New Spin on Organic Radical Batteries–An Isoindoline Nitroxide

Feb 7, 2018 - Organic electrode materials are a highly promising and environmentally benign class of battery materials with radical polymers being at ...
0 downloads 5 Views 683KB Size
Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES

Article

A New Spin on Organic Radical Batteries - An Isoindoline Nitroxide-Based High-Voltage Cathode Material Kai-Anders Hansen, Jawahar Nerkar, Komba Thomas, Steven E. Bottle, Anthony Peter O'Mullane, Peter Cade Talbot, and James P. Blinco ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.7b18252 • Publication Date (Web): 07 Feb 2018 Downloaded from http://pubs.acs.org on February 8, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

A New Spin on Organic Radical Batteries - An Isoindoline NitroxideBased High-Voltage Cathode Material Kai-Anders Hansen, Jawahar Nerkar, Komba Thomas, Steven E. Bottle, Anthony P. O’Mullane, Peter C. Talbot and James P. Blinco* School of Chemistry, Physics and Mechanical Engineering, Queensland University of Technology, 2 George Street, Brisbane, QLD 4000, Australia. ABSTRACT: Organic electrode materials are a highly promising and environmentally benign class of battery materials with radical polymers at the forefront of this research. Herein we report the first example of the 1,1,3,3tetramethylisoindolin-2-yloxyl (TMIO)-class of nitroxides as an organic electrode material and the synthesis and application of a novel styrenic nitroxide polymer, poly-5-vinyl-1,1,3,3-tetramethylisoindolin-2-yloxyl (PVTMIO). The polymer was synthesized from the precursor monomer 2-methoxy-5-vinyl-1,1,3,3-tetramethylisoindoline (VTMIOMe) and subsequent oxidative deprotection yielded the electroactive radical species. Cyclic voltammetry revealed a high oxidation potential of 3.7 V vs. Li, placing it among the top of the nitroxide-class of electrode materials. The suitability of PVTMIO for utilisation in a high voltage organic radical battery was confirmed with a discharge capacity of 104.7 mAh g-1, high-rate performance and stability under cycling conditions (90% capacity retention after 100 cycles), making it one of the highest reported organic p-dopable cathode materials.

The rapid technological advancement of modern society leads to an increasing need for higher density energy storage. With the advent of hybrid electrical cars, smart phones, tablets and drones, a pressing demand for reliable, scalable, portable and flexible high-performance energy solutions has been created. In particular the provision for on-demand battery power that is able to meet the future needs of our energy intensive society is required.1,2 Lithium-ion technology has dominated the battery market for portable electronics over the last few decades.3-5 However, this established technology, as well as other heavy metal-based battery materials, comes with its own inherent risks and questions about sustainability, safety and availability.6 Organic battery materials have attracted significant interest over the last decade as they have the potential to address the shortcomings of the more established metal-based technologies.7-12 Organic batteries employ carbon-based molecules that contain redox-active moieties that can undergo reversible electron-transfer processes. To prevent dissolution of the active material into the electrolyte and resultant degradation of the electrode, the redox-active molecules are often covalently attached to polymer-backbones.12-16 The benefits of organic battery materials include flexibility, lighter weights, and their decreased environmental impact upon disposal.17 Furthermore, functional group and structural modification allow the redox-properties to be tuned and inherently fast redox-processes lead to higher rate performances. Several classes of organic compounds have been investigated for battery applications, such as carbonyls18, organosulfurs19, triphenylamines20-22 and nitroxides.23-28 Ni-

troxides (stable organic radicals) in particular, are frequently employed in organic radical battery (ORB) designs, due to their high oxidation potential and fast electron transfer kinetics.9,29 The most prominent nitroxide class that has been investigated as ORBs to date is 2,2,6,6tetramethylpiperidinyloxyl (TEMPO) and its methacrylate polymer-derivative poly(2,2,6,6tetramethylpiperidinyloxyl-4-yl methacrylate) (PTMA), which provides a cell voltage of 3.58 V vs. Li.23,30-35 Several other nitroxides have also been studied, including 2,2,5,5tetramethyl pyrrolidinyloxyl (PROXYL),36 nitronyl37 and aryl nitroxides.25 One nitroxide motif that has yet to be investigated as an electrode material is isoindoline-based 1,1,3,3-tetramethylisoindolin-2-yloxyl (TMIO). The fused aromatic ring of TMIO is a highly useful synthetic handle enabling incorporation of a free radical into more complicated structures, including fluorescein-analogues for cellular imaging,38 antioxidant pharmacophore hybrids,39,40

Scheme 1. TMIO and PVTMIO.

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

polyaromatic analogues as probes for materials degradation,41,42 thermal and photoreactive linkers,43,44 and DNA spin labels.45 Herein, we report for the first time the evaluation of TMIO-based redox active materials for energy storage applications. TMIO was directly applied as a pure small molecule as well as in the form of a newly designed polymer, poly-5-vinyl-1,1,3,3-tetramethylisoindolin-2-oxyl (PVTMIO), structurally analogous to polystyrene bearing nitroxide rings fused onto the aryl system. This isoindoline-based nitroxide has a theoretical charge-discharge capacity of 141.0 mAh g-1 for the small molecule and 123.9 mAh g-1 for the polymer. Based on cyclic voltammetry of the small molecule, O n

a

b

N OMe 1

The solubility of the polymer in the battery electrolyte, an important factor for the long-term stability of the device, was investigated by saturating the electrolyte, 1M LiPF6 in ethylene carbonate/diethyl carbonate (1/1 V/V), with the polymer PVTMIO. Despite the high molecular weight the polymer displayed a significant solubility of 33.9 mg/mL (see Figures S10, S11).

c

N OMe 2

n

deprotection of methoxyamine to nitroxide was determined by UV-vis absorption spectroscopy (see ESI) and the formation of the radical character of the polymer was confirmed by EPR spectroscopy. The molecular weight analysis of the prepared PVTMIO 4 by gel permeation chromatography (GPC) indicated a broadening of the observed peaks and a significant molecular weight increase, with a molar mass of Mn = 41 500 g mol-1 and a molecular weight distribution of Đ = 2.34.

N OMe 3

N O 4 PVTMIO

Scheme 2. Schematic representation of the synthesis of PVTMIO 4. a) MePPh3I, 18-crown-6, K2CO3, KOH, 62%; b) AIBN, anisole, 80 °C; c) mCPBA, DCM. the oxidation potential of the isoindoline ring structure is expected to be approximately 3.7 V vs. Li, which, significantly, is 100 mV higher than PTMA, the most prevalent organic radical battery material. Immobilization on a polymer backbone is expected to improve the cycling stability of the device, as small organic molecules are prone to dissolution in the electrolytes. The synthetic pathway to PVTMIO is shown above in Scheme 2, while the small molecule comparative analogue TMIO was prepared as previously published.46 The starting material 5-formyl-2-methoxy-1,1,3,3tetramethylisoindoline 1 was synthesized as previously reported.43,47 The formyl group of 1 was transformed into a polymerizable vinyl functionality via a Wittig reaction with methyltriphenylphosphonium iodide, giving 2methoxy-5-vinyl-1,1,3,3-tetramethylisoindoline (VTMIOMe) 2. Poly-2-methoxy-5-vinyl-1,1,3,3tetramethylisoindoline (PVTMIOMe) 3 was prepared by conventional free radical polymerization of 2 with 2,2′azobisisobutyronitrile AIBN as initiator and toluene as solvent. To achieve a high molecular weight polymerizations with two different initiator concentrations were screened. Polymerization with 2 mol% of AIBN yielded an alkoxyamine polymer with a molecular weight of Mn = 10 900 g mol-1. A lowering of the initiator concentration to 0.5 mol% led to a molecular weight of Mn = 33 800 g mol-1 and a dispersity of Đ = 1.71. This higher molecular weight polymer was subsequently subjected to oxidative cleavage of the pendant methoxyamine-functionality via metachloroperbenzoic acid (mCPBA), yielding the final redoxactive polymer PVTMIO 4 in a yield of 76%. Quantitative

Fig. 1 a) Schematic representation of the redox-couple of polymer PVTMIO 4. b) Cyclic voltammogram of 2 mM TMIO (red) and PVTMIO (black) in dichloromethane, 0.1 M tetrabutylammonium hexafluorophosphate (C4H10)4N+PF6-; scan rate: 50 mV/s. c) Cyclic voltammograms of the assembled lithium-ion coin cells from TMIO (red) and PVTMIO (black): (10/80/10 wt% TMIO or PVTMIO/Super P®/PVDF), 1 M LiPF6/EC/DEC; scan rate: 5 mV/s.

ACS Paragon Plus Environment

Page 2 of 7

Page 3 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces Table 1 Experimental Redox Potentials (V) of investigated nitroxides materials.

TMIO PVTMIO a

Epa 1.02 0.89

vs Ag/AgCl Epc 0.90 0.84

E1/2a 0.97 0.87

Epa 3.80

vs Li/Li+ Epc 3.58

E1/2a 3.69

E1/2 = (Epa + Epc)/2

The electrochemical behavior of TMIO and the redoxactive polymer 4 was initially investigated by cyclic voltammetry (CV) undertaken in organic electrolyte. A 0.1 M solution of tetrabutylammonium hexafluorophospate (TBAPF6) in dichloromethane (DCM) served as the electrolyte in a three-electrode set-up with a glassy carbon working electrode, Ag/AgCl as the reference, and platinum wire as the counter electrode. The cyclic voltammogram displayed the expected48 reversible redox-wave at E1/2 = 0.97 V vs. Ag/AgCl for TMIO and E1/2 = 0.87 V vs. Ag/AgCl for PVTMIO representing the redox transition between the nitroxide radical and the oxoammonium cation (Figure 1a, 1b and Table 1). Notably, these oxidation potentials are 100-200 mV higher than for PTMA, the most prominent nitroxide-based material. This potential is amongst the highest reported for this class of materials and, in conjunction with the ~6% decrease in molar mass over PTMA, provides significantly increased energy density.9,23 Higher oxidation potentials are the main the advantage of isoindoline nitroxides, with most small molecule TMIO analogues displaying 100-200 mV higher oxidation potentials compared with their TEMPO counterparts.48

this peak belongs to the oxidation of the methoxyamine precursor and appeared when incomplete deprotection of the methoxyamine polymer occurred. However, the formation of the radical cation of the methoxyamine is unstable, leading to the cleavage of the alkoxyamine und partial regeneration of the nitroxide (Figure S14). Coincidentally, this phenomenon of nitroxide generation via electrochemical cleavage of alkoxyamines has recently been published for the first time by Ciampi et al. while we were summarizing the results of this study.49 While the increase in oxidation potential is desirable (over 300 mV increase), the cycling instability of the group demonstrates why the nitroxide moiety is preferred. As such, only cells with PVTMIO 4 as the active material were studied in more detail. Galvanostatic charge and discharge experiments between 2.5 V and 4.2 V at different C-rates were performed to determine the capacity and performance of the coin cells

To test the performance as organic cathode materials, coin cell batteries in a half-cell configuration were fabricated with lithium metal as anode material and carbon composite cathodes. Initially, the electrodes were prepared with 10% of active material (TMIO or 4/Super P®/PVDF; 10/80/10 wt%), a typical composition for this type of battery that requires large amounts of conductive carbon additive. Surprisingly, cyclic voltammetry of the coin cell with TMIO as the active material displayed only capacitive behaviour and did not feature any notable redox-activity. This indicates that sublimation of TMIO occurs and hence removal from the electrode surface during the drying process under reduced pressure at 50˚C over a period of 24h (Figures 1c and S12). This result indicates that the small molecule TMIO is not suitable for battery manufacturing under current assembly conditions and further highlights the necessity of immobilization of the small molecule onto a polymer backbone. Cyclic voltammetry of the polymer-based coin cell, on the other hand, displayed a highly reversible redox peak at E1/2 = 3.69 V (Figure 1c), analogous to the solution-based cyclic voltammetry. The redox peak corresponds to the nitroxideoxoammonium cation transition (Figure 1a). Interestingly, some cells exhibited a second minor peak at E1/2 = 3.98 V (Figure S13). This peak also appears reversible in nature, but with prolonged cycling did appear to undergo degradation. Upon further investigation, it was discovered that

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

age of 3.57 V and a specific discharge capacity of 68.4 mAh g-1 were retained, corresponding to 55% of the initial capacity at 1C. This corresponds to a full charge available in as little as 36 seconds. This excellent performance at high rates is based on the fast electrode kinetics and the rapid electron-transfer processes of nitroxide radicals.50 Upon return to 1C after 35 cycles the capacity returns to 99.8 mAh g-1, comparable to the initial value, demonstrating the cycling stability of the electrode at different rates (Figure 2b). To further investigate the stability of the PVTMIO electrode, galvanostatic cycling at a constant current of 1C was applied after conditioning the coin cell for 2 cycles at 0.1C (Figure 2c). After 100 cycles the capacity dropped from an initial 110 mAh g-1 (88% of theoretical capacity) to 98.1 mAh g-1 (79%) corresponding to a capacity retention of 90%, demonstrating a reasonable cycling stability. The gradual loss of capacity has also been observed in other polymeric electrode materials and is due to the slow dissolution of the polymer into the electrolyte over time (vide supra).12-16 During the first cycle the coulombic efficiency (CE) is only 96.5% (Figure 2c inset). This lower efficiency can be explained by the initial formation of a solid electrolyte interface (SEI) layer. Once this layer is stabilized the CE remains consistently at >99%. While all reported nitroxide active materials require a conductive carbon additive during cell fabrication, a higher active material fraction of 40 wt% was also assessed. Accordingly, due to the polymers having a much

Fig. 2

a) Galvanostatic charge/discharge at different C-rates of lithium-ion coin cell with composite electrode of 4/Super P®/PVDF 10/80/10 wt% in 1 M LiPF6/EC/DEC. b) Discharge capacities at different C-rates from 1C to 100C. c) Extended charge/discharge cycling of lithium-ion coin cell at 1C, 100 cycles. Inset: Coulombic efficiency (CE%) of corresponding charge/discharge cycles.

(Figure 2a). The device displayed a single voltage plateau at 3.69 V, consistent with the CV data obtained with the coin cell (Figure 1c). At 1C a specific discharge capacity of 104.7 mAh g-1 was observed, corresponding to 84% of the theoretical capacity. The rate capability of the PVTMIO electrode was tested by cycling through different C-rates ranging from 1C to 100C. The voltage plateau starts to separate to slightly higher and lower voltages for the charging and discharging process, respectively, indicating only minor polarization of the electrodes, and a decrease in capacity is observed with increasing C-rates. At 100C a discharge volt-

Fig. 3 Comparison of the galvanostatic charge/discharge at 1 C of lithium-ion coin cell with composite electrode of 4/Super P®/PVDF 10/80/10 wt% (black) and 40/50/10 wt% in 1 M LiPF6/EC/DEC (red).

higher density compared to the carbon additive, the mass loadings of the electrodes increased substantially, i.e. 3.05 mg/cm-2 vs. 0.8 mg/cm-2. This increased loading resulted in a cell with a specific capacity of 60.7 mAh g-1. This is approximately half of the theoretical capacity and well in line with other nitroxide-based battery materials when high loadings such as this are undertaken.15 However, longer preconditioning (50 cycles at 1C) was required in order for the cell to stabilize. This slow but steady increase can be explained by the time it takes for the electrolyte to fully penetrate the electrode microstructure which contains significantly more active material. In ab-

ACS Paragon Plus Environment

Page 4 of 7

Page 5 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces solute terms however, there is a significant increase in the actual capacity for the cells compared with the 10 wt% systems (see Figure 3). In summary, we report the synthesis of a novel redoxactive nitroxide-polymer poly-2-methoxy-5-vinyl-1,1,3,3tetramethylisoindoline (PVTMIO) and the first evaluation of isoindoline nitroxide-based redox-active materials for energy storage applications. This study establishes the TMIO structure as the basis for high voltage cathode materials in lithium-ion batteries. The synthetic strategy involved the protection of the nitroxide radical of a previously reported precursor and the two-step conversion of a formyl-functionality into a vinyl group, yielding a styrenetype monomer. Free radical polymerization and deprotection of the nitroxides yielded the redox-active polymer. Lithium-coin cells were fabricated from composite electrodes of TMIO and PVTMIO and were electrochemically characterized. While the small molecule TMIO proved to be unsuitable in this context the polymer PVMTIO was successfully applied in a prototype battery which displayed a discharge voltage of 3.69 V, putting it among the top of this class of nitroxide materials. The prototype cell produced a very good specific capacity of 104.7 mAh g-1, corresponding to 84% material activity. Good rechargeability over 100 cycles with minimal loss and >99% coulombic efficiency were observed, and the system demonstrated that higher loading of the active material was also possible. All of these properties demonstrate that PVTMIO and the TMIO motif are highly promising candidates as high voltage cathode materials in organic radical batteries.

ASSOCIATED CONTENT Supporting Information including full experimental details, NMR and EPR characterization, device assembly details are supplied. This material is available free of charge via the Internet at http://pubs.acs.org.”

AUTHOR INFORMATION Corresponding Author *Email: [email protected].

ACKNOWLEDGMENTS We gratefully acknowledge financial support for this work from the Ian Potter Foundation and the Queensland University of Technology. KAH acknowledges QUT for his postgraduate research award. The data reported in this paper was obtained at the Central Analytical Research Facility (CARF) operated by the Institute for Future Environments (QUT). Access to CARF is supported by funding from the Science and Engineering Faculty (QUT). We thank Yin Zhang for support in device assembly and Fawad Ali for assistance with the SEM measurement.

REFERENCES 1. Larcher, D.; Tarascon, J. M. Towards Greener and More Sustainable Batteries for Electrical Energy Storage. Nat. Chem. 2015, 7, 19-29.

2. Haight R.; Haensch W.; Friedman D. Solar-Powering the Internet of Things. Science 2016, 353, 124-125. 3. Song, Z.; Zhou, H. Towards Sustainable and Versatile Energy Storage Devices: An Overview of Organic Electrode Materials. Energy Environ. Sci. 2013, 6, 2280-2301. 4. Goodenough, J. B.; Park, K.-S. The Li-Ion Rechargeable Battery: A Perspective. J. Am. Chem. Soc. 2013, 135, 1167-1176. 5. Sun, Y.; Liu, N.; Cui, Y. Promises and Challenges of Nanomaterials for Lithium-Based Rechargeable Batteries. Nat. Energy 2016, 1, 16071. 6. Miroshnikov, M.; Divya, K. P.; Babu, G.; Meiyazhagan, A.; Reddy Arava, L. M.; Ajayan, P. M.; John, G. Power from Nature: Designing Green Battery Materials from Electroactive Quinone Derivatives and Organic Polymers. J. Mater. Chem. A 2016, 4, 12370-12386. 7. Zhu, Z.; Chen, J. Advanced Carbon-Supported Organic Electrode Materials for Lithium (Sodium)-Ion Batteries. J. Electrochem. Soc. 2015, 162, A2393-A2405. 8. Schon, T. B.; McAllister, B. T.; Li, P.-F.; Seferos, D. S. The Rise of Organic Electrode Materials for Energy Storage. Chem. Soc. Rev. 2016, 45, 6345-6404. 9. Janoschka, T.; Hager, M. D.; Schubert, U. S. Powering up the Future: Radical Polymers for Battery Applications. Adv. Mater. 2012, 24, 6397-6409. 10. Zhao, Q.; Lu Y.; Chen, J. Advanced Organic Electrode Materials for Rechargeable Sodium-Ion Batteries. Adv. Energy Mater. 2017, 7 (8), 1601792-n/a. 11. Casado, N.; Hernández, G.; Sardon, H.; Mecerreyes, D. Current Trends in Redox Polymers for Energy and Medicine. Prog. Polym. Sci. 2016, 52, 107-135. 12. Muench, S.; Wild, A.; Friebe, C.; Häupler, B.; Janoschka, T.; Schubert, U. S. Polymer-Based Organic Batteries. Chem. Rev. 2016, 116, 9438-9484. 13. Suga, T.; Konishi, H.; Nishide, H. Photocrosslinked Nitroxide Polymer Cathode-Active Materials for Application in an Organicbased Paper Battery. Chem. Commun. 2007, 1730-1732. 14. Bugnon, L.; Morton, C. J. H.; Novak, P.; Vetter J.; Nesvadba, P. Synthesis of Poly(4-methacryloyloxy-TEMPO) via Group-Transfer Polymerization and Its Evaluation in Organic Radical Battery. Chem. Mater. 2007, 19, 2910-2914. 15. Zhang, K.; Hu, Y.; Wang, L.; Fan, J.; Monteiro, M. J.; Jia, Z. The Impact of the Molecular Weight on the Electrochemical Properties of Poly(TEMPO methacrylate). Polym. Chem. 2017, 8, (11), 1815-1823. 16. Hansen, K.-A.; Blinco, J. P. Nitroxide Radical Polymers - A Versatile Material Class for High-Tech Applications. Polym. Chem. 2018, DOI: 10.1039/C7PY02001E. 17. Nishide, H.; Koshika K.; Oyaizu, K. Environmentally Benign Batteries Based on Organic Radical Polymers. Pure Appl. Chem. 2009, 81, 1961-1970. 18. Vlad, A.; Arnould, K.; Ernould, B.; Sieuw, L.; Rolland, J.; Gohy, J.-F. Exploring the Potential of Polymer Battery Cathodes with Electrically Conductive Molecular Backbone. J. Mater. Chem. A 2015, 3, 11189-11193. 19. Chung, W. J.; Griebel, J. J.; Kim, E. T.; Yoon, H.; Simmonds, A. G.; Ji, H. J.; Dirlam, P. T.; Glass, R. S.; Wie, J. J.; Nguyen, N. A.; Guralnick, B. W.; Park, J.; Somogyi, Á.; Theato, P.; Mackay, M. E.; Sung, Y.-E.; Char, K.; Pyun, J. The Use of Elemental Sulfur as an Alternative Feedstock for Polymeric Materials. Nat. Chem. 2013, 5, (6), 518-524. 20. Feng, J. K.; Cao, Y. L.; Ai, X. P.; Yang, H. X. Polytriphenylamine: A High Power and High Capacity Cathode Material for

ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Rechargeable Lithium Batteries. J. Power Sources 2008, 177, (1), 199-204. 21. Su, C.; Yang, F.; Xu, L.; Zhu, X.; He, H.; Zhang, C. Radical Polymer Containing a Polytriphenylamine Backbone: Its Synthesis and Electrochemical Performance as the Cathode of Lithium-Ion Batteries. ChemPlusChem 2015, 80, (3), 606-611. 22. Zhu, L.; Cao, X. p-Dopable Poly(4-cyano)triphenylamine: A High Voltage Organic Cathode for Lithium Ion Batteries. Mater. Lett. 2015, 150, 16-19. 23. Nakahara, K.; Iwasa, S.; Satoh, M.; Morioka, Y.; Iriyama, J.; Suguro M.; Hasegawa, E. Rechargeable Batteries with Organic Radical Cathodes. Chem. Phys. Lett. 2002, 359, 351-354. 24. Nakahara, K.; Iriyama, J.; Iwasa, S.; Suguro, M.; Satoh, M.; Cairns, E. J. Al-Laminated Film Packaged Organic Radical Battery for High-Power Applications. J. Power Sources 2007, 163, 11101113. 25. Suga, T.; Pu, Y.-J.; Kasatori, S.; Nishide, H. Cathode- and Anode-Active Poly(nitroxylstyrene)s for Rechargeable Batteries:  p- and n-Type Redox Switching via Substituent Effects. Macromolecules 2007, 40 (9), 3167-3173. 26. Nishide, H.; Oyaizu, K. Toward Flexible Batteries. Science 2008, 319, 737-738. 27. Katsumata, T.; Qu, J.; Shiotsuki, M.; Satoh, M.; Wada, J.; Igarashi, J.; Mizoguchi, K.; Masuda, T. Synthesis, Characterization, and Charge/Discharge Properties of Polynorbornenes Carrying 2,2,6,6-Tetramethylpiperidine-1-oxy Radicals at High Density. Macromolecules 2008, 41, 1175-1183. 28. Suga, T.; Ohshiro, H.; Sugita, S.; Oyaizu, K.; Nishide, H. Emerging N-Type Redox-Active Radical Polymer for a Totally Organic Polymer-Based Rechargeable Battery. Adv. Mater. 2009, 21, 1627-1630. 29. Zhang, K.; Monteiro, M. J.; Jia, Z. Stable Organic Radical Polymers: Synthesis and Applications. Polym. Chem. 2016, 7, 5589-5614. 30. Nishide, H.; Iwasa, S.; Pu, Y.-J.; Suga, T.; Nakahara, K.; Satoh, M. Organic Radical Battery: Nitroxide Polymers as a CathodeActive Material. Electrochim. Acta 2004, 50, 827-831. 31. Janoschka, T.; Martin, N.; Martin, U.; Friebe, C.; Morgenstern, S.; Hiller, H.; Hager, M. D.; Schubert, U. S. An Aqueous, PolymerBased Redox-Flow Battery Using Non-Corrosive, Safe, and LowCost Materials. Nature 2015, 527, 78-81. 32. Ernould, B.; Devos, M.; Bourgeois, J.-P.; Rolland, J.; Vlad A.; Gohy, J.-F. Grafting of a Redox Polymer onto Carbon Nanotubes for High Capacity Battery. J. Mater. Chem. A 2015, 3, 8832-8839. 33. Guo, W.; Yin, Y.-X.; Xin, S.; Guo, Y.-G.; Wan, L.-J. Superior Radical Polymer Cathode Material with a Two-Electron Process Redox Reaction Promoted by Graphene. Energy Environ. Sci. 2012, 5, 5221-5225. 34. Guo, W.; Su, J.; Li, Y.-H.; Wan, L.-J.; Guo, Y.-G. Nitroxide Radical Polymer/Graphene Nanocomposite as an Improved Cathode Material for Rechargeable Lithium Batteries. Electrochim. Acta 2012, 72, 81-86. 35. Li, Y.; Jian, Z.; Lang, M.; Zhang, C.; Huang, X. Covalently Functionalized Graphene by Radical Polymers for GrapheneBased High-Performance Cathode Materials. ACS Appl. Mater. Interfaces 2016, 8, 17352-17359. 36. Qu, J.; Katsumata, T.; Satoh, M.; Wada, J.; Masuda, T. Synthesis and Properties of Polyacetylene and Polynorbornene

Derivatives Carrying 2,2,5,5-Tetramethyl-1-pyrrolidinyloxy Moieties. Macromolecules 2007, 40, 3136-3144. 37. Hagemann, T.; Winsberg, J.; Haupler, B.; Janoschka, T.; Gruber, J. J.; Wild A.; Schubert, U. S. A Bipolar Nitronyl Nitroxide Small Molecule for an All-Organic Symmetric Redox-Flow Battery. NPG Asia Mater. 2017, 9, e340. 38. Morrow, B. J.; Keddie, D. J.; Gueven, N.; Lavin M. F.; Bottle, S. E. A Novel Profluorescent Nitroxide as a Sensitive Probe for the Cellular Redox Environment. Free Radical Biol. Med. 2010, 49, 6776. 39. Thomas, K. Novel Dual-Acting Nitroxide-based Hybrid Antioxidants. Ph.D. Thesis, Queensland University of Technology, Brisbane, Australia, 2016. 40. Walker, J.R.; Fairfull-Smith, K.E.; Kazunori, A.; Lau, S. White, P.J.; Scammells, P.J.; Bottle, S.E. Edaravone Containing Isoindoline Nitroxides for the Potential Treatment of Cardiovascular Ischaemia. Med. Chem. Commun. 2011, 2, 436441. 41. Blinco, J. P.; Keddie, D. J.; Wade, T. Barker, P. J.; George G. A.; Bottle, S. E. Profluorescent Nitroxides: Sensors and Stabilizers of Radical-Mediated Oxidative Damage. Polym. Degrad. Stab. 2008, 93, 1613-1618. 42. Lussini, V. C.; Colwell, J. M.; Fairfull-Smith K. E.; Bottle, S. E. Profluorescent Nitroxide Sensors for Monitoring Photo-Induced Degradation in Polymer Films. Sens. Actuators, B 2017, 241, 199209. 43. Lederhose, P.; Haworth, N. L.; Thomas, K.; Bottle, S. E.; Coote, M. L.; Barner-Kowollik C.; Blinco, J. P. Design of Redox/Radical Sensing Molecules via Nitrile Imine-Mediated Tetrazole-ene Cycloaddition (NITEC). J. Org. Chem. 2015, 80, 8009-8017. 44. Blinco, J. P.; Fairfull-Smith, K. E.; Micallef A. S.; Bottle, S. E. Highly Efficient, Stoichiometric Radical Exchange Reactions Using Isoindoline Profluorescent Nitroxides. Polym. Chem. 2010, 1, 1009-1012. 45. Chalmers, B. A.; Saha, S.; Nguyen, T.; McMurtrie, J.; Sigurdsson, S. T.; Bottle S. E.; Masters, K.-S. TMIO-PyrImid Hybrids are Profluorescent, Site-Directed Spin Labels for Nucleic Acids. Org. Lett. 2014, 16, 5528-5531. 46. Griffiths, P.; Moad, G.; Rizzardo, E. Synthesis of the Radical Scavenger 1,1,3,3-Tetramethylisoindolin-2-yloxyl. Aust. J. Chem. 1983, 36 (2), 397-401. 47. Chalmers, B. A.; Morris, J. C.; Fairfull-Smith, K. E.; Grainger R. S.; Bottle, S. E. A Novel Protecting Group Methodology for Syntheses Using Nitroxides. Chem. Commun. 2013, 49, 1038210384. 48. Blinco, J. P.; Hodgson, J. L.; Morrow, B. J.; Walker, J. R.; Will, G. D.; Coote, M. L.; Bottle, S. E. Experimental and Theoretical Studies of the Redox Potentials of Cyclic Nitroxides. J. Org. Chem. 2008, 73, 6763-6771. 49. Zhang, L.; Laborda, E.; Darwish, N.; Noble, B. B.; Tyrell, J. H.; Pluczyk, S.; Le Brun, A. P.; Wallace, G. G.; Gonzalez, J.; Coote, M. L.; Ciampi, S. Electrochemical and Electrostatic Cleavage of Alkoxyamines. J. Am. Chem. Soc., 2018, 140, 766-774. 50. Oyaizu, K.; Nishide, H. Radical Polymers for Organic Electronic Devices: A Radical Departure from Conjugated Polymers? Adv. Mater. 2009, 21, 2339-2344.

ACS Paragon Plus Environment

Page 6 of 7

Page 7 of 7 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

ACS Paragon Plus Environment

7