<<

Subscriber access provided by UNIV OF CALIFORNIA SAN DIEGO LIBRARIES C: Energy Conversion and Storage; Energy and Charge Transport Intercalation and Conversion Reactions of Nanosized #- MnO2 in the Secondary Zn/MnO2 Alkaline Battery Joon Kyo Seo, JaeWook Shin, Hyeseung Chung, Po Yu Meng, Xuefeng Wang, and Ying Shirley Meng J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b11685 • Publication Date (Web): 03 May 2018 Downloaded from http://pubs.acs.org on May 15, 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.

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 33 The Journal of Physical Chemistry

1 2 3 4 5 6 7 Intercalation and Conversion Reactions of 8 9 10 11 Nanosized βMnO Cathode in the Secondary 12 2 13 14 15 16 Zn/MnO2 Alkaline Battery 17 18 19 20 Joon Kyo Seo,†,‡ JaeWook Shin,‡,§ Hyeseung Chung,‡ Po Yu Meng,‡ Xuefeng Wang,‡ and Y. 21 22 *,†,‡ 23 Shirley Meng 24 25 26 †Materials Science and Engineering, University of California San Diego, La Jolla, CA 92093, 27 USA 28 29 ‡Department of NanoEngineering, University of California San Diego, La Jolla, CA 92093, USA 30 31 *Email: [email protected]. Tel: +1 (858) 8224247. 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49

50 51 52 §Present address: Advanced Battery Center, KAIST Institute for NanoCentury, Korea Advanced 53 54 Institute of Science and Technology (KAIST), 291 Daehakro, Yuseonggu, Daejeon, 34141, 55 Republic of Korea 56 57 58 59 60 ACS Paragon Plus Environment 1 The Journal of Physical Chemistry Page 2 of 33

1 2 3 ABSTRACT 4 5 6 7 This work reports rechargeable Zn/βMnO2 alkaline batteries as promising stationary energy 8 9 storage. Unlike commercial alkaline batteries with poor cyclic performance, the nanosized β 10 11 12 MnO2 cathode in the mixture of LiOH and KOH enables rechargeable reactions with 13 14 high capacity. To unveil the underlying reaction mechanisms of nanosized βMnO2, we combine 15 16 thermodynamic frameworks with experimental characterization including electrochemistry, X 17 18 19 ray diffraction, and Xray photoelectron spectroscopy. The results demonstrate a series of proton 20 21 intercalation reaction (βMnO2 → γMnOOH) and twophase conversion reactions (γMnOOH 22 23 → Mn(OH)2 → λMnO2) during the first cycle, and Li and H cointercalation in the host 24 25 structure of λMnO spinel during the hundredth cycle. It is remarkable that the addition of Bi O 26 2 2 3 27 28 in the nanosized βMnO2 cathode exhibits the outstanding capacity. After one hundred 29 30 discharging, the battery demonstrates a capacity of 316 mAh g1. Our findings can serve in the 31 32 tailored cathode design in high capacity and rechargeable Zn/βMnO alkaline batteries. 33 2 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 Paragon Plus Environment 2 Page 3 of 33 The Journal of Physical Chemistry

1 2 3 INTRODUCTION 4 5 6 Ever since Lewis Urry invented a primary Zn/MnO alkaline battery back in the late 1950’s, it 7 2 8 9 has been widely adopted as an energy source for lowpower electronics. Recently, the Zn/MnO2 10 11 aqueous battery is gaining attention as a rechargeable for smart grid technology.1 12 13 3 Compared to a Li battery, a Zn/MnO2 aqueous battery has several advantages such as low 14 15 16 material cost, better safety, and nontoxicity. To illustrate, the current rechargeable Zn/MnO2 17 2 18 aqueous battery technology costs under $90 per kWh, whereas the cost for Liion batteries are 19 3 20 unlikely to fall below $200 per kWh. In addition, Zn/MnO2 aqueous battery operates with 21 22 waterbased , furnishing a higher degree of safety making them an attractive option 23 24 25 for gridscale usage in residential homes and dense urban areas. 26 27 28 MnO2 exists in multiple crystallographic structures including α, β, γ, δ, λ, and RMnO2. 29 30 These MnO2 polymorphs are defined by different connection of the basic structural unit, MnO6 31 32 45 33 octahedron, and various types of tunnels within the crystal structure. Electrochemical reactions 34 35 vary in Zn/MnO2 aqueous batteries dependent on the phase of MnO2 and electrolyte used, 36 37 however, the most common discharge mechanism of MnO2 involves a homogeneous reaction 38 39 and followed by a heterogeneous reaction.68 Once the discharge initiates, protons intercalate into 40 41 42 tunnels of MnO2 through the homogeneous reaction without MnO bond breakage, however the 43 44 bonds begin to break in the heterogeneous reaction as the discharge proceeds further leading to 45 46 eventual deformation of the lattice. During the homogeneous reaction, the voltage changes 47 48 49 rapidly, however the heterogeneous reaction manifests the voltage plateaus due to twophase 50 51 reactions. 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment 3 The Journal of Physical Chemistry Page 4 of 33

1 2 3 Recently, many studies have reported the development of rechargeable Zn/MnO2 aqueous 4 5 911 6 batteries which utilize mild acidic electrolytes. Lee et al. describe that a Zn/αMnO2 system in 7 2+ 2+ 8 ZnSO4 electrolyte involves Zn/Zn redox at the and αMnO2/Mn at the cathode. An 9 10 additional reversible phase of Zn4(OH)6(SO4)‧5H2O ( hydroxide sulfate) precipitates on the 11 12 cathode’s surface due to Zn2+ and SO 2+ in an aqueous electrolyte.10 Pan et al. report that the 13 4 14 15 rechargeability of Zn/αMnO2 aqueous battery is significantly improved by adding MnSO4 salt in 16 11 17 the ZnSO4 electrolyte. They describe the same redox reactions proposed by Lee et al. with an 18 19 undefined hydration number for the zinc hydroxide sulfate. However, the addition of the MnSO 20 4 21 2+ 22 salt in the electrolyte suppresses the dissolution of Mn and enables longevity of over 5000 23 24 cycles. 25 26 27 In addition to the acidic solutions, the basic solutions have been investigated to understand the 28 29 rechargeability of Zn/MnO aqueous batteries. Gallaway et al. probed the phase transformation 30 2 31 12 32 of the cathode/anode in a commercial alkaline ‘AA’ cell. They observed ZnMn2O4 (hetaerolite) 33 34 forming in the cathode, implying that it is a side product of the chemical reaction between 35 36 2 2 2 MnOOH and Zn(OH)4 . The Zn(OH)4 (zincate ion) is generated from Zn(OH)4 /Zn redox at 37 38 39 the Zn anode which then migrates to the cathode where MnOOH is formed after MnO2 is 40 41 reduced. The ZnMn2O4 phase produced is electrochemically inactive phase and degrades the 42 43 cycle performance of Zn/MnO2 aqueous batteries. Turney et al. reported that the formation of 44 45 Zn(OH) 2 is favorable at the Zn anode where OH species are concentrated inducing the 46 4 47 3 48 production of ZnMn2O4. In another study, Hertzberg et al. found that the combination of LiOH 49 50 and KOH as electrolytes in a Bi species incorporated Zn/βMnO2 alkaline battery improved the 51 52 rechargeability demonstrating a reversible singleelectron reaction for 60 cycles.13 They 53 54 55 proposed that a reversible reaction occurs during the cycling between the reduced phases 56 57 58 59 60 ACS Paragon Plus Environment 4 Page 5 of 33 The Journal of Physical Chemistry

1 2 3 (Mn(OH)2 and LiMn2O4) and the oxidized phase (δMnO2). This reaction mechanism, however, 4 5 6 is contentious due to δMnO2 being measured solely during the first charged state and not in any 7 8 of the subsequent cycles. 9 10 11 Optimizing the utilization of two electrons in MnO2 is an essential for highenergy stationary 12 13 storage. Many studies have demonstrated that the capacity increases using nanostructured 14 15 5, 1415 16 materials. Chen et al. reported that nanosized γMnO2 shows almost twotimes higher 17 14 18 discharge capacity than bulk γMnO2. They suggested that broad surface area in nanoscale 19 20 MnO2 enlarges contact between MnO2 active materials and an electrolyte which lowers internal 21 22 resistance and facilitates proton diffusion. In order to utilize large capacity and improve the 23 24 25 reversibility of reactions in alkaline solutions, we synthesized nanoscale βMnO2 and assembled 26 27 the Zn/βMnO2 alkaline battery with the LiOH and KOH electrolytes. The discharge capacity 28 29 shows 225 mAh g1 at the hundredth cycle, which is higher than the capacity of the battery made 30 31 1 32 of commercial βMnO2 by 113 mAh g . Furthermore, we studied the phase transformation of 33 34 nanosized βMnO2 in Zn/βMnO2 alkaline battery. The result indicates that the proton 35 36 intercalation reaction is followed by twophase conversion reactions at the first cycle, however, a 37 38 39 different reaction mechanism is observed at the hundredth cycle. Li and H coinserted λMnO2 40 41 spinel and λMnO2 spinel are formed at the hundredth discharged and charged state, respectively. 42 43 Lastly, we incorporated βMnO2 cathode with 4% mole fraction of Bi2O3 to enhance the 44 45 electrochemical performance. The discharge capacity is 316 mAh g1 at the hundredth cycle. Bi 46 47 2 48 2O3 additive reacts with Zn(OH)4 , which alleviates the capacity loss driven by the formation of 49 50 irreversible ZnMn2O4. 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment 5 The Journal of Physical Chemistry Page 6 of 33

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 EXPERIMENTAL METHODS 17 18 19 βMnO2 synthesis. βMnO2 was synthesized by the multiple heating steps in order to confirm 20 21 4+ Mn , to remove NOx gases completely and crystalize nanosized βMnO2. The thermal 22 23 24 decomposition process begins with dissolving 10 g of Mn(NO3)2·4H2O in 16 mL deionized 25 26 (D.I.) water. The solution was stirred with a magnetic bar in a roundbottom flask while being 27 28 heated in the siliconoil bath to increase the reaction temperature to 140 °C. Once the 29 30 temperature reached 140 °C, the reaction was held at that temperature until the initially clear 31 32 4+ 33 solution developed a black color with high viscosity indicating the oxidation to the Mn state. 34 35 Then, the round flask was transferred to a preheated vacuum oven and kept at 125 °C overnight 36 37 to facilitate complete drying of the MnO2 product and to eliminate excess NOx gases. The sample 38 39 40 was removed from the vacuum oven after 12 h and transferred to a preheated box furnace heated 41 42 to 125 °C. The temperature of the oven was programmed to increase up to 325 °C with a step of 43 44 5 °C/min, then to hold the temperature at 325 °C for 5 h to crystalize nanosized βMnO2. After 45 46 47 natural cooling, the solid sample was ground with a mortar and a pestle to turn it into fine 48 49 powder. 50 51 Electrochemical characterization. The working is composed of the βMnO2 active 52 53 material (synthesized βMnO , commercial βMnO , or synthesized βMnO with 4 % mole 54 2 2 2 55 56 fraction of Bi2O3), carbon black SuperP, and polytetrafluoroethylene (PTFE) with a 60:30:10 57 58 59 60 ACS Paragon Plus Environment 6 Page 7 of 33 The Journal of Physical Chemistry

1 2 3 2 weight ratio. The typical mass loading of βMnO2 is 11.4 mg cm . The electrode was implanted 4 5 6 onto a Ni foam current collector and pressed using a hydraulic press with a pressure of eight 7 8 tons. The pressed electrode was then covered with one layer of polyvinylidene chloride, two 9 10 layers of nonwoven separator (FS 22145 from Freudenberg LLC) and a Zn plate counter 11 12 electrode. They were sandwiched between two acrylic plates, which were held together tightly 13 14 15 with screws. After an assembly, the battery was immersed in a 20 mL beaker with 10.5 mL of 1 16 17 M KOH and 3 M LiOH electrolyte. This molar ratio of 1 : 3 = KOH : LiOH was optimized for 18 19 alkaline batteries in previous report.13 Parafilm was used to seal the beaker. All batteries in this 20 21 22 study were tested at C/10 current rate for the first cycle and C/5 for the following cycles. 23 24 Considering the rate capability tests for synthesized βMnO2 with/without Bi2O3, the 25 26 discharge/charge process was carried out over the course 20 cycles for each stage except for the 27 28 29 last stage (9 cycles). Constant current rates were calculated using the theoretical twoelectron 30 1 31 capacity of MnO2 which is 616 mAh g . The voltage range used is 0.50 < V < 1.80. The constant 32 33 voltage of 1.80 V was applied at the end of constant current charging until the current decreased 34 35 to 10% of the current rate. The batteries were cycled using an Arbin battery cycler at room 36 37 38 temperature. In addition, the electrochemical stability window of electrolyte was measured using 39 40 a Solartron 1260 Impedance Analyzer. The linear sweep voltammetry was taken on Ni foam 41 42 compared against Hg/HgO reference electrode in 1M KOH and 3M LiOH electrolyte 43 44 45 at a scan rate of 1 mV/s. 46 47 In house Xray diffraction (XRD). The inhouse XRD samples were synthesized and 48 49 commercial βMnO2 powder, and cycled βMnO2 electrodes. Cycled electrodes were collected at 50 51 52 1.25 V, 1.05 V, 0.50 V and 1.80 V during the first cycle of synthesized βMnO2 in Zn/βMnO2 53 54 alkaline batteries. The collected electrodes were washed with running D.I. water and soaked in 55 56 57 58 59 60 ACS Paragon Plus Environment 7 The Journal of Physical Chemistry Page 8 of 33

1 2 3 the D.I. water overnight. After washing, the electrodes were dried in 60 ‧ for 3 hours in the 4 5 6 oven. All the XRD data were collected at an ambient temperature on a Bruker D8 Advance 7 8 diffractometer at 40 kV, 40 mA for Cu Kα (λ = 1.5418 Å), with a scan speed of 5 s per step, a 9 10 step size of 0.02° in 2θ, and a 2θ range of 10–80°. 11 12 Synchrotron Xray diffraction (SXRD). Cycled electrodes of synthesized βMnO collected at 13 2 14 15 the hundredth discharge/charged state were selected for SXRD analysis. The cycled electrodes 16 17 were washed with running D.I. water and soaked in the D.I. water overnight. After washing, 18 19 electrodes were dried at 60 ‧ for 3 hours in the oven. SXRD spectra were acquired at the 20 21 22 Advanced Photon Source (APS) at Argonne National Laboratory (ANL) on beamline 11BM (λ= 23 24 0.4145 Å). The beamline uses a sagittal focused Xray beam with a high precision diffractometer 25 26 circle and perfect Si(111) crystal analyzer detection for high sensitivity and resolution. SXRD 27 28 29 spectra were plotted based on λ= 1.5418 Å to compare with the inhouse XRD spectra. 30 31 Xray photoelectron spectroscopy (XPS). XPS was performed using a Kratos AXIS Supra at 32 33 the Laboratory for Electron and Xray Instrumentation. Cycled samples were prepared by 34 35 washing with running D.I. water and soaking the electrode in the D.I. water overnight, and 36 37 38 drying in 60 ‧ for 3 h in the oven. The XPS was operated using an Al anode source running at 39 40 15 kV. All XPS measurements were collected using a 300 m by 700 m spot size without using 41 42 a charge neutralizer during acquisition. A step size of 0.3eV was used for survey scan and 0.1 eV 43 44 3 45 was for highresolution scans. All the spectra were calibrated with carbon 1s sp at 284.8 eV. 46 47 Scanning electron microscope (SEM) and energy dispersive Xray spectroscopy (EDS). The 48 49 SEM/EDS samples selected were synthesized and commercial βMnO2 powders, Bi2O3 and 50 51 52 immersed Bi2O3 powder in 6M Zn nitrate solution. The sample of immersed Bi2O3 powder was 53 54 prepared by washing with running D.I. water, soaking the powder in the D.I. water overnight, 55 56 57 58 59 60 ACS Paragon Plus Environment 8 Page 9 of 33 The Journal of Physical Chemistry

1 2 3 and drying in 60 ‧ for 3 h in the oven. All powder samples were loaded onto a SEM holder 4 5 6 using a carbon adhesive tape. The images were obtained using a 15 kV energy source using the 7 8 FEI/Phillips XL30 ESEM. 9 10 Computational details. First principles calculations were conducted based on spinpolarized 11 12 GGA+U 1617 using the Perdew–Burke–Ernzerhof of exchange and correlation functionals.18 A 13 14 1920 15 planewave basis set and the projectoraugmented wave (PAW) method were used as 16 17 parameterized in the Vienna ab initio simulation package (VASP).21 A gamma point mesh is 18 19 performed with 7 × 7 × 7 kpoints for λMnO . All the atoms were fully relaxed to calculate the 20 2 21 22 optimized structure with a cutoff energy of 1.3 times the maximum cutoff specified by the 23 24 pseudopotentials on a planewave basis set. The scheme proposed by Persson et al22 was adopted 25 26 in order to obtain an accurate standard Gibbs free energy of formation. 27 28 29 30 31 32 RESULTS AND DISCUSSION 33 34 35 The influence of a particle size. The powder of synthesized and commercial βMnO2 are 36 37 measured by inhouse Xray diffractometer as shown in figure 1(a). The peak broadenings are 38 39 observed in synthesized βMnO , however, the peak positions match the reference βMnO . The 40 2 2 41 42 XPS result confirms Mn 3s multiplet splitting with E = 4.62 eV in figure 1(b); this value agrees 43 23 44 well with E = 4.65 eV found in commercial βMnO2 (Alfa Aesar) as shown in figure S1. The 45 46 SEM image shows agglomerated secondary particles from the synthesized βMnO2 (the red color 47 48 49 box) and chunky particles with small parasites from the commercial βMnO2 (the orange color 50 51 box) in figure 1(c). In the synthesized βMnO2 the primary particle is in a size range of less than 52 53 50 nm. The Brunauer–Emmett–Teller (BET) surface analysis conducted with the nitrogen 54 55 56 sorption isotherms at 196 °C by Autosorb iQ (Quantachrome) indicates ~21 times larger 57 58 59 60 ACS Paragon Plus Environment 9 The Journal of Physical Chemistry Page 10 of 33

1 2 3 surface area for the synthesized βMnO2 than the commercial βMnO2 in (figure S2). The cyclic 4 5 6 performance of Zn/βMnO2 alkaline batteries is tested in 1 M KOH and 3 M LiOH containing 7 8 electrolyte as shown in figure 1(d). The first discharge capacity of synthesized βMnO2 is 574 9 1 1 10 mAh g (approximately 93% of the MnO2’s theoretical capacity of 616 mAh g ). The 11 12 commercial βMnO produces 464 mAh g1 during the first discharge (approximately 75% of the 13 2 14 15 theoretical capacity). The discharge capacity of synthesized βMnO2 at the hundredth cycle is 16 1 1 17 225 mAh g , which is higher than that of commercial βMnO2, 112 mAh g . The coulombic 18 19 efficiency (CE) defined by the percentage of charge capacity over discharge capacity is less than 20 21 22 100 %. This is due to the different amount of additional capacity produced by H2 and O2 gases, 23 24 which are generated within the battery’s operating voltage (figure S3). The gases are released 25 26 from the customized beaker cell and its configuration is demonstrated in the experimental. 27 28 29 Another factor affecting the CE is the size of βMnO2. We estimate that bulk βMnO2 materials, 30 31 which have relatively small surfaceareatovolume ratio do not promote the formation of 32 33 irreversible ZnMn2O4 phase, resulting in relatively higher CE than the synthesized βMnO2 34 35 nanomaterials. In case of synthesized βMnO nanomaterials, it improves the specific capacity 36 2 37 38 during cycles, however it promotes the formation of ZnMn2O4 phase at the same time due to its 39 40 large surfaceareatovolume ratio. 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment 10 Page 11 of 33 The Journal of Physical Chemistry

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 Figure 1. (a) Inhouse XRD patterns for synthesized and commercial βMnO2 (Alfa Aesar); (b) 38 39 40 Mn 3s splitting in the XPS result of synthesized βMnO2; (c) SEM images for synthesized (red 41 42 box) and commercial βMnO2 (orange box); (d) Cyclic performance for synthesized βMnO2 and 43 44 commercial βMnO2 at C/10 current rate for the first cycle and C/5 for the following cycles. 45 46 47 The gap between the discharge capacities from two batteries is most likely due to the difference 48 49 50 between particle sizes of the active materials. The nanoscale βMnO2 particles have broader 51 52 electrochemical surface area than the bulk βMnO2 particles, which increases the effective 53 54 contact points with the electrolyte, lowering internal resistance and shortening diffusion 55 56 57 58 59 60 ACS Paragon Plus Environment 11 The Journal of Physical Chemistry Page 12 of 33

1 2 3 5, 15 pathways. In addition, insufficiently coordinated surface atoms of the synthesized βMnO2 4 5 6 nanoparticles are expected to facilitate chemical adsorptions from aqueous solutions and undergo 7 8 electrochemical reactions. Han et al. reported the effect of the nanosized particle on an 9 10 adsorption energy of chemicals in aqueous media.24 It shows that the adsorption energy of 11 12 chemicals on the surface of a nanoparticle is stronger than that on the surface of a bulk material. 13 14 15 Nanoparticles hold a high ratio of undercoordinated surface atoms to coordinated atoms, thus 16 17 their chemisorption is greater than that of the bulk materials. The nanosized βMnO2 case are 18 19 expected to undergo facile electrochemical reactions once H O adsorption from aqueous 20 2 21 22 electrolytes has occurred. The reaction mechanism of synthesized βMnO2 will be discussed 23 24 further in this study. Another factor to consider regarding surface chemisorption is the cohesive 25 26 energy per atom of a nanoparticle.25 Constituent atoms in a nanoparticle are bound together with 27 28 29 weak cohesive energy, which is a less stable formation than a bulk material. Thus nanoparticles 30 31 in aqueous media are likely to bind with nearby molecules for stabilization, which in turn 32 33 stimulates electrochemical reactions. 34 35 36 37 38 39 The reaction mechanism of nanosized βMnO during the first cycle. Ex situ inhouse XRD 40 2 41 42 patterns are measured for synthesized βMnO2 during the first cycle of Zn/βMnO2 43 44 alkaline batteries (figure 2). The reaction mechanism of synthesized βMnO2 nanomaterial is 45 46 proposed as follows: 47 48 49 Electrochemical reaction (the first discharge): 50 51 52 53 βMnO2 + H2O + e → γMnOOH + OH (1) 54 55 56 γMnOOH + H2O + e → Mn(OH)2 + OH (2) 57 58 59 60 ACS Paragon Plus Environment 12 Page 13 of 33 The Journal of Physical Chemistry

1 2 3 Electrochemical reaction (the first charge): 4 5 6 Mn(OH) + 2OH → λMnO + 2H O + 2e (3) 7 2 2 2 8 9 10 1/3Mn3O4 + 4/3OH → λMnO2 + 2/3H2O + 4/3e (4) 11 12 13 Chemical reaction: 14 15 16 γMnOOH + 1/2Mn(OH)2 → 1/2Mn3O4 + H2O (5) 17 18 19 To further validate the reaction mechanism characterized by ex situ inhouse XRD patterns, the 20 21 theoretical reaction voltage (E) for each electrochemical reaction proposed in Eq. (14) is 22 23 24 calculated by applying the Nernst equation as follows: 25 26 2 27 E = E° + 2.303 ‧ RT/F ‧ pOH – (–1.119) V vs. Zn(OH)4 | Zn (6) 28 29 30 where E° is the standard reduction potential (table S1) which is obtained by the standard Gibbs 31 32 free energy of formation (table S2); R is the gas constant (8.314 J K1 mol1); T is the room 33 34 temperature (298.15 K); F is the Faraday constant (96485 C mol1); pOH is the concentration of 35 36 37 hydroxide. The theoretical value of pOH = 0.6 (pH = 14.6) is used since the electrolyte consists 38 39 of 1 M KOH and 3 M LiOH, where 4 M OH exists in total. In alkaline media, the Zn anode has 40 41 2 12, 26 a redox couple with Zn(OH)4 . The theoretical reaction voltage is therefore measured with 42 43 2 27 44 respect to the standard reduction potential of Zn(OH)4 | Zn (1.119 V vs. SHE). The ideal 45 46 voltage curve of a Zn/βMnO2 alkaline battery fabricated with synthesized βMnO2 nanomaterial 47 48 is presented in figure 2(b). The theoretical voltage profile (dotted line) matches well with the 49 50 51 experimental voltage profile, which will be discussed in detail. 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment 13 The Journal of Physical Chemistry Page 14 of 33

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 Figure 2. (a) Ex situ inhouse XRD patterns for synthesized βMnO2 nanomaterial and its 27 28 intermediate phases during the first cycle. The vertical line presents XRD patterns from ICSD 29 30 database (βMnO2: #73716, γMnOOH: #84949, Mn3O4: #643199, and λMnO2: #193445). 31 32 Detailed XRD patterns are shown in figure S4. (b) The voltage profile for the first 33 34 35 discharge/charge of a Zn/βMnO2 alkaline battery at C/10 current rate. The colored circles 36 37 represent points that synthesized βMnO2 nanomaterial and cycled βMnO2 cathode samples are 38 39 collected. The dotted line indicates the theoretical voltage profile at pH=14.6. 40 41 42 43 As mentioned above, XRD analysis shown in figure 2(a) indicates that the synthesized powder is 44 45 βMnO2. βMnO2 has a tetragonal crystal lattice where MnO octahedrons are connected in a 46 47 cornersharing and an edgesharing configurations as shown in figure 3. Void tunnels and MnO 48 49 octahedrons are positioned 1 x 1. Upon the initiation of the first discharge, protons originating 50 51 52 from H2O molecules adsorbed on the surface of nanosized βMnO2 intercalate into the void 53 54 tunnels and form γMnOOH at the end of the firstelectron reaction (Eq. (1)). The XRD pattern 55 56 57 58 59 60 ACS Paragon Plus Environment 14 Page 15 of 33 The Journal of Physical Chemistry

1 2 3 of γMnOOH is obtained during the first discharge at 1.05 V (figure 2). γMnOOH is a 4 5 6 monoclinic crystal lattice where the original 1 x 1 void tunnels in βMnO2 are occupied with 7 8 protons, resulting in a distortion in the lattice (figure 3). The XRD spectrum acquired at 1.25 V 9 10 during the early stage of proton intercalation at the first discharge shows the overall peak 11 12 intensity of βMnO decreasing, yet the phase still holds the tetragonal crystal lattice (figure 13 2 14 15 S4(b)). The crystal lattice gets increasingly distorted as the proton concentration increases and 16 17 the structure becomes monoclinic at the end of the firstelectron reaction at 1.05 V (figure 2(a) 18 19 and figure 3). Kordesch et al. proposed that the firstelectron reaction of MnO in a Zn/MnO 20 2 2 21 28 22 alkaline battery is a solidsolution intercalation reaction. This reaction is a homogeneous step 23 24 because MnO bonding in the host MnO2 is retained during proton intercalation. The theoretical 25 26 voltage for the βMnO2 | γMnOOH redox couple averages 1.28 V upon proton intercalation, 27 28 29 which is located within the experimental voltage curve of the firstelectron reaction (figure 2(b)). 30 31 32 In addition the XRD results indicate Mn3O4 species at the end of first discharge (0.50 V) in 33 34 figure 2(a). Mn3O4 has the tetrahedral crystal lattice where cornersharing MnO tetrahedrons 35 36 and edgesharing MnO octahedrons are interconnected (figure 3). It is noted that Mn3O4 is 37 38 39 unlikely to reduce from γMnOOH by electrochemical reaction because Mn3O4 has incorporating 40 3+ 2+ 41 two Mn and only one Mn , which would indicate the involvement of only 1/3 electron. The 42 43 first discharge is a twoelectron reaction as its experimental capacity is 574 mAh g1, which is 44 45 approximately 93% of theoretical capacity of MnO ’s two electrons. Therefore Mn2+ species 46 2 47 48 should be formed at the end of the first discharge. It has been reported that MnOOH is reduced to 49 2930 50 Mn(OH)2 by a twophase conversion reaction (Eq. (2)) and Mn3O4 is formed by a parallel 51 52 chemical reaction (Eq. (5))12, 30 in the alkaline batteries. Mn(OH) has the trigonal crystal lattice 53 2 54 55 where the layer of edgesharing MO octahedrons and the layer of the proton are aligned along 56 57 58 59 60 ACS Paragon Plus Environment 15 The Journal of Physical Chemistry Page 16 of 33

1 2 3 the caxis as shown in figure 3. Mn3O4 is generated by the chemical reaction between MnOOH 4 5 6 and Mn(OH)2 while the βMnO2 reduces to Mn(OH)2 electrochemically. We estimate this 7 8 chemical reaction proceeds further while collecting the XRD sample for a cathode by ex situ 9 10 methods. Additional Mn3O4 begins to form after the electrochemical test stopped at the end of 11 12 first discharge (0.50 V) and its amount increases extensively until the ex situ sample is collected. 13 14 29 15 Donne et al. demonstrated that the formation of Mn3O4 is a kinetically limited reaction. They 16 17 measured Mn3O4 when experiments were tested over the extended time duration. Hertzberg et 18 19 al., however, did not detect the Mn O but Mn(OH) while conducting in situ Xray 20 3 4 2 21 13 22 diffractometer. Mn3O4 formation is alleviated in the relatively shorttime frame between 23 24 stopping the electrochemical reaction and measuring its XRD spectra. It is noted that the 25 26 theoretical voltage of γMnOOH | Mn(OH)2 redox couple demonstrates the formation of 27 28 29 Mn(OH)2 during the first discharge. The voltage is 0.93 V, which corresponds well with the 30 31 experimental voltage plateau during the secondelectron reaction (figure 2(b)). 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 Paragon Plus Environment 16 Page 17 of 33 The Journal of Physical Chemistry

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 Figure 3. Schematic illustration of the electrochemical reaction mechanism of nanosized β 42 43 MnO2 during the first cycle. 44 45 46 At the end of the first charge, λMnO2 is measured (figure 2(a)) as a result of Mn(OH)2 oxidation 47 48 (Eq. (3)). λMnO is a spinel phase with the cubic crystal lattice where MO octahedrons are 49 2 50 51 interconnected in an edgesharing manner creating a 3D network of 1 x 1 void channel (figure 3). 52 53 The XRD pattern of the cell at 1.80 V has a peak at ~19.1º, which corresponds to the (111) plane 54 55 of λMnO2 (figure S5(b)). This peak is the characteristic Bragg diffraction found in the λMnO2 56 57 58 59 60 ACS Paragon Plus Environment 17 The Journal of Physical Chemistry Page 18 of 33

1 2 3 phase exclusively among the MnO2 polymorphs (figure S5(a)). It is important to note that the 4 5 6 XRD pattern at 1.80 V is weak and broadened. We attribute this phenomenon to the formation of 7 8 nanostructured λMnO2 particles, which are less than ~50 nm (figure S6). In very 9 10 small particles, a large fraction of the atoms are located on the surface region, which undergo the 11 12 structural relaxation for thermodynamic stability. Relaxed atoms are distorted from the original 13 14 15 atomic positions of the crystal lattice, which could lead to the low intensity and peak broadening 16 17 in the XRD pattern. Gallaway et al. reported that the irreversible ZnMn2O4 phase can be 18 19 generated during the operation of primary alkaline batteries.12 In our study, however, the XRD 20 21 22 pattern at the first charged state does not indicate ZnMn2O4 phase (figure S4(e)). Also, no Zn 2p 23 24 signal is found from XPS at the first charged state (figure S7(a)). We estimate that the formation 25 26 of ZnMn2O4 is dependent on the rate of reaction. It is reported that ZnMn2O4 is not formed at 27 28 12 29 fast Crate, but generated at slow Crate during the first discharge. 30 31 Also, we investigated the formation of possible Ni hydroxide phases including βNiOOH as a 32 33 result of Ni oxidation (figure S8), however their XRD patterns were not aligned with the XRD 34 35 pattern at the first charged state (1.80 V). Considering the λMnO , the theoretical voltage of 36 2 37 38 Mn(OH)2 | λMnO2 redox couple (1.12 V) supports its formation, which is approximately the 39 40 experimental voltage plateau during the first charge shown in figure 2(b). The first charge 41 42 capacity is 353 mAh g1, which is 57 % of the theoretical capacity for the twoelectron reaction. 43 44 45 We ascribe the capacity loss to any unreacted Mn(OH)2 which dissolves in the electrolyte as OH 46 n 31 2+ 47 coordinatedcomplex or [Mn(OH)n+2] , because Mn in Mn(OH)2 is soluble in highly 48 49 concentrated basic solution. It is interesting that Mn3O4 which is generated during the first 50 51 52 discharge is not measured at 1.80 V because it is also oxidized to λMnO2 (Eq. (4)). Mn3O4 has 53 8 54 poor electrical conductivity (10 ohm‧cm) which hinders electrochemical reactions in Zn/MnO2 55 56 57 58 59 60 ACS Paragon Plus Environment 18 Page 19 of 33 The Journal of Physical Chemistry

1 2 3 12 alkaline batteries. However, Mn3O4 converts into λMnO2 in this study and we attribute it to 4 5 6 the particle size of Mn3O4. We think that our synthesized active materials remain as nanoscale 7 8 particles during phase transformations, which compensates for the intrinsic high resistivity of 9 10 Mn3O4 and facilitates electrochemical reactions. Compared to bulky Mn3O4, nanoscale Mn3O4 11 12 has advantages in terms of large electrode/electrolyte interface for electrochemical reactions, 13 14 15 superior conductivity through its ability to embed into carbon networks, and short electron/ion 16 32 17 transport distance. Recently, nanosized Mn3O4 has been reported on for displaying high 18 19 reversible capacity and enhanced CE as an anode material in Liion batteries.3234 The theoretical 20 21 22 redox potential of Mn3O4 | λMnO2 is calculated as 1.28 V, which is also close to the 23 24 experimental voltage plateau observed during the charge (figure 2(b)). 25 26 27 28 29 30 The reaction mechanism at the hundredth cycle. The ex situ SXRD pattern is measured at the 31 32 33 hundredth discharged/charged state of the synthesized βMnO2 cathode in the Zn/βMnO2 34 35 alkaline battery (figure 4(a,c)). It is interesting to note that the ZnMn2O4 phase is found, which 36 37 was not formed during first cycle. This phase exists in both discharged/charged states because of 38 39 its irreversibility in alkaline solutions. A Li and H coinserted λMnO spinel is also formed at 40 2 41 42 the hundredth discharged state as shown in figure 4(a). The 111 and 311 peaks from the 43 44 reference Li and H coinserted λMnO2 spinel align with the peaks from ex situ SXRD. The 45 46 presence of Li is confirmed at the hundredth discharge state by XPS spectra with a signal in the 47 48 49 Li 1s region shown in figure 4(b), which supports the formation of Li contained phase. We 50 51 investigated other lithiated λMnO2 spinel phases as the function of Li concentration to see if 52 53 their XRD patterns match the ex situ SXRD pattern at the hundredth discharged state. However, 54 55 56 there are no other phases that demonstrate as good a match as shown in figure S9. As the lithium 57 58 59 60 ACS Paragon Plus Environment 19 The Journal of Physical Chemistry Page 20 of 33

1 2 3 concentration in the λMnO2 spinel increases, the peak positions of XRD shift to low diffraction 4 5 6 angle (2θ) values. These values do not correspond to the ex situ SXRD pattern at the hundredth 7 8 discharged state. It should be noted that the specific stoichiometry of the Li and H coinserted λ 9 10 MnO2 spinel is not defined quantitatively, however, we anticipate the Li and H coinserted λ 11 12 MnO spinel is generated based on our ex situ SXRD and XPS. 13 2 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 Paragon Plus Environment 20 Page 21 of 33 The Journal of Physical Chemistry

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 Figure 4. (a) Ex situ SXRD pattern from the cathode at the hundredth discharged state, the 42 43 reference ZnMn2O4 (ICSD #15305) and the reference Li & H coinserted λMnO2 spinel (ICSD 44 45 #85606); (b) XPS spectra for Li 1s and Mn 3p region from the cathode at the hundredth 46 47 48 discharged state; (c) Ex situ SXRD pattern from the cathode at the hundredth charged state, the 49 50 reference ZnMn2O4 (ICSD #15305) and the reference λMnO2 spinel (ICSD #193445); (d) XPS 51 52 spectra for Li 1s and Mn 3p region from the cathode at the hundredth charged state. 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment 21 The Journal of Physical Chemistry Page 22 of 33

1 2 3 We estimate delithiated λMnO2 spinel is formed at the hundredth charged state as shown in 4 5 6 figure 4(c). It is noted that the right side of asymmetrically broadened peak at ~19.1º in the 7 8 SXRD pattern matches the (111) peak in the reference XRD pattern of λMnO2 (figure S10). 9 10 This asymmetric peak belongs to the convoluted peak from PTFE and λMnO2. We believe this 11 12 phase is formed by the extraction of Li and H process during charging as shown in figure 5. 13 14 15 There is no detection of Li in the Li 1s region from the XPS spectrum, which is evidence for the 16 17 formation of the delithiated spinel phase. In addition, the XPS spectrum for surface Mn species 18 19 can be used for qualitative understanding of elements’ change. The increase in the ratio of Mn4+ 20 21 3+ 22 species to Mn species indicates delithiation upon charging as shown in figure 4(b, d). 23 24 Assignments for Mn 3p of ~50 eV (Mn4+) and ~48.5 eV (Mn3+) were made based on the previous 25 26 reports.3536 The ratio is approximately 1:1 at the hundredth discharge which increases to 3:2 at 27 28 3+ 4+ + + 29 the hundredth charge. Mn is partially oxidized to Mn when Li and H leave the cathode. We 30 31 demonstrate the voltage curve at the hundredth cycle combining with the voltage curve at the 32 33 first cycle for comparison in figure S11. 34 35 36 We measure irreversible ZnMn2O4 from both the hundredth discharge/charge states in the ex situ 37 38 2 39 SXRD pattern (figure 4(a, c)). In Zn/MnO2 alkaline batteries, Zn(OH)4 is formed during the 40 26 41 discharge reaction: 42 43 2 44 Zn + 4OH → Zn(OH)4 + 2e (7) 45 46 2 37 47 Once Zn(OH)4 is formed, it chemically reacts with MnOOH at the cathode: 48 49 50 2 2MnOOH + Zn(OH)4 → ZnMn2O4 + 2H2O + 2OH (8) 51 52 53 ZnMn O has a tetrahedral crystal lattice where cornersharing ZnO tetrahedrons and edge 54 2 4 55 56 sharing MnO octahedrons are interconnected (figure 5). Mn and Zn have 3+ and 2+ oxidation 57 58 59 60 ACS Paragon Plus Environment 22 Page 23 of 33 The Journal of Physical Chemistry

1 2 3 states, respectively. ZnMn2O4 is known to have higher resistivity than MnO2 by six orders of 4 5 12 6 magnitude, which causes a loss of conductivity in the cathode. In this study, ZnMn2O4 is not 7 8 found after the first cycle as shown in figure 2(a), however, it is seen in the hundredth cycle in 9 10 both the discharged/charged states as shown in figure 4. The amount of ZnMn2O4 in the cathode 11 12 increases as the cycle of Zn/βMnO alkaline battery proceeds, which leads to capacity fading. 13 2 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 Figure 5. The reaction mechanism of βMnO2 during the hundredth cycle. 43 44 45 46 47 3+ 48 The effect of Bi2O3 additive. It has been reported that the addition of Bi either chemically or 49 50 physically to the MnO cathode improves the cycling performance of Znbased aqueous 51 2 52 3841 53 batteries. Dzieciuch et al. reported that Bi2O3 in the MnO2 cathode of alkaline batteries using 54 40 55 KOH electrolytes inhibits ZnMn2O4 formation to some level. In another study, Shin et al. found 56 57 58 59 60 ACS Paragon Plus Environment 23 The Journal of Physical Chemistry Page 24 of 33

1 2 3 41 that the zincate ion is deposited on Bi2O3 and prolongs the cycle life of Znbased batteries. In 4 5 3+ 6 order to study the effect of the Bi compound on nanosized βMnO2 in the 1 M KOH and 3 M 7 8 LiOH electrolyte, we physically mix Bi2O3 into the synthesized βMnO2 nanomaterial with 4% 9 10 mole fraction and measure the performance of Zn/βMnO2 alkaline battery. The discharge 11 12 capacity measured is 316 mAh g1 at the hundredth cycle (figure 6(a)). This indicates increased 13 14 15 capacity compared to the commercial βMnO2 and synthesized βMnO2 nanomaterial (112 and 16 1 17 225 mAh g , respectively). The CE is less than 100 % due to the H2 and O2 gas evolution within 18 19 the operating voltage window of a battery as was discussed earlier. Furthermore, the synthesized 20 21 22 βMnO2 nanomaterial with Bi2O3 incorporated demonstrates improved specific capacity over a 23 24 range of rates as shown in figure 6(b). This exhibits higher discharge capacity and better cycling 25 26 stability than the synthesized βMnO2 without Bi2O3. To be specific, the synthesized βMnO2 27 28 1 29 with Bi2O3 has an average capacity of 506 mAh g during the first stage and retains 86% of that 30 1 31 initial capacity (434 mAh g ) at the last stage. Comparatively, the synthesized βMnO2 without 32 1 1 33 Bi2O3 has an average capacity of 403 mAh g at the first stage and 292 mAh g at the last stage, 34 35 indicating 72% recovery. In order to investigate the role of Bi O in the rechargeable Zn/MnO 36 2 3 2 37 38 alkaline chemistry, Bi2O3 powder was immersed in the 6 M Zn nitrate solution. SEM images 39 40 show that original disordered structures transform to relatively ordered structures with island 41 42 meshes in a Zn2+ solution (figure S12). The EDS mapping for immersed Bi O powder that was 43 2 3 44 45 carefully cleansed and dried indicates the presence of Zn species on the surface of the Bi2O3 46 2+ 47 (figure S13). These SEM and EDS analyses imply that Zn reacts with the surface Bi2O3. In 48 49 addition, we took EDS of synthesized βMnO2 with and without Bi2O3 at the hundredth charged 50 51 52 state as shown in figure S14. We find that the atomic percent of Zn out of Mn and Zn reduces 53 54 from 11.71 % to 8.84 % in the synthesized βMnO2 with Bi2O3, indicating Bi2O3 additive 55 56 57 58 59 60 ACS Paragon Plus Environment 24 Page 25 of 33 The Journal of Physical Chemistry

1 2 3 alleviates the formation of ZnMn2O4. As was discussed for Zn/MnO2 alkaline batteries, 4 5 2 2+ 6 Zn(OH)4 or potentially a combination of Zn and OH in alkaline solution undergo chemical 7 8 reactions with Mn species to produce an electrochemically irreversible phase consisting of 9 10 ZnMn2O4. Our results suggest that the formation of ZnMn2O4 can be alleviated through 11 12 Zn(OH) 2 reactions with Bi O which lessens the likelihood that Mn species will react with 13 4 2 3 14 2 15 Zn(OH)4 . 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 Figure 6. (a) Cyclic performance for the synthesized βMnO with Bi O additive. The inset 37 2 2 3 38 39 indicates the voltage curve at the hundredth cycle; (b) Rate capability test for the synthesized β 40 41 MnO2 with and without the Bi2O3 additive are conducted with C/10, C/5, C/2, 1C, and C/10 in 42 43 sequence. 44 45 46 47 48 49 50 CONCLUSIONS 51 52 53 This study has demonstrated that synthesized βMnO2 nanomaterials have high utility for 54 55 rechargeable Zn/βMnO2 alkaline batteries. The nanosized βMnO2 cathode in the 1 M KOH and 56 57 58 59 60 ACS Paragon Plus Environment 25 The Journal of Physical Chemistry Page 26 of 33

1 2 3 3 M LiOH mixed electrolyte shows higher rechargeable capacity compared to the bulky βMnO2 4 5 6 cathode’s capacity. The βMnO2 nanomaterials have large surface area and strong surface 7 8 chemisorption property to facilitate electrochemical reactions in alkaline solutions. In addition, 9 10 the phase transformation of synthesized βMnO2 nanomaterial involves proton intercalation 11 12 reactions followed by twophase conversion reactions during the first cycle. At the hundredth 13 14 15 cycle, Li and H intercalate together into the host structure of λMnO2 spinel. Additionally the 16 17 physical incorporation of Bi2O3 into synthesized βMnO2 nanomaterials shows an excellent 18 19 discharge capacity of 316 mAh g1 at the hundredth cycle. The findings of this study may provide 20 21 22 useful insights towards the development of rechargeable Zn/MnO2 alkaline batteries in high 23 24 energy and lowcost stationary energy storage. 25 26 27 28 29 30 ASSOCIATED CONTENT 31 32 33 34 Supporting Information 35 36 37 Standard reduction potentials, standard Gibbs free energy of formation, Mn 3s splitting in the 38 39 XPS result of commercial βMnO2, Nitrogen adsorption/desorption isotherm of the βMnO2 and 40 41 BET surface areas, the electrochemical stability window of 1M KOH and 3M LiOH electrolytes 42 43 44 on Ni foam electrodes, inhouse XRD patterns for synthesized βMnO2 nanomaterial and its 45 46 intermediates during the first cycle, XRD patterns of MnO2 polymorphs from the ICSD database, 47 48 and inhouse XRD pattern for synthesized βMnO2 nanomaterial at the first charged state (1.80 49 50 51 V), the SEM image of MnO2 cathode at the first charged state (1.80 V), the XPS spectra for Zn 52 53 2p region from the cathode at the first and the hundredth charged state, the XRD pattern 54 55 measured at the first charged state (1.80 V) and possible Ni hydroxides phases, XRD patterns of 56 57 58 59 60 ACS Paragon Plus Environment 26 Page 27 of 33 The Journal of Physical Chemistry

1 2 3 lithiated Mn oxide spinels from ICSD database and the SXRD pattern at the hundredth 4 5 6 discharge, the SXRD pattern measured at the hundredth charged state, the comparison of voltage 7 8 curves between the first and the hundredth cycle, SEM images of Bi2O3 and immersed Bi2O3 in 9 10 6M Zn nitrate solution with corresponding EDS mapping for Zn Kα1 and Bi Mα1, and SEM 11 12 images and the corresponding EDS of the hundredth charged state for MnO cathode with and 13 2 14 15 without Bi2O3 (PDF) 16 17 18 19 20 AUTHOR INFORMATION 21 22 23 Corresponding Author 24 25 26 *Email: [email protected]. 27 28 29 30 31 32 Present Addresses 33 34 §(J.S.) Advanced Battery Center, KAIST Institute for NanoCentury, Korea Advanced Institute of 35 36 37 Science and Technology (KAIST), 291 Daehakro, Yuseonggu, Daejeon, 34141, Republic of 38 39 Korea. 40 41 42 Notes 43 44 45 The authors declare no competing financial interest. 46 47 48 49 50 ACKNOWLEDGMENTS 51 52 53 Joon Kyo Seo is grateful for William Getz and Fei Ji for fruitful discussions. This research was 54 55 funded by the Advanced Research Projects Agency Energy (ARPAE), U.S. Department of 56 57 58 59 60 ACS Paragon Plus Environment 27 The Journal of Physical Chemistry Page 28 of 33

1 2 3 Energy (award number DEAR0000535). The authors wish to acknowledge the computation 4 5 6 facilities at the Extreme Science and Engineering Discovery Environment (XSEDE), which is 7 8 supported by National Science Foundation (No. ACI1053575). The SEM analysis in this work 9 10 was performed at the San Diego Nanotechnology Infrastructure (SDNI), a member of the 11 12 National Nanotechnology Coordinated Infrastructure, which is supported by the National 13 14 15 Science Foundation (No. ECCS1542148). We performed XPS experiments at the University of 16 17 California, Irvine Materials Research Institute (IMRI) using instrumentation funded in part by 18 19 the National Science Foundation Major Research Instrumentation Program (grant number CHE 20 21 22 1338173). The authors extend their gratitude for the precious beam time and assistance from 23 24 scientists at beamlines 11BM in Argonne National Laboratory. The separator (FS 22145) was 25 26 provided from Freudenberg LLC. 27 28 29 30 31 32 33 34 REFERENCES 35 36 (1) Gallaway, J. W.; Erdonmez, C. K.; Zhong, Z.; Croft, M.; Sviridov, L. A.; Sholklapper, T. Z.; 37 38 Turney, D. E.; Banerjee, S.; Steingart, D. A. RealTime Materials Evolution Visualized within 39 40 Intact Cycling Alkaline Batteries. J. Mater. Chem. A 2014, 2, 27572764. 41 42 43 (2) Ingale, N. D.; Gallaway, J. W.; Nyce, M.; Couzis, A.; Banerjee, S. Rechargeability and 44 45 Economic Aspects of Alkaline Zinc– Dioxide Cells for Electrical Storage and Load 46 47 Leveling. J. Power Sources 2015, 276, 718. 48 49 50 (3) Turney, D. E., et al. Rechargeable Zinc Alkaline for LongCycle Energy Storage. 51 52 Chem. Mater. 2017, 29, 48194832. 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment 28 Page 29 of 33 The Journal of Physical Chemistry

1 2 3 (4) Julien, C. M.; Massot, M.; Poinsignon, C. Lattice Vibrations of Manganese Oxides: Part I. 4 5 6 Periodic Structures. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 7 8 2004, 60, 689700. 9 10 (5) Zhang, K.; Han, X.; Hu, Z.; Zhang, X.; Tao, Z.; Chen, J. Nanostructured MnBased Oxides 11 12 for Electrochemical Energy Storage and Conversion. Chem. Soc. Rev. 2015, 44, 699728. 13 14 15 (6) Patrice, R.; Gérand, B.; Leriche, J. B.; Seguin, L.; Wang, E.; Moses, R.; Brandt, K.; 16 17 Tarascon, J. M. Understanding the Second Electron Discharge Plateau in Mno2Based Alkaline 18 19 Cells. J. Electrochem. Soc. 2001, 148, A448A455. 20 21 22 (7) McBreen, J. The Electrochemistry of ΒMno2 and ΓMno2 in Alkaline Electrolyte. 23 24 Electrochim. Acta 1975, 20, 221225. 25 26 (8) Boden, D.; Venuto, C. J.; Wisler, D.; Wylie, R. B. The Alkaline 27 28 29 Electrode: I . The Discharge Process. J. Electrochem. Soc. 1967, 114, 415417. 30 31 (9) Lee, B.; Yoon, C. S.; Lee, H. R.; Chung, K. Y.; Cho, B. W.; Oh, S. H. Electrochemically 32 33 Induced Reversible Transition from the Tunneled to Layered Polymorphs of Manganese 34 35 Dioxide. Sci. Rep. 2014, 4, 6066. 36 37 38 (10) Lee, B.; Seo, H. R.; Lee, H. R.; Yoon, C. S.; Kim, J. H.; Chung, K. Y.; Cho, B. W.; Oh, S. 39 40 H. Critical Role of Ph Evolution of Electrolyte in the Reaction Mechanism for Rechargeable 41 42 Zinc Batteries. ChemSusChem 2016, 9, 29482956. 43 44 45 (11) Pan, H., et al. Reversible Aqueous Zinc/Manganese Oxide Energy Storage from 46 47 Conversion Reactions. Nat. Energy 2016, 1, 16039. 48 49 (12) Gallaway, J. W.; Menard, M.; Hertzberg, B.; Zhong, Z.; Croft, M.; Sviridov, L. A.; Turney, 50 51 52 D. E.; Banerjee, S.; Steingart, D. A.; Erdonmez, C. K. Hetaerolite Profiles in Alkaline Batteries 53 54 Measured by High Energy Edxrd. J. Electrochem. Soc. 2015, 162, A162A168. 55 56 57 58 59 60 ACS Paragon Plus Environment 29 The Journal of Physical Chemistry Page 30 of 33

1 2 3 (13) Hertzberg, B. J., et al. Effect of Multiple Cation Electrolyte Mixtures on Rechargeable Zn– 4 5 6 Mno2 Alkaline Battery. Chem. Mater. 2016, 28, 45364545. 7 8 (14) Chen, J.; Cheng, F. Combination of Lightweight Elements and Nanostructured Materials 9 10 for Batteries. Acc. Chem. Res. 2009, 42, 713723. 11 12 (15) Cheng, F.; Zhao, J.; Song, W.; Li, C.; Ma, H.; Chen, J.; Shen, P. Facile Controlled 13 14 15 Synthesis of Mno2 Nanostructures of Novel Shapes and Their Application in Batteries. Inorg. 16 17 Chem. 2006, 45, 20382044. 18 19 (16) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio TotalEnergy 20 21 22 Calculations Using a PlaneWave Basis Set. Phys. Rev. B 1996, 54, 1116911186. 23 24 (17) Dudarev, S. L.; Botton, G. A.; Savrasov, S. Y.; Humphreys, C. J.; Sutton, A. P. Electron 25 26 EnergyLoss Spectra and the Structural Stability of Nickel Oxide:An Lsda+U Study. Phys. Rev. 27 28 29 B 1998, 57, 15051509. 30 31 (18) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made 32 33 Simple. Phys. Rev. Lett. 1996, 77, 38653868. 34 35 (19) Blöchl, P. E. Projector AugmentedWave Method. Phys. Rev. B 1994, 50, 1795317979. 36 37 38 (20) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector AugmentedWave 39 40 Method. Phys. Rev. B 1999, 59, 17581775. 41 42 (21) Kresse, G.; Furthmüller, J. Efficiency of AbInitio Total Energy Calculations for Metals 43 44 45 and Semiconductors Using a PlaneWave Basis Set. Comput. Mater. Sci. 1996, 6, 1550. 46 47 (22) Persson, K. A.; Waldwick, B.; Lazic, P.; Ceder, G. Prediction of SolidAqueous Equilibria: 48 49 Scheme to Combine FirstPrinciples Calculations of Solids with Experimental Aqueous States. 50 51 52 Phys. Rev. B 2012, 85, 235438. 53 54 (23) Stranick, M. A. Mno2 by Xps. Surf. Sci. Spectra 1999, 6, 3138. 55 56 57 58 59 60 ACS Paragon Plus Environment 30 Page 31 of 33 The Journal of Physical Chemistry

1 2 3 (24) Han, B. C.; Miranda, C. R.; Ceder, G. Effect of Particle Size and Surface Structure on 4 5 6 Adsorption of O and Oh on Platinum Nanoparticles: A FirstPrinciples Study. Phys. Rev. B 2008, 7 8 77, 075410. 9 10 (25) Seo, J. K.; Khetan, A.; Seo, M. H.; Kim, H.; Han, B. FirstPrinciples Thermodynamic 11 12 Study of the Electrochemical Stability of Pt Nanoparticles in Applications. J. Power 13 14 15 Sources 2013, 238, 137143. 16 17 (26) Dirkse, T. P. The Behavior of the Zinc Electrode in Alkaline Solutions: Iii . The 18 19 Equilibrium Potential. J. Electrochem. Soc. 1979, 126, 14561459. 20 21 22 (27) Bratsch, S. G. Standard Electrode Potentials and Temperature Coefficients in Water at 23 24 298.15 K. J. Phys. Chem. Ref. Data 1989, 18, 121. 25 26 (28) Kordesch, K.; Gsellmann, J.; Tomantschger, K. The Alkaline Manganese DioxideZinc 27 28 29 Cell. J. Electroanal. Chem. Interfacial Electrochem. 1981, 118, 187201. 30 31 (29) Donne, S. W.; Lawrance, G. A.; Swinkels, D. A. J. Redox Processes at the Manganese 32 33 Dioxide Electrode: Ii. SlowScan Cyclic Voltammetry. J. Electrochem. Soc. 1997, 144, 2954 34 35 2961. 36 37 38 (30) , L.; Kordesch, K.; Urdl, P. Improvements of the Rechargeable Alkaline Mno2 ‧ Zn 39 40 Cell. J. Electrochem. Soc. 1996, 143, 1317. 41 42 (31) Kozawa, A.; KalnokiKis, T.; Yeager, J. F. Solubilities of Mn(Ii) and Mn(Iii) Ions in 43 44 45 Concentrated Alkaline Solutions. J. Electrochem. Soc. 1966, 113, 405409. 46 47 (32) Wang, J.G.; Jin, D.; Zhou, R.; Li, X.; Liu, X.r.; Shen, C.; Xie, K.; Li, B.; Kang, F.; Wei, 48 49 B. Highly Flexible Graphene/Mn3o4 Nanocomposite Membrane as Advanced Anodes for LiIon 50 51 52 Batteries. ACS Nano 2016, 10, 62276234. 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment 31 The Journal of Physical Chemistry Page 32 of 33

1 2 3 (33) Gao, J.; Lowe, M. A.; Abruña, H. D. Spongelike Nanosized Mn3o4 as a HighCapacity 4 5 6 Anode Material for Rechargeable Lithium Batteries. Chem. Mater. 2011, 23, 32233227. 7 8 (34) Wang, H.; Cui, L.F.; Yang, Y.; Sanchez Casalongue, H.; Robinson, J. T.; Liang, Y.; Cui, 9 10 Y.; Dai, H. Mn3o4−Graphene Hybrid as a HighCapacity Anode Material for Lithium Ion 11 12 Batteries. J. Am. Chem. Soc. 2010, 132, 1397813980. 13 14 15 (35) McCloy, J. S.; Leslie, C.; Kaspar, T.; Jiang, W.; Bordia, R. K. Magnetic Behavior of Ni and 16 17 Co Doped Cumn2o4 Spinels. Journal of Applied Physics 2012, 111, 07E149. 18 19 (36) Töpfer, J.; Feltz, A.; Gräf, D.; Hackl, B.; Raupach, L.; Weissbrodt, P. Cation Valencies and 20 21 22 Distribution in the Spinels Nimn2o4 and Mznimn2−Zo4 (M = Li, Cu) Studied by Xps. physica 23 24 status solidi (a) 1992, 134, 405415. 25 26 (37) Sharma, Y.; Haynes, A.; Binder, L.; Kordesch, K. The Effect of the Amount of Electrolyte 27 28 29 in the Anode Gel on the Rechargeability of Alkaline Manganese Dioxide  Zinc Cells. J. Power 30 31 32 Sources 1989, 27, 145153. 33 34 (38) Yao, Y. F.; Gupta, N.; Wroblowa, H. S. Rechargeable Manganese Oxide Electrodes. J. 35 36 37 Electroanal. Chem. Interfacial Electrochem. 1987, 223, 107117. 38 39 (39) Wroblowa, H. S.; Gupta, N. Rechargeable Manganese Oxide Electrodes. J. Electroanal. 40 41 Chem. Interfacial Electrochem. 1987, 238, 93102. 42 43 44 (40) Dzieciuch, M. A.; Gupta, N.; Wroblowa, H. S. Rechargeable Cells with Modified Mno2 45 46 Cathodes. J. Electrochem. Soc. 1988, 135, 24152418. 47 48 (41) Shin, J.; You, J.M.; Lee, J. Z.; Kumar, R.; Yin, L.; Wang, J.; Meng, Y. S. Deposition of 49 50 Zno on Bismuth Species Towards Rechargeable ZnBased Aqueous Battery. Phys. Chem. Chem. 51 52 53 Phys. 2016, 18, 26376. 54 55 56 57 58 59 60 ACS Paragon Plus Environment 32 Page 33 of 33 The Journal of Physical Chemistry

1 2 3 TABLE OF CONTENTS 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 Paragon Plus Environment 33