Electrochemical Energy Storage with an Aqueous ... - ACS Publications

Jan 12, 2018 - KEYWORDS: energy storage, zinc anode, Quinone cathode, solid electrolyte, mediator ion. Green energy sources, such as wind ... reliable...
1 downloads 5 Views 1MB Size
Subscriber access provided by READING UNIV

Letter

Electrochemical Energy Storage with an Aqueous ZincQuinone Chemistry Enabled by a Mediator-ion Solid Electrolyte Xingwen Yu, and Arumugam Manthiram ACS Appl. Energy Mater., Just Accepted Manuscript • DOI: 10.1021/acsaem.7b00089 • Publication Date (Web): 12 Jan 2018 Downloaded from http://pubs.acs.org on January 12, 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 free 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 accessible to all readers and 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 Energy Materials 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 Energy Materials

Electrochemical Energy Storage with an Aqueous Zinc-Quinone Chemistry Enabled by a Mediator-ion Solid Electrolyte Xingwen Yu and Arumugam Manthiram* Materials Science and Engineering Program and Texas Materials Institute, The University of Texas at Austin, Austin, TX 78712, USA

ABSTRACT: Quinone series of organics are promising electrode materials for the development of low-cost, sustainable, environmentally benign electrochemical energy storage technologies. However, the redox potential of quinones is sensitively dependent on the acidity and alkalinity of aqueous solutions. This study demonstrates a high-voltage aqueous Zn-quinone battery with an alkaline anode electrolyte (anolyte) and an acidic cathode electrolyte (catholyte), which are separated by a sodium mediator-ion solid electrolyte membrane. The redox chemistries of the acidic quinone cathode and the alkaline Zn anode are ionically linked by the shuttling of sodium mediator ions through the solid electrolyte.

Green energy sources, such as wind and solar, are increasingly explored in recent years and are expected to be deployed in larger numbers in the future.1,2 To efficiently use these energy sources, appropriate energy storage technologies are required, since both the solar and the wind energies are harvested intermittently. Rechargeable batteries, a class of important electrochemical energy technologies, offer a reliable, efficient, and simple way of storing the wind and solar energies.3-11 Among the currently available secondary battery technologies, lithium-ion (Li+-ion) batteries deliver the highest energy density and, therefore, are in the forefront in impacting the portable electronics market and electric vehicles. However, large-scale grid storage requires costeffective battery systems with low maintenance metrics. In this regard, aqueous batteries are relatively preferable than

nonaqueous batteries. At the current stage, most aqueous battery systems are based on redox reactions of inorganic electrode materials, which sometimes are constrained by material abundance and environmental concerns. Therefore, there have been increasing attention in exploring organic-based electrode materials which offer great opportunities of constructing low-cost, environmentally benign, sustainable energy storage systems/devices.12,13 As a class of redox-active moieties, quinones have widely been used in biological systems in controlling electron-transport processes. 14,15 Recently, their applications in electrochemical energy storage have been explored as well.16,17 In this study, we explore using hydroquinone (HQ) solution as a catholyte for the development of an aqueous battery system. The redox reaction of the hydroquinone is based on a reorganization process of the conjugated double bond, as schematized in Figure 1a. Generally, it is believed that the reduction process involves the transformation of cyclic diketone core into an aromatic ring after releasing two electrons. The oxidation reaction is a reverse process. The redox reaction is based on a 2 e- / 2 H+ transfer process, which results in the reorganization of the double-bonds.18,19 Based on the Nernst equation and a series of electrochemical experiments we carried out, the reduction potential of the hydroquinone depends strongly on the acidity and alkalinity of aqueous solutions.20 Figure. 1b and c show the linear sweep voltammograms of hydroquinone respectively in an acid and an alkaline solution on a carbon foam electrode matrix (The

Figure 1. (a) Redox reaction and electrochemical transition of hydroquinone and benzoquinone. (b, c) Linear sweep voltammograms (LSVs) of a carbon nanofoam electrode in (b) 0.1 M hydroquinone + 0.1 M NaOH solution and (c) 0.1 M hydroquinone + 0.05 M H 2SO4 + 0.05 M Na2SO4 solution at a scan rate of 1.0 mV s-1.

ACS Paragon Plus Environment

1

ACS Applied Energy Materials 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

Page 2 of 7

Figure 2. (a) Schematic of a Zn-quinone battery with an alkaline anolyte, an acidic catholyte, and a Na+-ion solid electrolyte (termed as Zn(base) ǁ Na-SSE ǁ HQ(acid) cell). (b) Scanning electron microscopy (SEM) image of the carbon nanofoam matrix, showing a carbon fiber surrounded/attached by the carbon nanofoam. (c) High-magnification image of the carbon nanofoam.

acidity and alkalinity of the acidic and alkaline quinone solutions are provided in the caption to Figure 1). It is noticed that the reduction potential of hydroquinone under an acidic condition is ~ 0.6 V higher than that under an alkaline condition, which is consistent with the calculation results based on the Nernst equation.20 Therefore, an acidic catholyte is preferred when developing a battery with the hydroquinone as an active positive electrode. From the negative electrode perspective, zinc is supposed to be one of the best options due to its sustainability, environmentally benignity and low-cost feature. Coupling the Zn negative electrode with the hydroquinone positive electrode is expected to develop a high-energy and low-cost battery system. As a general scenario, the Zn negative electrode has to be operated under an alkaline condition owing to its intrinsic corrosion problem in acidic solutions. 21,22 However, it is impossible to develop a battery with an alkaline negative electrode and an acidic positive electrode under the traditional battery operating principle with a conventional porous battery separator. 23,24 A mediator-ion battery development strategy provides this possibility. Figure 2a shows the schematic of a proposed mediator-ion Zn-quinone battery with an alkaline Zn negative electrode, an acidic hydroquinone positive electrode and a sodium-ion solid-state electrolyte (Na-SSE). The Na-SSE physically and electrically separates the alkaline anode/anolyte and the acidic cathode/catholyte. Na+-ions shuttle through the Na-SSE as an ionic mediator to sustain redox chemistries of the acidic hydroquinone positive electrode and the alkaline Zn negative electrode. The cell is accordingly annotated as Zn(base) ǁ NaSSE ǁ HQ(acid) cell. The anolyte, catholyte and solid electrolyte used for this study are summarized in Table 1. During the cell operation, the Na-SSE provides ionic channels for Na+-ions to migrate back and forth as ionic mediators to facilitate charge balance between the two electrodes. As schematically illustrated in Figure 2a, during the charge process, the zinc oxide (ZnO) film on the surface of the negative electrode (may also exists as Zn(OH) 42-) is re-

duced into Zn metal and OH -. The metallic Zn is deposited onto the negative electrode and the OH- associates with Na+ to form NaOH. At the positive electrode, the hydroquinone is oxidized to benzoquinone. To balance the ionic charge between the negative electrode and the positive electrode, Na+ ions transport through the Na-SSE from the catholyte to the anolyte. During the discharge, the processes are reversed. At the negative electrode, the Zn metal is oxidized to Zn2+, which combines with OH- to form the Zn(OH)42- species. At the positive electrode, the benzoquinone is reduced to hydroquinone. The Na+ ions in the anolyte migrate through the NaSSE to the catholyte to form Na2SO4. It should be noted here that the use of a Na-SSE to separate an acidic electrolyte and an alkaline electrolyte may impact both ion transport and mass transport in the cell. According to the positive electrode reaction illustrated in Figure 1a, there would be a pH change during the charge-discharge of the cell. However, it is not clear whether or not the Na-SSE membrane transports protons during cell cycling. This issue needs to be investigated in detail in the future. The Na-SSE used in this study is a NASICON-type solid electrolyte, Na3Zr2Si2PO12, purchased from 421 Energy Corporation, South Korea. It provides a Na+-ion conductivity of ~ 1.0 x 10-3 S. cm-1 at room temperature. Na3Zr2Si2PO12 has previously been proved as a good solid electrolyte in a sodium-seawater battery that does not allow aqueous solutions to migrate through.25 The use of Na3Zr2Si2PO12 ensures the elimination of the mixing of the two liquid electrolytes. A carbon nanofoam (purchased from MarkeTech, average pore size: 0.7 nm, surface area: 400 m2 g-1) was used as the positive electrode matrix to provide a facile environment for the electrochemical reactions of hydroquinone/benzoquinone. Figure 2b and c show the scanning electron microscope (SEM) images of the carbon nanofoam electrode. The carbon fibers (as marked in Figure 2b) act as a reinforcement to enhance the mechanical property of the carbon nanofoam. The carbon nanofoam with a high surface area provides a facile environment for the redox reactions of liquid electrode materials.

Table 1. Anolyte, catholyte, and solid-state electrolyte used in the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell Cell notation

solid electrolyte

anolyte

Catholyte

Zn(base) ǁ Na-SSE ǁ HQ(acid)

Na3Zr2Si2PO12 (NZSP)

0.1 M NaOH

0.1 M hydroquinone + 0.05 M H2SO4 + 0.05 M Na2SO4

ACS Paragon Plus Environment

2

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 Energy Materials

Figure 3. (a) A full charge curve of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell. (b) Ultraviolet-visible (UV-Vis) spectroscopy data of the fresh catholyte and those after charging the cell to the states indicated with the red open ovals in the discharge curve in Figure 3(a).

The hydroquinone is in its discharged state after the cell was assembled. Figure 3a shows the first charge curve of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell. At the end of charge with a sharp rising-voltage, the hydroquinone delivered a high specific capacity of ~ 400 mA h g -1. Comparison of this discharge capacity with the theoretical capacity of the hydroquinone (487 mA h g-1, based on a 2-electron redox reaction shown in Figure 1a and a molecular weight of 110 g mol-1 for hydroquinone), ~ 82% of the active hydroquinone was utilized. To investigate the depth of charge of the hydroquinone positive electrode, ultraviolet-visible spectroscopy (UV-Vis) experiments were performed. Figure 3b presents the UV-Vis spectra of the hydroquinone catholyte collected after the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell was charged, respectively, to the half-charge and the full-charge states, as indicated in Figure 3a with the red circles on the first charge profile. The samples collected are as pictured in Figure S1. The pristine catholyte shows an absorption peak of the hydroquinone at a wavelength of 295 nm. 26,27 After the cell was charged to the half–way state, the UV-Vis spectrum exhibits mixed absorption characteristics of the hydroquinone and benzoquinone species.28,29 The benzoquinone species usually has a characteristic absorption peak at 245 nm.30 Upon a deep charge to a cutoff voltage of 2.6 V, the hydroquinone absorption peak becomes very small, and the benzoquinone absorption becomes dominant. The concentrations of hydroquinone and benzoquinone species have also been quantitatively analyzed. To do this, a set of standard hydroquinone solutions with various assigned concentrations were prepared and measured with UV-Vis. With a set of calibration curves obtained (as shown in Figure S2), concentrations of hydroquinone species in the catholytes at the half-charge and full-charge states can be determined. The results are summarized in Table 2, which are consistent with the results obtained

Table 2. Concentrations of hydroquinone species in the catholyte of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell at different charge states. The margin of error is + 0.001 M. Concentration of hydroquinone, M

Before charge 0.1

Half-way charge 0.056

from the electrochemical experiments in Figure 3a. Figure 4 summaries the electrochemical performances of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell. The polarization behavior of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell is presented in Figure 4a, b, and c. The polarization curve at each current density is highly reproducible. After a first charge, the open-circuit voltage (OCV) of the cell is ~ 1.6 - 1.7 V. The cell voltage responds with the applied charge-discharge currents. The higher the current density applied, the higher the charge voltage and the lower the discharge voltage. The polarization behavior of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell could be due to the high thickness of the Na-SSE pellet and the relatively low ionic conductivity of the solid electrolyte. This issue can be overcome by the development of highconductivity Na-SSE material and thin Na-SSE membranes. Figure S3 provides a cyclic voltammetry (CV) profile of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell. The anodic and cathodic peaks in the CV are consistent with the charge-discharge profiles in Figure 4a-c. Figure 4d shows the consecutive discharge-charge curves of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell at an operating current density of 1.0 mA cm -2 (the first 25 cycles and the last 25 cycles are presented here). The discharge-charge profiles show consistent shape throughout the 100 cycles. The voltage polarization did not increase significantly with the continuous cycling, indicating that the NaSSE is compatible with the catholyte and the anolyte. Figure S4 shows the electrochemical impedance spectroscopy (EIS) of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell at the first cycle and the 100th cycle. Impedance behaviour for an aqueous battery can generally be expressed with an equivalent circuit shown in Figure S5. Accordingly, the bulk impedance RL can be determined from the real axis value at the high frequency intercept in Figure S4. The impedance of charge transfer Rct can be derived from the semicircle in Figure S4. The Rct values are, respectively, 151 and 157 Ω before and after cycling, whereas the RL values are, respectively, 130 and 133 Ω before and after cycling. There was no significant change in both the Rct and the RL of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell after 100 cycles, implying good compatibility/stability of the components during cell cycling.

Full charge 0.012

ACS Paragon Plus Environment

3

ACS Applied Energy Materials 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

Page 4 of 7

Figure 4. (a, b, and c) Voltage profiles of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell operated at (a) 0.5 mA cm-2, (b) 1.0 mA cm-2, and (c) 2.0 mA cm-2. (d) Consecutive charge-discharge profiles of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell at 1.0 mA cm-2 (the first 25 cycles and the last 25 cycles are presented here). (e) Discharge capacity and Coulombic efficiency versus cycle number of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell. Figure 4e shows the discharge capacity and Coulombic efficiency as a function of the cycle number of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell. Over the 100 cycles, the cell delivered a consistent discharge capacity. The capacity loss throughout the 100 cycles was not significant, revealing that chemical crossover of the hydroquinone liquid electrode did not occur during the cell operation. The Coulombic efficiency of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell also maintained at a constant value over the 100 cycles. Figure S6 shows the SEM images of the surface of Zn negative electrode before and after cycling. There was no significant passivation on the surface of Zn plate after 100 cycles. The results presented above successfully validate a mediator-ion Zn-quinone battery concept. However, both practical and fundamental aspects of this novel battery system/concept need a comprehensive study in the future. The solid-electrolyte Zn-quinone cell delivered a reasonable power density (3.0 mW cm-2) at ambient temperatures. However, due to the high thickness of the Na-SSE pellet and the relatively low ionic conductivity of the solid electrolyte, the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell can only be operated at relatively low current densities at present. With an enhancement in the ionic conductivity of the Na-SSE and the development of thin Na-SSE membrane,31,32 the Zn(base) ǁ NaSSE ǁ HQ(acid) cell is expected to deliver high power densities under higher operating current densities. Our major efforts in the future will focus on (1) optimization of the cycling performances of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell in collaboration with solid-electrolyte membrane researchers; (2) fundamental investigation of the overpotentials of the hydroquinone positive electrodes under cell-operating conditions and deep-insight into the ohmic drop across the NaSSE; (3) Rigorous investigation of the cycling stability of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell including the possible passivation of the Na-SSE and the possible degradation/regeneration behaviour of the carbon electrode matrix;

(4) The changes in the electrode composition upon cycling of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell, (5) Improvement in the power density and the energy density of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell by optimizing the cell components and the cell configuration. In summary, we have demonstrated a Zn(solid)Hydroquinone(liquid) battery study with a “mediator-ion solid electrolyte” strategy. Under an acidic condition, the reduction potential of hydroquinone is ~ 0.6 V higher than that under an alkaline condition. Therefore, an acidic positive electrode chemistry is preferred when employing quinone as a positive electrode. On the other hand, to couple with the quinone positive electrode, zinc negative electrode is one the best options. However, the Zn negative electrode has to be operated under alkaline conditions due to it intrinsic corrosion problem under acidic environments. Therefore, under the traditional battery operating principal with a porous separator,33,34 it is impossible to establish a high-voltage Zn-quinone energy storage system. In this study, the Zn(base) ǁ Na-SSE ǁ HQ(acid) battery was operated with an alkaline anolyte and an acidic catholyte, which are separated by a solid-state electrolyte. The redox reactions at the negative electrode and positive electrode are sustained and ionically linked by a shuttling of mediator-ions (Na+-ions) in the solid electrolyte. The transporting Na+-ions play an ionic mediator role to balance the charge transfer between the negative electrode and positive electrode. An alkaline chemistry of zinc negative electrode and an acidic chemistry of hydroquinone positive electrode result in a high-voltage Zn(base)Hydroquinone(acid) battery system. In addition to the Zn(base) ǁ Na-SSE ǁ HQ(acid) battery system presented in this study, the “mediator-ion” strategy offers a versatile approach for the development of a broad range of new battery systems with a vast range of low-cost, high energy redox couples.

ACS Paragon Plus Environment

4

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 Energy Materials

ASSOCIATED CONTENT Supporting Information Experimental Section; Photographs of the samples collected for ultraviolet-visible spectroscopy (UV-Vis) experiments (Figure S1); UV-Vis calibration curve for the hydroquinone solutions (Figure S2); A cyclic voltammetry (CV) profile of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell (Figure S3); Electrochemical impedance spectroscopy (EIS) Nyquist plots of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell at the 1st cycle and after 100 cycles (Figure S4); Equivalent circuit of the Zn(base) ǁ Na-SSE ǁ HQ(acid) cell (Figure S5); SEM images of a fresh Zn negative electrode and a Zn negative electrode after 100 cycles (Figure S6).

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

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT This work was supported by the U.S. Department of Energy, Office of Basic Energy Sciences, Division of Materials Science and Engineering under award number DE-SC0005397.

REFERENCES (1) Al-Falahi, M. D. A.; Jayasinghe, S. D. G.; Enshaei, H. A Review on Recent Size Optimization Methodologies for Standalone Solar and Wind Hybrid Renewable Energy System. Energ. Convers. Manage. 2017, 143, 252-274. (2) Bird, L.; Lew, D.; Milligan, M.; Carlini, E. M.; Estanqueiro, A.; Flynn, D.; Gomez-Lazaro, E.; Holttinen, H.; Menemenlis, N.; Orths, A.; Eriksen, P. B.; Smith, J. C.; Soder, L.; Sorensen, P.; Altiparmakis, A.; Yasuda, Y.; Miller, J. Wind and Solar Energy Curtailment: A Review of International Experience. Renew. Sust. Energ. Rev. 2016, 65, 577-586. (3) Dunn, B.; Kamath, H.; Tarascon, J. M. Electrical Energy Storage for the Grid: A Battery of Choices. Science 2011, 334, 928-935. (4) Yang, Z. G.; Zhang, J. L.; Kintner-Meyer, M. C. W.; Lu, X. C.; Choi, D. W.; Lemmon, J. P.; Liu, J. Electrochemical Energy Storage for Green Grid. Chem. Rev. 2011, 111, 35773613. (5) Sun, T.; Li, Z. J.; Wang, H. G.; Bao, D.; Meng, F. L.; Zhang, X. B. A Biodegradable Polydopamine-Derived Electrode Material for High-Capacity and Long-Life Lithium-Ion and Sodium-Ion Batteries. Angew. Chem., Int. Edit. 2016, 55, 1066210666. (6) Wang, H. G.; Yuan, S.; Si, Z. J.; Zhang, X. B. Multiring Aromatic Carbonyl Compounds Enabling High Capacity and Stable Performance of Sodium-Organic Batteries. Energ. Environ. Sci. 2015, 8, 3160-3165. (7) Wang, H. G.; Yuan, S.; Ma, D. L.; Zhang, X. B.; Yan, J. M. Electrospun Materials for Lithium and Sodium Rechargeable Batteries: from Structure Evolution to Electrochemical Performance. Energ. Environ. Sci. 2015, 8, 1660-1681. (8) Yuan, S.; Liu, Y. B.; Xu, D.; Ma, D. L.; Wang, S.; Yang, X. H.; Cao, Z. Y.; Zhang, X. B. Pure Single-Crystalline Na1.1V3O7.9 Nanobelts as Superior Cathode Materials for Rechargeable Sodium-Ion Batteries. Adv. Sci. 2015, 2, 1400018. (9) Yuan, S.; Huang, X. L.; Ma, D. L.; Wang, H. G.; Meng, F. Z.; Zhang, X. B. Engraving Copper Foil to Give Large-Scale Binder-Free Porous CuO Arrays for a High-Performance SodiumIon Battery Anode. Adv. Mater. 2014, 26, 2273-2279.

(10) Huang, X. L.; Xu, D.; Yuan, S.; Ma, D. L.; Wang, S.; Zheng, H. Y.; Zhang, X. B. Dendritic Ni-P-Coated Melamine Foam for a Lightweight, Low-Cost, and Amphipathic ThreeDimensional Current Collector for Binder-Free Electrodes. Adv. Mater. 2014, 26, 7264-7270. (11) Wang, H. G.; Yuan, S.; Ma, D. L.; Huang, X. L.; Meng, F. L.; Zhang, X. B. Tailored Aromatic Carbonyl Derivative Polyimides for High-Power and Long-Cycle Sodium-Organic Batteries. Adv. Energy Mater. 2014, 4, 1301651. (12) Liang, Y. L.; Tao, Z. L.; Chen, J. Organic Electrode Materials for Rechargeable Lithium Batteries. Adv. Energy Mater. 2012, 2, 742-769. (13) Xie, J.; Zhang, Q. C. Recent Progress in Rechargeable Lithium Batteries with Organic Materials as Promising Electrodes. J. Mater. Chem. A 2016, 4, 7091-7106. (14) Scott, D. T.; McKnight, D. M.; Blunt-Harris, E. L.; Kolesar, S. E.; Lovley, D. R. Quinone Moieties Act as Electron Acceptors in the Reduction of Humic Substances by HumicsReducing Microorganisms. Environ. Sci. Technol. 1998, 32, 29842989. (15) Steinberg-Yfrach, G.; Rigaud, J. L.; Durantini, E. N.; Moore, A. L.; Gust, D.; Moore, T. A. Light-driven Production of ATP Catalysed by F0F1-ATP Synthase in an Artificial Photosynthetic Membrane. Nature 1998, 392, 479-482. (16) Park, M.; Shin, D. S.; Ryu, J.; Choi, M.; Park, N.; Hong, S. Y.; Cho, J. Organic-Catholyte-Containing Flexible Rechargeable Lithium Batteries. Adv. Mater. 2015, 27, 51415146. (17) Huskinson, B.; Marshak, M. P.; Suh, C.; Er, S.; Gerhardt, M. R.; Galvin, C. J.; Chen, X. D.; Aspuru-Guzik, A.; Gordon, R. G.; Aziz, M. J. A Metal-free Organic-inorganic Aqueous Flow Battery. Nature 2014, 505, 195-198. (18) Wass, J. R. T. J.; Ahlberg, E.; Panas, I.; Schiffrin, D. J. Quantum Chemical Modeling of the Reduction of Quinones. J. Phys. Chem. A 2006, 110, 2005-2020. (19) Nikitina, V. A.; Nazmutdinov, R. R.; Tsirlina, G. A. Quinones Electrochemistry in Room-Temperature Ionic Liquids. J. Phys. Chem. B 2011, 115, 668-677. (20) Walczak, M. M.; Dryer, D. A.; Jacobson, D. D.; Foss, M. G.; Flynn, N. T. pH-dependent Redox Couple: Illustrating the Nernst Equation Using Cyclic Voltammetry. J. Chem. Educ. 1997, 74, 1195-1197. (21) Licht, S.; Yu, X. W. An Alkaline Periodate Cathode and Its Unusual Solubility Behavior in KOH. Electrochem. Solid. St. Lett. 2007, 10, A36-A39. (22) Yu, X. W.; Licht, S. Zirconia Coating Stabilized SuperIron Alkaline Cathodes. J. Power Sources 2007, 173, 1012-1016. (23) Yu, X. W.; Licht, S. Advances in Fe(VI) Charge Storage Part II. Reversible Alkaline Super-iron Batteries and Nonaqueous Super-iron Batteries. J. Power Sources 2007, 171, 1010-1022. (24) Yu, X. W.; Lich, S. Advances in Fe(VI) Charge Storage Part I. Primary Alkaline Super-iron Batteries. J. Power Sources 2007, 171, 966-980. (25) Kim, J. K.; Lee, E.; Kim, H.; Johnson, C.; Cho, J.; Kim, Y. Rechargeable Seawater Battery and Its Electrochemical Mechanism. Chemelectrochem 2015, 2, 328-332. (26) Moldovan, Z.; Popa, D. E.; David, I. G.; Buleandra, M.; Badea, I. A. A Derivative Spectrometric Method for Hydroquinone Determination in the Presence of Kojic Acid, Glycolic Acid, and Ascorbic Acid. J. Spectrosc. 2017, 2017, 6929520. (27) Lu, J. M.; Rosokha, S. V.; Neretin, I. S.; Kochi, J. K. Quinones as Electron Acceptors. X-ray Structures, Spectral (EPR, UV-vis) Characteristics and Electron-transfer Reactivities of Their Reduced Anion Radicals as Separated vs Contact Ion Pairs. J. Am. Chem. Soc. 2006, 128, 16708-16719.

ACS Paragon Plus Environment

5

ACS Applied Energy Materials 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

Page 6 of 7

(28) Barone, V.; Improta, R.; Morelli, G.; Santoro, F. UV-vis Spectra of p-benzoquinone Anion Radical in Solution by a TDDFT/PCM Approach. Theor. Chem. Acc. 2007, 118, 143-148. (29) Silva, C. H. B.; Ferreira, D. C.; Ando, R. A.; Temperini, M. L. A. Aniline-1,4-benzoquinone as a Model System for the Characterization of Products from Aniline Oligomerization in Low Acidic Media. Chem. Phys. Lett. 2012, 551, 130-133. (30) Gulaboski, R.; Bogeski, I.; Mirceski, V.; Saul, S.; Pasieka, B.; Haeri, H. H.; Stefova, M.; Stanoeva, J. P.; Mitrev, S.; Hoth, M.; Kappl, R. Hydroxylated Derivatives of Dimethoxy-1,4benzoquinone as Redox Switchable Earth-alkaline Metal Ligands and Radical Scavengers. Sci. Rep. - UK 2013, 3, 1865. (31) Goodenough, J. B.; Singh, P. Review-Solid Electrolytes in Rechargeable Electrochemical Cells. J. Electrochem. Soc. 2015, 162, A2387-A2392. (32) Bachman, J. C.; Muy, S.; Grimaud, A.; Chang, H. H.; Pour, N.; Lux, S. F.; Paschos, O.; Maglia, F.; Lupart, S.; Lamp, P.; Giordano, L.; Shao-Horn, Y. Inorganic Solid-State Electrolytes for Lithium Batteries: Mechanisms and Properties Governing Ion Conduction. Chem. Rev. 2016, 116, 140-162. (33) Posada, J. O. G.; Rennie, A. J. R.; Villar, S. P.; Martins, V. L.; Marinaccio, J.; Barnes, A.; Glover, C. F.; Worsley, D. A.; Hall, P. J. Aqueous Batteries as Grid Scale Energy Storage Solutions. Renew. Sust. Energ. Rev. 2017, 68, 1174-1182. (34) Huggins, R. A. Review-A New Class of High Rate, Long Cycle Life, Aqueous Electrolyte Battery Electrodes. J. Electrochem. Soc. 2017, 164, A5031-A5036.

ACS Paragon Plus Environment

6

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 Energy Materials

Table of Contents

ACS Paragon Plus Environment

7