long, J. R.; Spiccia, L.; Bach, U. A New Direc(cid:415)on in Dye-Sensi(cid:415)zed Solar Cells Redox Mediator Development: In Situ Fine-Tuning of the Cobalt(II)/(III) Redox Poten(cid:415)al through Lewis Base Interac(cid:415)ons. J Am Chem Soc 2012, 134 (40), 16646–16653. h(cid:425)ps://doi.org/10.1021/ja305897k. (36) Nakade, S.; Kanzaki, T.; Kubo, W.; Kitamura, T.; Wada, Y.; Yanagida, S. Role of Electrolytes on Charge Recombina(cid:415)on in Dye-Sensi(cid:415)zed TiO 2 Solar Cell (1): The Case of Solar Cells Using the I - /I 3 - Redox Couple. J Phys Chem B 2005, 109 (8), 3480–3487. h(cid:425)ps://doi.org/10.1021/jp0460036. (37) Kusama, H.; Konishi, Y.; Sugihara, H.; Arakawa, H. Influence of Alkylpyridine Addi(cid:415)ves in Electrolyte Solu(cid:415)on on the Performance of Dye-Sensi(cid:415)zed Solar Cell. Solar Energy Materials and Solar Cells 2003, 80 (2), 167–179. h(cid:425)ps://doi.org/10.1016/j.solmat.2003.08.001. (38) Wang, Y.; Hamann, T. W. Improved Performance Induced by in Situ Ligand Exchange Reac(cid:415)ons of Copper Bipyridyl Redox Couples in Dye-Sensi(cid:415)zed Solar Cells. Chemical Communica(cid:415)ons 2018, 54 (87), 12361–12364. h(cid:425)ps://doi.org/10.1039/C8CC07191H. 102 (39) Fürer, S. O.; Milhuisen, R. A.; Kashif, M. K.; Raga, S. R.; Acharya, S. S.; Forsyth, C.; Liu, M.; Frazer, L.; Duffy, N. W.; Ohlin, C. A.; Funston, A. M.; Tachibana, Y.; Bach, U. The Performance-Determining Role of Lewis Bases in Dye-Sensi(cid:415)zed Solar Cells Employing Copper-Bisphenanthroline Redox Mediators. Adv Energy Mater 2020, 10 (37). h(cid:425)ps://doi.org/10.1002/aenm.202002067. (40) Hu, M.; Shen, J.; Yu, Z.; Liao, R.-Z.; Gurzadyan, G. G.; Yang, X.; Hagfeldt, A.; Wang, M.; Sun, L. Efficient and Stable Dye-Sensi(cid:415)zed Solar Cells Based on a Tetradentate Copper(II/I) Redox Mediator. ACS Appl Mater Interfaces 2018, 10 (36), 30409–30416. h(cid:425)ps://doi.org/10.1021/acsami.8b10182. (41) Rui, H.; Shen, J.; Yu, Z.; Li, L.; Han, H.; Sun, L. Stable Dye-Sensi(cid:415)zed Solar Cells Based on Copper(II/I) Redox Mediators Bearing a Pentadentate Ligand. Angewandte Chemie Interna(cid:415)onal Edi(cid:415)on 2021, 60 (29), 16156–16163. h(cid:425)ps://doi.org/10.1002/anie.202104563. (42) Michaels, H.; Benesperi, I.; Edvinsson, T.; Muñoz-Garcia, A.; Pavone, M.; Boschloo, G.; Freitag, M. Copper Complexes with Tetradentate Ligands for Enhanced Charge Transport in Dye-Sensi(cid:415)zed Solar Cells. Inorganics (Basel) 2018, 6 (2), 53. h(cid:425)ps://doi.org/10.3390/inorganics6020053. (43) Jeffrey D. Sachs. The Road to Clean Energy Starts Here. Sci Am 2007. (44) Bierbaum, R. M.; Matson, P. A. Energy in the Context of Sustainability. Daedalus 2013, 142 (1), 146–161. h(cid:425)ps://doi.org/10.1162/DAED_a_00191. (45) El Chaar, L.; lamont, L. A.; El Zein, N. Review of Photovoltaic Technologies. Renewable and Sustainable Energy Reviews 2011, 15 (5), 2165–2175. h(cid:425)ps://doi.org/10.1016/j.rser.2011.01.004. (46) Parida, B.; Iniyan, S.; Goic, R. A Review of Solar Photovoltaic Technologies. Renewable and Sustainable Energy Reviews 2011, 15 (3), 1625–1636. h(cid:425)ps://doi.org/10.1016/j.rser.2010.11.032. (47) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pe(cid:425)ersson, H. Dye-Sensi(cid:415)zed Solar Cells. Chem Rev 2010, 110 (11), 6595–6663. h(cid:425)ps://doi.org/10.1021/cr900356p. (48) Liu, Q.; Jiang, Y.; Jin, K.; Qin, J.; Xu, J.; Li, W.; Xiong, J.; Liu, J.; Xiao, Z.; Sun, K.; Yang, S.; Zhang, X.; Ding, L. 18% Efficiency Organic Solar Cells. Sci Bull (Beijing) 2020, 65 (4), 272– 275. h(cid:425)ps://doi.org/10.1016/j.scib.2020.01.001. (49) Lu, M. N.; Su, T.; Pylnev, M.; Long, Y.; Wu, T.; Tsai, M.; Wei, T. Stepwise Op(cid:415)mizing Photovoltaic Performance of Dye-sensi(cid:415)zed Cells Made under 50-lux Dim Light. Progress 103 in Photovoltaics: Research and Applica(cid:415)ons 2021, 29 (5), 533–545. h(cid:425)ps://doi.org/10.1002/pip.3406. (50) Cao, Y.; Liu, Y.; Zakeeruddin, S. M.; Hagfeldt, A.; Grätzel, M. Direct Contact of Selec(cid:415)ve Charge Extrac(cid:415)on Layers Enables High-Efficiency Molecular Photovoltaics. Joule 2018, 2 (6), 1108–1117. h(cid:425)ps://doi.org/10.1016/j.joule.2018.03.017. (51) Freitag, M.; Teuscher, J.; Saygili, Y.; Zhang, X.; Giordano, F.; Liska, P.; Hua, J.; Zakeeruddin, S. M.; Moser, J.-E.; Grätzel, M.; Hagfeldt, A. Dye-Sensi(cid:415)zed Solar Cells for Efficient Power Genera(cid:415)on under Ambient Ligh(cid:415)ng. Nat Photonics 2017, 11 (6), 372–378. h(cid:425)ps://doi.org/10.1038/nphoton.2017.60. (52) Kakiage, K.; Aoyama, Y.; Yano, T.; Oya, K.; Fujisawa, J.; Hanaya, M. Highly-Efficient Dye- Sensi(cid:415)zed Solar Cells with Collabora(cid:415)ve Sensi(cid:415)za(cid:415)on by Silyl-Anchor and Carboxy-Anchor Dyes. Chemical Communica(cid:415)ons 2015, 51 (88), 15894–15897. h(cid:425)ps://doi.org/10.1039/C5CC06759F. (53) Grätzel, M. Dye-Sensi(cid:415)zed Solar Cells. Journal of Photochemistry and Photobiology C: Photochemistry Reviews 2003, 4 (2), 145–153. h(cid:425)ps://doi.org/10.1016/S1389- 5567(03)00026-1. (54) Sapp, S. A.; Ellio(cid:425), C. M.; Contado, C.; Caramori, S.; Bignozzi, C. A. Subs(cid:415)tuted Polypyridine Complexes of Cobalt(II/III) as Efficient Electron-Transfer Mediators in Dye- Sensi(cid:415)zed Solar Cells. J Am Chem Soc 2002, 124 (37), 11215–11222. h(cid:425)ps://doi.org/10.1021/ja027355y. (55) Nusbaumer, H.; Zakeeruddin, S. M.; Moser, J.; Grätzel, M. An Alterna(cid:415)ve Efficient Redox Couple for the Dye-Sensi(cid:415)zed Solar Cell System. Chemistry – A European Journal 2003, 9 (16), 3756–3763. h(cid:425)ps://doi.org/10.1002/chem.200204577. (56) Yum, J.-H.; Baranoff, E.; Kessler, F.; Moehl, T.; Ahmad, S.; Bessho, T.; Marchioro, A.; Ghadiri, E.; Moser, J.-E.; Yi, C.; Nazeeruddin, Md. K.; Grätzel, M. A Cobalt Complex Redox Shu(cid:425)le for Dye-Sensi(cid:415)zed Solar Cells with High Open-Circuit Poten(cid:415)als. Nat Commun 2012, 3 (1), 631. h(cid:425)ps://doi.org/10.1038/ncomms1655. (57) Boschloo, G.; Hagfeldt, A. Characteris(cid:415)cs of the Iodide/Triiodide Redox Mediator in Dye- Sensi(cid:415)zed Solar Cells. Acc Chem Res 2009, 42 (11), 1819–1826. h(cid:425)ps://doi.org/10.1021/ar900138m. (58) Saygili, Y.; Söderberg, M.; Pellet, N.; Giordano, F.; Cao, Y.; Muñoz-García, A. B.; Zakeeruddin, S. M.; Vlachopoulos, N.; Pavone, M.; Boschloo, G.; Kavan, L.; Moser, J.-E.; Grätzel, M.; Hagfeldt, A.; Freitag, M. Copper Bipyridyl Redox Mediators for Dye-Sensi(cid:415)zed Solar Cells with High Photovoltage. J Am Chem Soc 2016, 138 (45), 15087–15096. h(cid:425)ps://doi.org/10.1021/jacs.6b10721. 104 (59) Feldt, S. M.; Wang, G.; Boschloo, G.; Hagfeldt, A. Effects of Driving Forces for Recombina(cid:415)on and Regenera(cid:415)on on the Photovoltaic Performance of Dye-Sensi(cid:415)zed Solar Cells Using Cobalt Polypyridine Redox Couples. The Journal of Physical Chemistry C 2011, 115 (43), 21500–21507. h(cid:425)ps://doi.org/10.1021/jp2061392. (60) Mosconi, E.; Yum, J.-H.; Kessler, F.; Gómez García, C. J.; Zuccaccia, C.; Cin(cid:415), A.; Nazeeruddin, M. K.; Grätzel, M.; De Angelis, F. Cobalt Electrolyte/Dye Interac(cid:415)ons in Dye- Sensi(cid:415)zed Solar Cells: A Combined Computa(cid:415)onal and Experimental Study. J Am Chem Soc 2012, 134 (47), 19438–19453. h(cid:425)ps://doi.org/10.1021/ja3079016. (61) Rorabacher, D. B. Electron Transfer by Copper Centers. Chem Rev 2004, 104 (2), 651–698. h(cid:425)ps://doi.org/10.1021/cr020630e. (62) Kaim, W.; Rall, J. Copper—A“Modern” Bioelement. Angewandte Chemie Interna(cid:415)onal Edi(cid:415)on in English 1996, 35 (1), 43–60. h(cid:425)ps://doi.org/10.1002/anie.199600431. (63) Birker, P. J. M. W. L.; Helder, J.; Henkel, G.; Krebs, B.; Reedijk, J. Synthesis and Spectroscopic Characteriza(cid:415)on of Copper(I) and Copper(II) Complexes with 1,6-Bis(2- Benzimidazolyl)-2,5-Dithiahexane (BBDH). X-Ray Structure of Trigonal-Bipyramidal Coordinated [Cu(BBDH)Cl]Cl.2C2H5OH. Inorg Chem 1982, 21 (1), 357–363. h(cid:425)ps://doi.org/10.1021/ic00131a064. (64) Freitag, M.; Daniel, Q.; Pazoki, M.; Sveinbjörnsson, K.; Zhang, J.; Sun, L.; Hagfeldt, A.; Boschloo, G. High-Efficiency Dye-Sensi(cid:415)zed Solar Cells with Molecular Copper Phenanthroline as Solid Hole Conductor. Energy Environ Sci 2015, 8 (9), 2634–2637. h(cid:425)ps://doi.org/10.1039/C5EE01204J. (65) Rodrigues, R. R.; Lee, J. M.; Taylor, N. S.; Cheema, H.; Chen, L.; Fortenberry, R. C.; Delcamp, J. H.; Jurss, J. W. Copper-Based Redox Shu(cid:425)les Supported by Preorganized Tetradentate Ligands for Dye-Sensi(cid:415)zed Solar Cells. Dalton Transac(cid:415)ons 2020, 49 (2), 343–355. h(cid:425)ps://doi.org/10.1039/C9DT04030G. (66) Boschloo, G.; Häggman, L.; Hagfeldt, A. Quan(cid:415)fica(cid:415)on of the Effect of 4- Tert - Butylpyridine Addi(cid:415)on to I - /I 3 - Redox Electrolytes in Dye-Sensi(cid:415)zed Nanostructured TiO 2 Solar Cells. J Phys Chem B 2006, 110 (26), 13144–13150. h(cid:425)ps://doi.org/10.1021/jp0619641. (67) Koops, S. E.; O’Regan, B. C.; Barnes, P. R. F.; Durrant, J. R. Parameters Influencing the Efficiency of Electron Injec(cid:415)on in Dye-Sensi(cid:415)zed Solar Cells. J Am Chem Soc 2009, 131 (13), 4808–4818. h(cid:425)ps://doi.org/10.1021/ja8091278. (68) Long, H.; Zhou, D.; Zhang, M.; Peng, C.; Uchida, S.; Wang, P. Probing Dye-Correlated Interplay of Energe(cid:415)cs and Kine(cid:415)cs in Mesoscopic Titania Solar Cells with 4- Tert - 105 Butylpyridine. The Journal of Physical Chemistry C 2011, 115 (29), 14408–14414. h(cid:425)ps://doi.org/10.1021/jp202826m. (69) Giordano, M.; Volpi, G.; Bonomo, M.; Mariani, P.; Garino, C.; Viscardi, G. Methoxy- Subs(cid:415)tuted Copper Complexes as Possible Redox Mediators in Dye-Sensi(cid:415)zed Solar Cells. New Journal of Chemistry 2021, 45 (34), 15303–15311. h(cid:425)ps://doi.org/10.1039/D1NJ02577E. (70) Deng, Z.; Yang, X.; Yang, K.; Zhang, L.; Wang, H.; Wang, X.; Sun, L. Helical Copper Redox Mediator with Low Electron Recombina(cid:415)on for Dye-Sensi(cid:415)zed Solar Cells. ACS Sustain Chem Eng 2021, 9 (15), 5252–5259. h(cid:425)ps://doi.org/10.1021/acssuschemeng.0c08195. (71) Kavan, L.; Saygili, Y.; Freitag, M.; Zakeeruddin, S. M.; Hagfeldt, A.; Grätzel, M. Electrochemical Proper(cid:415)es of Cu(II/I)-Based Redox Mediators for Dye-Sensi(cid:415)zed Solar Cells. Electrochim Acta 2017, 227, 194–202. h(cid:425)ps://doi.org/10.1016/j.electacta.2016.12.185. (72) Kannanku(cid:425)y, K.; Chen, C.-C.; Nguyen, V. S.; Lin, Y.-C.; Chou, H.-H.; Yeh, C.-Y.; Wei, T.-C. Tert - Butylpyridine Coordina(cid:415)on with [Cu(Dmp) 2 ] 2+/+ Redox Couple and Its Connec(cid:415)on to the Stability of the Dye-Sensi(cid:415)zed Solar Cell. ACS Appl Mater Interfaces 2020, 12 (5), 5812– 5819. h(cid:425)ps://doi.org/10.1021/acsami.9b19119. (73) Addison, A. W.; Burke, P. J.; Henrick, K.; Rao, T. N.; Sinn, E. Pentacoordinate Copper Complexes of Nitrogen-Sulfur Donors: Structural Chemistry of Two Complexes of Bis(2-(2- Benzimidazolyl)Ethyl) Sulfide with the Sulfur Alterna(cid:415)vely in Equatorial and Axial Coordina(cid:415)on Modes. Inorg Chem 1983, 22 (24), 3645–3653. h(cid:425)ps://doi.org/10.1021/ic00166a030. (74) Yang, L.; Powell, D. R.; Houser, R. P. Structural Varia(cid:415)on in Copper( i ) Complexes with Pyridylmethylamide Ligands: Structural Analysis with a New Four- Coordinate Geometry Index, τ 4. Dalton Trans. 2007, No. 9, 955–964. h(cid:425)ps://doi.org/10.1039/B617136B. (75) Niu, J.; Zhou, H.; Li, Z.; Xu, J.; Hu, S. An Efficient Ullmann-Type C−O Bond Forma(cid:415)on Catalyzed by an Air-Stable Copper(I)−Bipyridyl Complex. J Org Chem 2008, 73 (19), 7814– 7817. h(cid:425)ps://doi.org/10.1021/jo801002c. (76) Linfoot, C. L.; Richardson, P.; Hewat, T. E.; Moudam, O.; Forde, M. M.; Collins, A.; White, F.; Robertson, N. Subs(cid:415)tuted [Cu(i)(POP)(Bipyridyl)] and Related Complexes: Synthesis, Structure, Proper(cid:415)es and Applica(cid:415)ons to Dye-Sensi(cid:415)sed Solar Cells. Dalton Transac(cid:415)ons 2010, 39 (38), 8945. h(cid:425)ps://doi.org/10.1039/c0dt00190b. (77) Zimmer, K. D.; Shoemaker, R.; Ruminski, R. R. Synthesis and Characteriza(cid:415)on of a Fluxional Re(I) Carbonyl Complex Fac-[Re(CO)3(Dpop(cid:3387))Cl] with the Nominally Tri-Dentate 106 Ligand Dipyrido(2,3-a:3(cid:3387),2(cid:3387)-j)Phenazine (Dpop(cid:3387)). Inorganica Chim Acta 2006, 359 (5), 1478–1484. h(cid:425)ps://doi.org/10.1016/j.ica.2005.11.042. (78) Zhang, J.; Siu, K.; Lin, C. H.; Canary, J. W. Conforma(cid:415)onal Dynamics of Cu(i) Complexes of Tripodal Ligands: Steric Control of Molecular Mo(cid:415)on. New Journal of Chemistry 2005, 29 (9), 1147. h(cid:425)ps://doi.org/10.1039/b509050d. (79) Colombo, A.; Dragone(cid:427), C.; Magni, M.; Roberto, D.; Demar(cid:415)n, F.; Caramori, S.; Bignozzi, C. A. Efficient Copper Mediators Based on Bulky Asymmetric Phenanthrolines for DSSCs. ACS Appl Mater Interfaces 2014, 6 (16), 13945–13955. h(cid:425)ps://doi.org/10.1021/am503306f. (80) Ha(cid:425)ori, S.; Wada, Y.; Yanagida, S.; Fukuzumi, S. Blue Copper Model Complexes with Distorted Tetragonal Geometry Ac(cid:415)ng as Effec(cid:415)ve Electron-Transfer Mediators in Dye- Sensi(cid:415)zed Solar Cells. J Am Chem Soc 2005, 127 (26), 9648–9654. h(cid:425)ps://doi.org/10.1021/ja0506814. (81) Saygili, Y.; Söderberg, M.; Pellet, N.; Giordano, F.; Cao, Y.; Muñoz-García, A. B.; Zakeeruddin, S. M.; Vlachopoulos, N.; Pavone, M.; Boschloo, G.; Kavan, L.; Moser, J.-E.; Grätzel, M.; Hagfeldt, A.; Freitag, M. Copper Bipyridyl Redox Mediators for Dye-Sensi(cid:415)zed Solar Cells with High Photovoltage. J Am Chem Soc 2016, 138 (45), 15087–15096. h(cid:425)ps://doi.org/10.1021/jacs.6b10721. (82) Le Fur, M.; Beyler, M.; Le Poul, N.; Lima, L. M. P.; Le Mest, Y.; Delgado, R.; Platas-Iglesias, C.; Pa(cid:415)nec, V.; Tripier, R. Improving the Stability and Inertness of Cu( ii ) and Cu( i ) Complexes with Methylthiazolyl Ligands by Tuning the Macrocyclic Structure. Dalton Transac(cid:415)ons 2016, 45 (17), 7406–7420. h(cid:425)ps://doi.org/10.1039/C6DT00385K. (83) Cong, J.; Kinschel, D.; Daniel, Q.; Safdari, M.; Gabrielsson, E.; Chen, H.; Svensson, P. H.; Sun, L.; Kloo, L. Bis(1,1-Bis(2-Pyridyl)Ethane)Copper( i / ii ) as an Efficient Redox Couple for Liquid Dye-Sensi(cid:415)zed Solar Cells. J Mater Chem A Mater 2016, 4 (38), 14550–14554. h(cid:425)ps://doi.org/10.1039/C6TA06782D. (84) Yang, K.; Yang, X.; Deng, Z.; Jiang, M. High Stability Tetradentate Ligand Copper Complexes and Organic Small Molecule Hybrid Electrolyte for Dye-Sensi(cid:415)zed Solar Cells. Electrochim Acta 2022, 432, 141108. h(cid:425)ps://doi.org/10.1016/j.electacta.2022.141108. (85) Noviandri, I.; Brown, K. N.; Fleming, D. S.; Gulyas, P. T.; Lay, P. A.; Masters, A. F.; Phillips, L. The Decamethylferrocenium/Decamethylferrocene Redox Couple: A Superior Redox Standard to the Ferrocenium/Ferrocene Redox Couple for Studying Solvent Effects on the Thermodynamics of Electron Transfer. Journal of Physical Chemistry B 1999, 103 (32), 6713–6722. h(cid:425)ps://doi.org/10.1021/jp991381+. 107 (86) Yang, E. S.; Chan, M.-S.; Wahl, A. C. Electron Exchange between Ferrocene and Ferrocenium Ion. Effects of Solvent and of Ring Subs(cid:415)tu(cid:415)on on the Rate. J Phys Chem 1980, 84 (23), 3094–3099. h(cid:425)ps://doi.org/10.1021/j100460a025. (87) Michaels, H.; Rinderle, M.; Freitag, R.; Benesperi, I.; Edvinsson, T.; Socher, R.; Gagliardi, A.; Freitag, M. Dye-Sensi(cid:415)zed Solar Cells under Ambient Light Powering Machine Learning: Towards Autonomous Smart Sensors for the Internet of Things. Chem Sci 2020, 11 (11), 2895–2906. h(cid:425)ps://doi.org/10.1039/C9SC06145B. (88) Bechlars, B.; D’Alessandro, D. M.; Jenkins, D. M.; Iavarone, A. T.; Glover, S. D.; Kubiak, C. P.; Long, J. R. High-Spin Ground States via Electron Delocaliza(cid:415)on in Mixed-Valence Imidazolate-Bridged Divanadium Complexes. Nat Chem 2010, 2 (5), 362–368. h(cid:425)ps://doi.org/10.1038/nchem.585. (89) Bechlars, B.; D’Alessandro, D. M.; Jenkins, D. M.; Iavarone, A. T.; Glover, S. D.; Kubiak, C. P.; Long, J. R. High-Spin Ground States via Electron Delocaliza(cid:415)on in Mixed-Valence Imidazolate-Bridged Divanadium Complexes. Nat Chem 2010, 2 (5), 362–368. h(cid:425)ps://doi.org/10.1038/nchem.585. (90) Raithel, A. L.; Kim, T.-Y.; Nielsen, K. C.; Staples, R. J.; Hamann, T. W. Low-Spin Cobalt( ii ) Redox Shu(cid:425)le by Isocyanide Coordina(cid:415)on. Sustain Energy Fuels 2020, 4 (5), 2497–2507. h(cid:425)ps://doi.org/10.1039/D0SE00314J. (91) Swarts, P. J.; Conradie, J. Redox Data of Ferrocenylcarboxylic Acids in Dichloromethane and Acetonitrile. Data Brief 2020, 30, 105650. h(cid:425)ps://doi.org/10.1016/j.dib.2020.105650. (92) Baillargeon, J.; Xie, Y.; Hamann, T. W. Bifurca(cid:415)on of Regenera(cid:415)on and Recombina(cid:415)on in Dye-Sensi(cid:415)zed Solar Cells via Electronic Manipula(cid:415)on of Tandem Cobalt Redox Shu(cid:425)les. ACS Appl Mater Interfaces 2017, 9 (39), 33544–33548. h(cid:425)ps://doi.org/10.1021/acsami.7b01626. (93) Kannanku(cid:425)y, K.; Chen, C.-C.; Nguyen, V. S.; Lin, Y.-C.; Chou, H.-H.; Yeh, C.-Y.; Wei, T.-C. Tert -Butylpyridine Coordina(cid:415)on with [Cu(Dmp) 2 ] 2+/+ Redox Couple and Its Connec(cid:415)on to the Stability of the Dye-Sensi(cid:415)zed Solar Cell. ACS Appl Mater Interfaces 2020, 12 (5), 5812–5819. h(cid:425)ps://doi.org/10.1021/acsami.9b19119. (94) Yang, K.; Yang, X.; Deng, Z.; Jiang, M. High Stability Tetradentate Ligand Copper Complexes and Organic Small Molecule Hybrid Electrolyte for Dye-Sensi(cid:415)zed Solar Cells. Electrochim Acta 2022, 432, 141108. h(cid:425)ps://doi.org/10.1016/j.electacta.2022.141108. (95) Devdass, A.; Watson, J.; Firestone, E.; Hamann, T. W.; Delcamp, J. H.; Jurss, J. W. An Efficient Copper-Based Redox Shu(cid:425)le Bearing a Hexadentate Polypyridyl Ligand for DSCs 108 under Low-Light Condi(cid:415)ons. ACS Appl Energy Mater 2022, 5 (5), 5964–5973. h(cid:425)ps://doi.org/10.1021/acsaem.2c00344. (96) Bernardo, M. M.; Heeg, M. J.; Schroeder, R. R.; Ochrymowycz, L. A.; Rorabacher, D. B. Comparison of the Influence of Saturated Nitrogen and Sulfur Donor Atoms on the Proper(cid:415)es of Copper(II/I)-Macrocyclic Polyamino Polythiaether Ligand Complexes: Redox Poten(cid:415)als and Protona(cid:415)on and Stability Constants of CuIL Species and New Structural Data. Inorg Chem 1992, 31 (2), 191–198. h(cid:425)ps://doi.org/10.1021/ic00028a013. (97) Rorabacher, D. B. Electron Transfer by Copper Centers. Chem Rev 2004, 104 (2), 651–698. h(cid:425)ps://doi.org/10.1021/cr020630e. (98) Chaka, G.; Kandegedara, A.; Heeg, M. J.; Rorabacher, D. B. Compara(cid:415)ve Study of Donor Atom Effects on the Thermodynamic and Electron-Transfer Kine(cid:415)c Proper(cid:415)es of Copper( ii / i ) Complexes with Sexadentate Macrocyclic Ligands. [Cu II/I ([18]AneS 4 N 2 )] and [Cu II/I ([18]AneS 4 O 2 )]. Dalton Trans. 2007, No. 4, 449–458. h(cid:425)ps://doi.org/10.1039/B612252C. (99) Aronne, L.; Yu, Q.; Ochrymowycz, L. A.; Rorabacher, D. B. Effect of Macrocyclic Ligand Constraints upon the Kine(cid:415)cs of Complex Forma(cid:415)on and Dissocia(cid:415)on and Metal Ion Exchange. Copper(II) Complexes with Cyclohexanediyl Deriva(cid:415)ves of the Cyclic Tetrathiaether [14]AneS4 in 80% Methanol. Inorg Chem 1995, 34 (7), 1844–1851. h(cid:425)ps://doi.org/10.1021/ic00111a036. (100) Deng, H.; Kebarle, P. Bond Energies of Copper Ion−Ligand L Complexes CuL 2 + Determined in the Gas Phase by Ion−Ligand Exchange Equilibria Measurements. J Am Chem Soc 1998, 120 (12), 2925–2931. h(cid:425)ps://doi.org/10.1021/ja973814x. 109 APPENDIX A: CHAPTER 2 SUPPLEMENTARY DATA Figure 2.1A: 1H-NMR spectrum (500 MHz, CDCl3) of dbmed. 110 Figure 2.2A: 1H-NMR spectrum (500 MHz, CDCl3) of [Cu(dbmed)](OTf). 111 3 l c d c 6 2 . 7 M C D 0 3 . 5 r e h t E 9 4 . 3 r e h t E 7 4 . 3 r e h t E 1 2 . 1 19 18 17 16 15 14 13 12 11 10 f1 (ppm) 9 8 7 6 5 4 3 2 1 Figure 2.3A: 1H-NMR spectrum (500 MHz, CDCl3) of [Cu(dbmed)](OTf)2. 250 240 230 220 210 200 190 180 170 160 150 140 130 120 110 100 90 80 70 60 50 40 30 20 10 0 -10 -20 112 Figure 2.4A: Crystal structure of the [Cu(dbmed)]OTf species. Table 2.1A: Selected bond lengths and angles for the [Cu(dbmed)]OTf complex. The bond lengths are reported in angstroms (Å) and bond angles are in degrees (°). The standard devia(cid:415)ons of each value are shown in parenthesis. Atom Atom Length/ Å Atom Atom Atom Angle/° Cu1 Cu1 Cu1 Cu1 N1 N2 N3 N4 2.268(5) N1 1.950(5) N2 2.302(5) N2 1.953(5) N2 N4 N4 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 N3 N1 N3 N4 N1 N3 83.46(18 ) 120.1(2) 82.8(2) 151.0(2) 83.8(2) 118.5(2) 113 APPENDIX B: CHAPTER 3 SUPPLEMENTARY DATA Table 3.1A: Selected bond lengths and angles for the [Cu(bpyPY4)]OTf complex. The bond lengths are reported in angstroms (Å) and bond angles are in degrees (°). The standard devia(cid:415)ons of each value are shown in parenthesis. Atom Atom Length/Å Atom Atom Atom Angle/° Cu1 Cu1 Cu1 Cu1 N1 N2 N3 N4 2.019(2) N1 Cu1 N2 87.61(9) 2.086(2) N1 Cu1 N3 135.49(10) 2.078(2) N3 Cu1 N2 78.64(10) 2.011(2) N4 Cu1 N1 126.30(9) N4 Cu1 N2 137.50(9) N4 Cu1 N3 89.34(9) Table 3.2A: Selected bond lengths and angles for the [Cu(bpyPY4)]OTf2 complex. The bond lengths are reported in angstroms (Å) and bond angles are in degrees (°). The standard devia(cid:415)ons of each value are shown in parenthesis. Atom Atom Length/Å Atom Atom Atom Angle/° Cu1 Cu1 Cu1 Cu1 Cu1 N1 N2 N3 N4 N5 1.969(3) N1 Cu1 N2 81.88(11) 2.208(3) N1 Cu1 N3 88.16(12) 2.053(3) N1 Cu1 N4 81.95(12) 1.993(3) N1 Cu1 N5 172.85(13) 1.980(3) N3 Cu1 N2 86.34(11) N4 Cu1 N2 133.89(12) N4 Cu1 N3 135.76(12) N5 Cu1 N2 96.68(11) N5 Cu1 N3 98.76(12) N5 Cu1 N4 94.20(12) 114 r e h t E 6 4 . 3 r e h t E 5 4 . 3 r e h t E 3 4 . 3 r e h t E 2 4 . 3 e n o t e c A 4 2 . 2 n c 3 d c 7 9 . 1 n c 3 d c 6 9 . 1 N C A 9 9 . 1 r e h t E 5 1 . 1 r e h t E 4 1 . 1 r e h t E 3 1 . 1 19 18 17 16 15 14 13 12 11 10 f1 (ppm ) 9 8 7 6 5 4 3 2 1 Figure 3.1A: : 1H-NMR spectrum (500 MHz, CD3CN) of [Cu(bpyPY4)](TFSI)2. 400 350 300 250 200 150 100 50 0 115 APPENDIX C: CHAPTER 4 SUPPLEMENTARY DATA Figure 4.1A: 1H-NMR of [Cu(PY5)]OTf in deuterated acetonitrile. 116 Figure 4.2A: 1H-NMR of synthesized [Cu(PY5)]OTf2 in deuterated acetonitrile. 117 M C D 7 4 . 5 r e h t e 6 4 . 3 r e h t e 5 4 . 3 r e h t e 3 4 . 3 r e h t e 2 4 . 3 n c 3 d c 9 9 . 1 n c 3 d c 8 9 . 1 n c 3 d c 7 9 . 1 n c 3 d c 7 9 . 1 n c 3 d c 6 9 . 1 n c 3 d c 6 9 . 1 n c 3 d c 5 9 . 1 r e h t e 6 1 . 1 r e h t e 4 1 . 1 r e h t e 3 1 . 1 19 18 17 16 15 14 13 12 11 10 f1 (ppm) 9 8 7 6 5 4 3 2 1 Figure 4.3A: 1H-NMR of dispropor(cid:415)onated [Cu(PY5)]OTf2 in deuterated acetonitrile. 2400 2300 2200 2100 2000 1900 1800 1700 1600 1500 1400 1300 1200 1100 1000 900 800 700 600 500 400 300 200 100 0 -100 -200 118 Figure 4.4A: 1H-NMR of [Cu(PY5)]TFSI in deuterated acetonitrile. 119 Figure 4.5A: 1H-NMR of [Cu(PY5)]TFSI2 in deuterated acetonitrile. Table 4.1A: Selected bond lengths and angles for the [Cu(PY5)]OTf complex. The bond lengths are reported in angstroms (Å) and bond angles are in degrees (°). The standard devia(cid:415)ons of each value are shown in parenthesis. Atom Atom Cu1 Cu1 Cu1 Cu1 N15 N31 N37 N44 Length/Å 2.057(3) 2.100(4) 1.980(4) 1.938(4) Atom Atom Atom Angle/° N15 N37 N37 N44 N44 N44 90.20(14) 88.69(15) 90.51(15) 97.32(14) 118.19(15) 150.50(16) Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 N31 N15 N31 N15 N31 N37 120 Table 4.2A: Selected bond lengths and angles for the [Cu(PY5)]OTf2 complex. The bond lengths are reported in angstroms (Å) and bond angles are in degrees (°). The standard devia(cid:415)ons of each value are shown in parenthesis. N1 N2 N3 N4 N5 N6 2.082(3) 2.035(3) 2.160(3) 2.038(3) 2.090(3) 2.369(3) Atom Atom Length/Å Atom Atom Atom Angle/° 87.45(10) Cu1 175.41(11) Cu1 90.63(11) Cu1 82.47(11) Cu1 87.50(11) Cu1 176.29(11) Cu1 96.93(11) 91.25(12) 177.83(11) 99.16(11) 89.23(11) 81.17(11) 92.07(12) 87.98(11) 93.94(12) Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 Cu1 N1 N1 N1 N2 N2 N2 N2 N2 N3 N4 N4 N4 N4 N5 N5 N3 N5 N6 N1 N3 N4 N5 N6 N6 N1 N3 N5 N6 N3 N6 Table 4.3A: Selected bond lengths and angles for the [Cu(PY5)]TFSI2 complex. The bond lengths are reported in angstroms (Å) and bond angles are in degrees (°). The standard devia(cid:415)ons of each value are shown in parenthesis. Atom Atom N(2) Cu(1) N(3) Cu(1) N(4) Cu(1) N(5) Cu(1) N(6) Cu(1) Length/Å 2.0224(14) 2.0675(14) 2.1130(14) 2.0491(14) 2.0262(14) Atom Atom Atom N(3) Cu(1) N(2) N(4) Cu(1) N(2) N(5) Cu(1) N(2) N(6) Cu(1) N(2) N(4) Cu(1) N(3) N(3) Cu(1) N(5) N(4) Cu(1) N(5) N(3) Cu(1) N(6) N(4) Cu(1) N(6) N(5) Cu(1) N(6) Angle/° 82.00(6) 92.78(6) 98.97(6) 175.34(6) 87.35(6) 175.30(6) 88.00(6) 97.08(6) 91.74(6) 82.32(6) 121