79
Rechargeable Aqueous Batteries Based on Available Resources Investigation and Development towards Efficient Battery Performance Mylad Chamoun Academic dissertation for the Degree of Doctor of Philosophy in Inorganic Chemistry at Stockholm University to be publicly defended on Friday 15 February 2019 at 13.00 in Magnélisalen, Kemiska övningslaboratoriet, Svante Arrhenius väg 16 B. Abstract Batteries employing water based electrolytes enable extremely low manufacturing costs and are inherently safer than Li-ion batteries. Batteries based on zinc, manganese dioxide, iron, and air have high energy relevancy, are not resource restricted, and can contribute to large scale energy storage solutions. Zinc has a rich history as electrode material for primary alkaline Zn–MnO 2 batteries. Historically, its use in secondary batteries has been limited because of morphological uncertainties and passivation effects that may lead to cell failure. Manganese dioxide electrodes are ineffective as rechargeable electrodes because of failure mechanisms associated with phase transformations during cycling. The irreversibility of manganese dioxide is strongly correlated to the formation of the electrochemically inactive spinel, Mn 3 O 4 /ZnMn 2 O 4 . The development of the iron electrode for Fe–air batteries was initiated in late the 1960s and these batteries still suffer from charging inefficiency, due to the unwanted hydrogen evolution reaction. Meanwhile, the air electrode is limited in long-term operation because of the sluggish oxygen evolution and reduction kinetics. These limitations of the Fe–air battery yield poor overall efficiencies, which bring vast energy losses upon cycling. Herein, the limitations described above were countered for rechargeable Zn–MnO 2 and Fe–air batteries by synthesizing electrode materials and modifying electrolyte compositions. The electrolyte mixture of 1 M KOH + 3 M LiOH for rechargeable alkaline Zn–MnO 2 batteries limited the formation of the inactive spinels and improved their cycle life significantly. Further, the formation of the inactive spinels was overcome in mildly acidic electrolytes containing 2 M ZnSO 4 , enabling the cells to cycle reversibly at lower pH via a distinctive reaction mechanism. The iron electrodes were improved with the addition of stannate, which suppressed hydrogen evolution. Furthermore, optimal charge protocols of the iron electrodes were identified to minimize the hydrogen evolution rate. On the air electrode, the synthesized NiCo 2 O 4 showed excellent bifunctional catalytic activity for oxygen evolution and reduction, and was incorporated to a flow assisted rechargeable Fe–air battery, in order to prove the practicability of this technology. Studies of the electrode materials on the micro, macro, nano, and atomic scales were carried out to increase the understanding of the nature of and interactions between of these materials. This included both in operando and ex situ characterization. X-ray and neutron radiation, and analytical- and electrochemical methods provided insight to improve the performance and cycle life of the batteries. Keywords: rechargeable aqueous batteries, alkaline electrolytes, aqueous sulfate electrolytes, zinc electrodes, manganese dioxide electrodes, iron electrodes, air electrodes, oxygen electrocatalysts. Stockholm 2019 http://urn.kb.se/resolve?urn=urn:nbn:se:su:diva-163154 ISBN 978-91-7797-552-6 ISBN 978-91-7797-553-3 Department of Materials and Environmental Chemistry (MMK) Stockholm University, 106 91 Stockholm

DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

Rechargeable Aqueous Batteries Based onAvailable ResourcesInvestigation and Development towards Efficient BatteryPerformanceMylad Chamoun

Academic dissertation for the Degree of Doctor of Philosophy in Inorganic Chemistry atStockholm University to be publicly defended on Friday 15 February 2019 at 13.00 inMagnélisalen, Kemiska övningslaboratoriet, Svante Arrhenius väg 16 B.

AbstractBatteries employing water based electrolytes enable extremely low manufacturing costs and are inherently safer than Li-ionbatteries. Batteries based on zinc, manganese dioxide, iron, and air have high energy relevancy, are not resource restricted,and can contribute to large scale energy storage solutions. Zinc has a rich history as electrode material for primary alkalineZn–MnO2 batteries. Historically, its use in secondary batteries has been limited because of morphological uncertainties andpassivation effects that may lead to cell failure. Manganese dioxide electrodes are ineffective as rechargeable electrodesbecause of failure mechanisms associated with phase transformations during cycling. The irreversibility of manganesedioxide is strongly correlated to the formation of the electrochemically inactive spinel, Mn3O4/ZnMn2O4. The developmentof the iron electrode for Fe–air batteries was initiated in late the 1960s and these batteries still suffer from charginginefficiency, due to the unwanted hydrogen evolution reaction. Meanwhile, the air electrode is limited in long-termoperation because of the sluggish oxygen evolution and reduction kinetics. These limitations of the Fe–air battery yieldpoor overall efficiencies, which bring vast energy losses upon cycling.

Herein, the limitations described above were countered for rechargeable Zn–MnO2 and Fe–air batteries by synthesizingelectrode materials and modifying electrolyte compositions. The electrolyte mixture of 1 M KOH + 3 M LiOH forrechargeable alkaline Zn–MnO2 batteries limited the formation of the inactive spinels and improved their cycle lifesignificantly. Further, the formation of the inactive spinels was overcome in mildly acidic electrolytes containing 2 MZnSO4, enabling the cells to cycle reversibly at lower pH via a distinctive reaction mechanism. The iron electrodes wereimproved with the addition of stannate, which suppressed hydrogen evolution. Furthermore, optimal charge protocols ofthe iron electrodes were identified to minimize the hydrogen evolution rate. On the air electrode, the synthesized NiCo2O4showed excellent bifunctional catalytic activity for oxygen evolution and reduction, and was incorporated to a flow assistedrechargeable Fe–air battery, in order to prove the practicability of this technology. Studies of the electrode materials onthe micro, macro, nano, and atomic scales were carried out to increase the understanding of the nature of and interactionsbetween of these materials. This included both in operando and ex situ characterization. X-ray and neutron radiation, andanalytical- and electrochemical methods provided insight to improve the performance and cycle life of the batteries.

Keywords: rechargeable aqueous batteries, alkaline electrolytes, aqueous sulfate electrolytes, zinc electrodes,manganese dioxide electrodes, iron electrodes, air electrodes, oxygen electrocatalysts.

Stockholm 2019http://urn.kb.se/resolve?urn=urn:nbn:se:su:diva-163154

ISBN 978-91-7797-552-6ISBN 978-91-7797-553-3

Department of Materials and EnvironmentalChemistry (MMK)

Stockholm University, 106 91 Stockholm

Page 2: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron
Page 3: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

RECHARGEABLE AQUEOUS BATTERIES BASED ON AVAILABLERESOURCES

Mylad Chamoun

Page 4: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron
Page 5: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

Rechargeable AqueousBatteries Based on AvailableResources

Investigation and Development towards Efficient Battery Performance

Mylad Chamoun

Page 6: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

©Mylad Chamoun, Stockholm University 2019 ISBN print 978-91-7797-552-6ISBN PDF 978-91-7797-553-3 Cover: Vector images adapted from Vecteezy.com & Freepik.com Printed in Sweden by Universitetsservice US-AB, Stockholm 2019

Page 7: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

Doctoral Thesis 2019Department of Materials and Environmental ChemistryArrhenius Laboratory, Stockholm UniversitySE-10691 Stockholm, Sweden Faculty opponent: Prof. Ann Mari SvenssonDepartment of Materials Science and EngineeringNorwegian University of Science and Technology (NTNU) Evaluation committee: Prof. Göran LindberghDepartment of Chemical Engineering and TechnologyThe Royal Institute of Technology (KTH), Sweden Dr. Helena BergCEO & Owner, AB Libergreen Prof. Daniel BrandellDepartment of Chemistry - Ångström LaboratoryUppsala University Substitute: Prof. Jiayin YuanDepartment of Materials and Environmental ChemistryStockholm University Cover: Investigated rechargeable aqueous battery chemistriesfor electrical power systems with renewable energy sourcesinstalled such as solar and wind power.

Page 8: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron
Page 9: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

i

List of publications

This thesis is based on the following publications:

Paper I:

Effect of Multiple Cation Electrolyte Mixtures on Rechargeable Zn-MnO2 Alkaline Battery

B. Hertzberg, A. Huang, A. Hsieh, M. Chamoun, G. Davies, K. J. Seo, Z. Zhong, M. Croft, C. Erdonmez,

S. Meng, D. Steingart.

Chemistry of Materials, 2016, 28 (13), 4536-4545

My contribution: Synthesized the MBDB material, contributed to the collection, and processing of

the operando EDXRD data, conducted parts of the electrochemical characterization, and wrote parts

of the manuscript.

Paper II:

Stannate Increases Hydrogen Evolution Overpotential on Rechargeable Alkaline Iron Electrodes

M. Chamoun, B. Skårman, H. Vidarsson, R. I. Smith, S. Hull, M. Lelis, D. Milcius, D. Noréus.

Journal of The Electrochemical Society, 2017, 164 (6), 1251-1257

My contribution: Conducted the electrochemical and structural characterization (SEM, XRD and

EDS), contributed to the collection and processing of the operando neutron diffraction data, and

wrote most of the manuscript except the XPS part.

Paper III:

Rechargeability of Aqueous Sulfate Zn/MnO2 Batteries Enhanced by Accessible Mn2+ ions

M. Chamoun, W. R. Brant, CW. Tai, G. Karlsson, D. Noréus.

Energy Storage Materials, 2018, 15, 351-360

My contribution: Conducted the electrochemical and structural characterization (SEM, XRD, EDS and

ICP-AES), contributed to data collection and the processing of the operando XRD data, performed

the operando pH measurements and quantification of hydrogen on zinc electrodes, and wrote most

of the manuscript except the TEM/EELS parts.

Paper IV:

Electrochemical Performance and in Operando Charge Efficiency Measurements of Cu/Sn-Doped

Nano Iron Electrodes

A. R. Paulraj, Y. Kiros, M. Chamoun, H. Svengren, D. Noréus, M. Göthelid, B. Skårman, H. Vidarsson,

M. B. Johansson.

Batteries, 2019, 5, 1-15

My contribution: Contributed to data collection and processing of quantifying hydrogen on iron

electrodes and wrote parts of the manuscript.

Page 10: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

ii

Paper V:

Bifunctional Performance of Flow Assisted Rechargeable Iron-Air Alkaline Batteries

M. Chamoun, A. R. Paulraj, B. Skårman, H. Vidarsson, Y. Kiros, D. Noréus.

In manuscript

My contribution: Synthesized the oxygen electrocatalysts (except LCMO), conducted the

electrochemical and structural characterization (SEM and XRD), developed the Fe–air cell setup, and

wrote most of the manuscript.

Publications not included in this thesis:

Paper VI:

Water Splitting Catalysis Studied by using Real-Time Faradaic Efficiency Obtained through Coupled

Electrolysis and Mass Spectrometry

Svengren H., Chamoun M., Grins J., Johnsson M.

ChemElectroChem, 2017, 5 (1), 44-50

Reprints were made with permission from the publishers.

Page 11: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

iii

Contents

List of publications ................................................................................................................................... i

Abbreviations .......................................................................................................................................... v

1. Introduction .................................................................................................................................... 1

1.1. Large-scale energy storage systems ....................................................................................... 1

1.2. Electrochemical energy storage .............................................................................................. 2

1.3. Rechargeable aqueous batteries based on available resources ............................................. 2

1.4. Investigated electrode materials for rechargeable aqueous batteries .................................. 3

1.4.1. Zinc .................................................................................................................................. 3

1.4.2. Electrochemical challenges of zinc ................................................................................. 4

1.4.3. Manganese dioxide ......................................................................................................... 5

1.4.4. Electrochemical challenges of manganese dioxide ........................................................ 7

1.4.5. Iron .................................................................................................................................. 8

1.4.6. Electrochemical challenges of iron ................................................................................. 9

1.4.7. Oxygen electrocatalysts .................................................................................................. 9

1.4.8. Electrochemical challenges of oxygen electrocatalysts ................................................ 10

1.5. The aim of the thesis ............................................................................................................. 11

2. Experimental ................................................................................................................................. 13

2.1. Synthesis ............................................................................................................................... 13

2.1.1. Bismuth-doped β-MnO2 ................................................................................................ 13

2.1.2. Oxygen electrocatalysts ................................................................................................ 13

2.2. Cell preparation and electrochemical characterization ........................................................ 13

2.2.1. Alkaline Zn–MnO2 ......................................................................................................... 13

2.2.2. Aqueous sulfate Zn–MnO2 ............................................................................................ 14

2.2.3. Iron electrode ................................................................................................................ 14

2.2.4. Air electrode.................................................................................................................. 15

2.2.5. Fe–air prototype ........................................................................................................... 15

2.2.6. Coulombic efficiency ..................................................................................................... 16

2.3. Operando techniques ........................................................................................................... 17

2.3.1. Energy-dispersive X-ray diffraction (EDXRD) ................................................................ 17

2.3.2. Neutron diffraction ....................................................................................................... 17

2.3.3. X-ray diffraction (XRD) .................................................................................................. 17

Page 12: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

iv

2.4. Scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDS) ...... 18

2.5. XRD ........................................................................................................................................ 18

2.6. Electron energy loss spectroscopy (EELS) ............................................................................. 19

2.7. X-ray photoelectron spectroscopy (XPS) .............................................................................. 19

3. Results and discussion .................................................................................................................. 20

3.1. Cation electrolyte mixtures for rechargeable Zn–MnO2 alkaline batteries (Paper I) ........... 20

3.1.1. Structural characterization of bismuth-doped β-MnO2 (MBDB) .................................. 20

3.1.2. Electrochemical performance of KOH:LiOH electrolytes .............................................. 21

3.1.3. Phase evolution investigation of MBDB electrodes ...................................................... 21

3.2. Reversible aqueous sulfate Zn–MnO2 batteries with Mn2+ (Paper III) .................................. 23

3.2.1. Electrochemical characterization of MnO2 electrodes ................................................. 23

3.2.2. Characterization of cycled MnO2 electrodes ................................................................ 24

3.2.3. Progression of the MnO2 charge and discharge mechanism ........................................ 27

3.3. Effect of stannate on rechargeable iron electrodes (Paper II) .............................................. 30

3.3.1. Electrochemical and structural characterization of iron electrodes ............................ 30

3.3.2. Phase evolution characterization and structure refinement ........................................ 33

3.4. Hydrogen evolution on iron and zinc electrodes (Papers III and IV) .................................... 35

3.5. Flow-assisted rechargeable Fe–air batteries (Paper V) ........................................................ 38

3.5.1. Structural characterization of oxygen electrocatalysts ................................................ 38

3.5.2. Electrochemical performance of air electrodes ............................................................ 40

3.5.3. Electrochemical performance of the Fe–air prototype ................................................ 42

4. Conclusions ................................................................................................................................... 46

5. Future perspectives ...................................................................................................................... 48

6. Sammanfattning ............................................................................................................................ 49

Acknowledgements ............................................................................................................................... 51

References ............................................................................................................................................ 53

Page 13: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

v

Abbreviations

CAES Compressed-air energy storage

CV Cyclic voltammetry

DOD Depth-of-discharge

EDS Energy-dispersive X-ray spectroscopy

EDXRD Energy-dispersive X-ray diffraction

EELS Electron energy loss spectroscopy

EMD Electrolytic manganese dioxide

EPDM Ethylene propylene diene monomer

ESS Energy-storage systems

GDL Gas diffusion layer

HER Hydrogen-evolution reaction

IHP Intermediate hydride phase

LCMO La1–xCaxMnO3

LDH Layered double hydroxide

NMP N-Methyl-2-pyrrolidone

MBDB Bismuth doped β-MnO2

MS Mass spectrometer

PDF Powder diffraction file

PE Polyethylene

PHS Pumped hydroelectric storage

PTFE Polytetrafluoroethylene

PVC Polyvinyl chloride

PVP Polyvinylpyrrolidone

OER Oxygen-evolution reaction

ORR Oxygen-reduction reaction

SEM Scanning electron microscopy

SHE Standard hydrogen electrode

SMES Superconducting magnetic energy storage

SOC State-of-charge

SS Stainless steel

XPS X-ray photoelectron spectroscopy

XRD X-ray diffraction

Page 14: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

1

1. Introduction 1.1. Large-scale energy storage systems

We are now shifting our electrical power systems from using fossil fuels to renewable energy

sources. The incentive is to replace large traditional power plants with smaller non-dispatchable

renewable energy sources. The paradigm shift to renewable energy needs, however, to cope with

the increasing global energy demand, estimated to double by 2050, and without increasing carbon

dioxide emission levels or reliance on restricted fossil fuel resources1.

Shifting the entire electrical power system to renewable energy sources such as wind, solar, and

hydro is a complex task. Their inherently intermittent character is dependent on time and weather

and results in unpredictable generation profiles, which may not match with the energy-consumption

profile. Imbalance between power generation and demand already exists in current power systems.

With increasing renewable energy sources installed, the uncertainty in predicting power generation

adequacy will increase2. Thus, technologies that improve the resiliency of the power system are

necessary. To make the best use of a power system with renewable energy installed, we need

efficient energy-storage systems (ESS) to handle load fluctuations and ensure reliable power delivery

whenever needed. In spite of the advantages with energy storage, only 1% of the global energy

presently used had been stored, mostly through pumped hydroelectric storage, which accounts for

98% of the total installed storage systems1.

ESS that are available on a large scale can be divided into four groups: 1) mechanical, 2) electrical, 3)

electrochemical and 4) chemical. These consist of technologies such as 1) pumped hydroelectric

storage (PHS), compressed-air energy storage (CAES) and flywheels, 2) superconducting magnetic

energy storage (SMES) and electrical supercapacitors, 3) batteries and electrochemical

supercapacitors, and 4) power-to-hydrogen or synthetic natural-gas production3,4. In Table 1.1, the

characteristics of these ESS are compared with regard to power output and discharge time.

Depending on the time scale of service, these technologies can support electrical power systems by

facilitating frequency regulation and load balancing, enhancing power quality, and providing an

uninterruptible power supply. These assets will improve power systems quality, stability and

reliability2,5.

Table 1.1. Characteristics of different energy storage technologies6.

Technology Power output (MW) Discharge time Efficiency (%) Start time

PHS < 5000 1 – 24 h 65 – 85 s – min CAES Depending on storage size 1 – 24 h 42 – 70 min Flywheels 0.002 – 20 s – min 95 s – min SMES 0.001 – 10 s 90 ms Supercapacitors 0.01 – 1 ms – s 95 ms Lead-acid batteries 0.001 – 50 s – 3 h 60 – 95 < ms Lithium-ion batteries 0.001 – 2 min – h 85 – 99 < ms Vanadium redox flow batteries 0.03 – 20 s – 10 h 85 ms Sodium sulfur batteries 0.5 – 50 s – h 85 – 90 < ms Power to H2 gas kW – GW s – months 62 – 82 s – min Power to CH4 gas kW – GW s – months 49 – 56 min – h

Page 15: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

2

1.2. Electrochemical energy storage

Electrochemical energy-storage technologies include batteries, redox flow batteries, electrochemical

supercapacitors, and fuel cells. The technologies are distinguished by their energy storage

mechanisms. Batteries store energy through electron transfer reactions, wherein the oxidation

states of reactants change. Redox flow batteries follow the same mechanism except that the redox

species circulate. By comparison, supercapacitors undergo capacitive charging from the electric

double layer at the interface between the electrode and the electrolyte. The technologies all

function as closed systems except for fuel cells, which store energy from external reactants fed to

the cells7.

Batteries are of high interest and contain electrochemical cells with two electrodes where the redox

reactions ensue. An electrolyte separates the electrodes and contains dissolved ions that can be

transferred freely from one electrode to the other. The electrodes are connected externally and, as

the redox reactions proceed, electrons transferred via the outer circuit create a current. These cells

can then be connected in series and/or in parallel to provide the desired voltage and/or capacity,

respectively8. Important characteristics of batteries are their energy density (Wh L–1), specific energy

(Wh kg–1) and specific power (W kg–1), i.e. how much energy they can store per unit mass or volume,

and how quickly they can deliver this energy. There is a trade-off in batteries between short-term

(power density) and long-term storage (energy density), as described by Ragone9, and these can be

customized to the application. Table 1.2 lists energy and power characteristics of mature battery

technologies for large-scale applications. In large-scale energy storage the primary factors are not

energy and power density, but rather low installation cost, long cycle life, high energy efficiency and

the ease of scaling up the storage capacity10.

Table 1.2. Energy and power characteristics of mature battery technologies for large-scale applications5.

Battery Energy density (Wh L–1)

Specific energy (Wh kg–1)

Specific power (W kg–1)

Cycle life

Lead-acid 60 – 75 30 – 40 60 – 110 100 – 500 Nickel-Cadmium 130 – 150 40 – 60 40 – 100 2000 Nickel-Metal hydride 250 – 330 70 – 100 70 – 200 1000 Lithium-ion - Li(Ni,Co,Mn)O2 – C 200 – 250 120 – 160 200 – 300 300 –1000 Lithium-ion - LiFePO4 – C 120 – 150 80 – 90 200 – 300 1500 – 2000 Sodium-Sulfur 70 – 150 60 – 120 15 – 70 4000 Vanadium Redox Flow 10 – 20 10 – 20 1 – 4 5000

1.3. Rechargeable aqueous batteries based on available

resources

The most important aspect of manufacturing batteries for large-scale energy storage is the price set

by the market. The availability of the materials and the processes used to manufacture the devices

drive the cost. These two factors must coincide with a sustainable life cycle for large-volume

markets. Thus, scrutinizing feasible materials to develop sustainable and efficient batteries is critical

and not an easy task. Forecasts of the availability of materials are inaccurate and vary depending on

the state of the art in industrial sectors1. Among the various types of batteries available today, non-

aqueous lithium-ion batteries are the most prominent choice because of their high energy density

and versatile design capabilities that allow them to meet energy and power demands11. However,

Page 16: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

3

cost12, safety13 and lifetime will limit their full-scale implementation in electrical power systems, for

which low-cost and long service life are the main concerns. For instance, the use of cobalt in the

layered LiCoO2 electrode material has been the benchmark in lithium-ion batteries despite that the

availability of cobalt is low14. The European Commission identified cobalt as a critical raw material in

2017 because of its significant supply risk15. The supply risk originates from geopolitical issues with

the Democratic Republic of Congo, which is the dominant global producer of cobalt (64%). The

unsustainable cobalt supply has encouraged research into other battery chemistries16–20.

Batteries using aqueous electrolytes are inherently safer and less expensive than their non-aqueous

counterparts. Aqueous electrolytes have significantly higher ionic conductivities (up to 1 S cm–1),

than non-aqueous ones (typically around 1 – 10 mS cm–1 21). This favors aqueous electrolytes for

high-rate operations e.g. when sudden energy deliveries are needed, in particular in quick-response

balancing systems in the electrical grid2. At present, lead-acid batteries dominate the aqueous

battery market because of their high rate capability and low system-installation price. Lead-acid

batteries find its use as start battery or backup battery with low demand for cycle life10.

Proposed potential electrode materials for large-scale energy storage are zinc, manganese dioxide,

iron, and non-precious-metal-based catalysts for the air electrode. These electrode materials are

adopted in rechargeable aqueous Zn–MnO222,23 and Fe–air24 batteries, both having high energy

relevancy and unrestricted availability25. Their challenges concerning irreversibility and inefficiency

are in this thesis investigated and alleviated in order to make viable systems. The studied electrode

materials are not exclusively applicable for rechargeable aqueous batteries and the work is intended

to shed light on the importance of using available materials in the development of future electrical

power system. The batteries designed in this thesis are relevant for large scale energy storage with

assured safe operation and low total cost.

1.4. Investigated electrode materials for rechargeable

aqueous batteries

1.4.1. Zinc

Zinc has a rich history in alkaline26–28, mildly acidic29 and redox-flow rechargeable batteries30–32. It

possesses attractive attributes for an electrode material, such as abundancy, low toxicity and a high

specific theoretical capacity of 820 mAh g–1. Furthermore, it is the most electropositive metal that

does not have noteworthy corrosion issues in aqueous electrolytes between pH 4 and 1433, making it

an outstanding electrode material34. The alkaline battery has been the working horse in the primary

battery market for over 60 years. This battery contains zinc and manganese dioxide and delivers a

specific energy density of 150 Wh kg–1, comparable to some lithium-ion chemistries35. In recent

years, zinc has been coupled with several electrolyte and electrode combinations for high-

performance rechargeable batteries. Electrodes used in alkaline electrolytes include Ni–Zn36,37, Zn–

air38–40 and Zn–MnO223,41,42. For mildly acidic electrolytes, electrode materials with open crystal

structures that are capable of hosting zinc ions20,43–46 are used, and in redox-flow cells, Zn–Br247,48,

Zn–I249 and Zn–Fe50,51 have been used as electrodes. In static cells, zinc is found as composites,

pastes, or powders, whereas flow cells use dissolved zinc ions sourced from various salts.

In alkaline electrolytes operating above pH 14, zinc exists in equilibrium with zincate ions, Zn(OH)42–,

and zinc oxide, ZnO, precipitates when zincate exceeds its supersaturated concentration limit34:

Page 17: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

4

𝑍𝑛(𝑂𝐻)42− + 2𝑒− 𝑍𝑛 + 4𝑂𝐻− E° = –1.12 V vs. Standard hydrogen electrode (SHE) (1.1)

𝑍𝑛(𝑂𝐻)42− 𝑍𝑛𝑂 + 𝐻2𝑂 + 2𝑂𝐻− (1.2)

where E° is the standard potential relative to the SHE at 25 °C.

In mildly acidic electrolytes, at pH 4–6, zinc dissolves to Zn2+ during discharge and is electrodeposited

as zinc metal during charge22:

𝑍𝑛2+ + 2𝑒− 𝑍𝑛 E° = -0.76 V vs. SHE (1.3)

In Figure 1.1, the Pourbiax diagram shows that the redox potential of zinc is below that at which the

hydrogen evolution reaction (HER) occurs. Pourbaix diagrams depict possible stable phases of an

electrochemical system at equilibrium and do not consider kinetic effects. Based on

thermodynamics, the HER should dominate at zinc redox potentials, but luckily that reaction is

sluggish, which enables zinc to be used as electrode material.

Figure 1.1. Pourbaix diagram of 10–5 M Zn2+(aq) at 25 °C, created by the software Medusa®. Highlighted arrows

in red and blue depict the mildly acidic and alkaline pH regions, respectively, and the dashed green lines

correspond to the oxygen evolution and hydrogen evolution reactions.

1.4.2. Electrochemical challenges of zinc

Zinc faces several challenges when adopted in aqueous rechargeable alkaline batteries. Shape

changes and morphological uncertainties affect the deposition of the metal during charge. Zinc

tends to plate anisotropically and this induces localized mass-transport-limited regions52. The

anisotropic growth ramifies, and dendrite formation increases as the mass-transport limitation

increase. These dendrites eventually cause short-circuiting if they penetrate the separator53. Figure

1.2 illustrates the uneven deposition of zinc and the formation of dendrites that may short-circuit

the Zn–MnO2 cell.

mildly acidic

alkaline

Page 18: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

5

Figure 1.2. Illustration of uneven zinc deposition and the formation of dendrites that short-circuit the cell.

Engineering zinc structures can extend the cycle life but shape changes upon cycling are not easily

avoided. Reported fine three-dimensional structures of zinc have provided longer cycle life while

sacrificing energy density54–56. Nevertheless, these three-dimensional structures have limited

utilization due to rapid dissolution of the metal57. Low zinc utilization during dissolution is mainly

associated with corrosion and passivation effects33. Hence, utilization is limited to 60% or less58.

Limiting passivation and corrosion effects is a major challenge related to the particle size of zinc and

the solubility of zincate. The smaller the zinc particles are (or the higher their surface area is), the

more aggravated is the corrosion. The corrosion reaction is caused by the HER and is parasitic

because it consumes water without contributing any useful charge capacity to the battery:

2𝐻2𝑂 + 2𝑒− → 𝐻2 + 2𝑂𝐻− E°= –0.83 V vs. SHE (1.4)

Using zinc particles with low surface area limits corrosion at the expense of passivation35. Passivation

is dependent on the solubility of zincate and is caused by precipitation of ZnO when the solution

becomes saturated. Initially, porous ZnO forms, but this densifies overtime and eventually passivates

the zinc59.

Mildly acidic electrolytes are more forgiving than alkaline electrolytes. Passivation from ZnO is

prevented because the pH is lower. Instead, zinc dissolves to Zn2+ during discharge and electroplates

back during charge (1.3). The reversibility has been extensively studied and is good60–64. Concerns of

the HER lowering the Coulombic efficiency and potential dendrite formation remain. Other work has

focused on substituting sulfate with other anions, adding surfactants or adjusting the concentration

of the salts used43. The pH of the electrolyte must be maintained at 4–6 during battery operation to

avoid severe corrosion and passivation33.

1.4.3. Manganese dioxide

Manganese dioxide is used in primary alkaline batteries for a wide range of power electronics65, as

well as in secondary batteries such as lithium-66, sodium-67, magnesium-68, and zinc-ion batteries43.

Electrolyte

Zn

Mn

O2

e

-

Short-circuit

Page 19: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

6

Inexpensive production, high theoretical energy capacity, high redox potential, and low

environmental impact make manganese dioxide an excellent electrode material69.

Manganese dioxide comes in a variety of polymorphs depending on the synthesis conditions:

tetragonal pyrolusite β-MnO2, orthorhombic ramsdellite, tetragonal hollandite ɑ-MnO2, hexagonal

birnessite δ-MnO2, monoclinic romanèchite and cubic λ-MnO2. These structures can be described by

different distributions of Mn4+ cations over octahedral sites in the oxygen atom arrangement. MnO6

octahedra sharing opposite octahedral edges form MnO6 chains parallel to the c-axis. The chains

further connect to each other in different ways and tunnels form along the c-axis. The tunnels can be

classified by the number of MnO6 units and chains between two basal planes70. The stable form of

manganese dioxide β-MnO2 is built by MnO6 units forming a 1 1 tunnel structure. Among the

polymorphs, β-MnO2 is the least electrochemically active. Electrolytic manganese dioxide (EMD),

also known as γ-MnO2, is composed of intergrown β-MnO2 and ramsdellite. In modern alkaline,

lithium or other types of batteries, EMD is used because of its high electrochemical activity, high

manganese content, and purity. EMD is produced via electrochemical deposition from acidic sulfate

baths containing Mn2+ ions and undergoes a two-electron oxidation65:

𝑀𝑛2+ + 2𝐻2𝑂 → 𝑀𝑛𝑂2 + 4𝐻+ + 2𝑒− (1.5)

The redox reactions of manganese dioxide vary with pH. Two pH regions are of interest: one above

pH 14 (alkaline electrolytes) and the other at pH 4–6 (mildly acidic electrolytes). Figure 1.3 shows the

Pourbaix diagram of these two pH regions and corresponding redox reactions.

Figure 1.3. Pourbaix diagram of 10–5 M Mn2+(aq) at 25 °C, created by the software Medusa®. Highlighted

arrows in red and blue depict mildly acidic and alkaline pH regions, respectively, and the dashed green lines

correspond to the oxygen evolution and hydrogen evolution reactions.

The discharge reaction of EMD in alkaline electrolytes includes a homogeneous one-electron

reduction via proton insertion to form MnOOH:

𝑀𝑛𝑂2 + 𝐻2𝑂 + 𝑒− → 𝑀𝑛𝑂𝑂𝐻 + 𝑂𝐻− E° = +0.36 V vs. SHE (1.6)

mildly acidic

alkaline

Page 20: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

7

Further one-electron reduction proceeds through heterogeneous dissolution to a give soluble

hydroxymanganese complex, which then precipitates as Mn(OH)271:

𝑀𝑛𝑂𝑂𝐻 + 𝐻2𝑂 + 𝑒− → 𝑀𝑛(𝑂𝐻)2 + 𝑂𝐻− E° = –0.28 V vs. SHE (1.7)

In mildly acidic electrolytes, EMD initially dissolves to trivalent manganese (𝑀𝑛4+ + 𝑒− → 𝑀𝑛3+).

This Jahn–Teller Mn3+ cation is unstable because of its high-spin electronic configuration and

disproportionates to Mn4+ and Mn2+ ions (2𝑀𝑛3+ → 𝑀𝑛4+ + 𝑀𝑛2+)61. The total discharge reaction

mechanism may be simplified to72:

𝑀𝑛𝑂2 + 4𝐻+ + 2𝑒− → 𝑀𝑛2+ + 2𝐻2𝑂 E° = +1.23 V vs. SHE (1.8)

The two-electron transfer in both electrolytes corresponds to a theoretical capacity of 617 mAh g–1.

1.4.4. Electrochemical challenges of manganese dioxide

Manganese dioxide in alkaline electrolytes can evolve several failure mechanisms upon battery

cycling. These mechanisms are strongly correlated to the irreversible formation of inactive phases

over time. Figure 1.4 shows the reactions during charge-discharge cycling of EMD. The figure

highlights the significant phase transformations leading to irreversibility.

Figure 1.4. Reaction mechanism of EMD (γ-MnO2) upon charge-discharge cycling in alkaline electrolyte. Figure

reproduced from Paper I69 with permission from American Chemical Society.

The first discharge involves a phase transformation of γ-MnO2 into ɑ-MnOOH (2 1 tunnel structure)

and γ-MnOOH (1 1 tunnel structure) via proton insertion, with a change of the manganese valence

state from 4+ to 3+73. The reduction proceeds through the partial formation of the spinel phase

Mn3O4, or ZnMn2O4 if zinc is present. The reduction continues from Mn3O4 to the final discharge

product Mn(OH)2, with the manganese valence state of 2+73,74. The partial formation of

Page 21: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

8

Mn3O4/ZnMn2O4 does not involve the exchange of electrons but is based on whichever complex ion

is available to fill the tetrahedral lattice73:

2𝑀𝑛𝑂𝑂𝐻 + 𝑀𝑛(𝑂𝐻)42−/𝑍𝑛(𝑂𝐻)4

2− → 𝑀𝑛3𝑂4/𝑍𝑛𝑀𝑛2𝑂4 + 2𝐻2𝑂 + 2𝑂𝐻− (1.9)

The following charge step includes the oxidation of Mn(OH)2 to β-MnOOH (layered structure), γ-

MnOOH and γ-Mn2O3 (spinel structure) during the first electron transfer. Charging continues with

the formation of δ-MnO2 upon the second electron transfer. The layered δ-MnO2, with large

interlayer spacing, hosts cations and structural water to stabilize its crystal structure75. The

electrochemically inactive Mn3O4/ZnMn2O4 can only be partially reduced to Mn(OH)2 and the rest

remains inert in the electrode, resulting in a significant loss of capacity. This does progress over

multiple cycles, generating more of the inactive spinel phase and eventually leading to failure. The

presence of zinc accelerates this process, as ZnMn2O4 is less electrochemically active than Mn3O4. A

similar reaction mechanism has been reported elsewhere74,76,77. The formation of Mn3O4/ZnMn2O4

limits the rechargeability of alkaline manganese dioxide based batteries78.

In mildly acidic electrolytes, a different reaction mechanism takes place. The electrochemically

inactive spinel phase is suppressed because the pH is buffered below 6 by basic salts precipitating.

These electrolytes containing zinc sulfate have been reported to deliver long cycle life and excellent

battery performance60,64,79. The proposed discharge reaction mechanism of EMD in zinc sulfate

electrolytes can be described as the co-insertion of protons and zinc ions72,79,80. Upon the first stage

of discharge, protons are inserted into the EMD structure, leading to an increased local pH at the

electrode surface. With continuous pH increase, the second discharge regime proceeds where zinc is

being inserted, precipitating zinc hydroxide sulfate pentahydrate, Zn4SO4(OH)6·5H2O:

4𝑍𝑛2+ + 𝑆𝑂42− + 6𝑂𝐻− + 5𝐻2𝑂 → 𝑍𝑛4𝑆𝑂4(𝑂𝐻)6 ⋅ 5𝐻2𝑂 (1.10)

This does not involve any electron exchange and occurs at pH 561. The formation of the precipitate

buffers the pH and thus prevents the formation of Mn3O4/ZnMn2O4, which would be formed at

higher pH. However, it is electrochemically inactive and forms large crystalline flakes on active

particles. These flakes block active sites, plug pores, and impede mass transport. Thus, preventing

the precipitate from insulating the surface is critical for maintaining reversibility. Another challenge

is manganese dissolution, which generates manganese vacancies where zinc ions can be inserted.

Excessive zinc-ion insertion into EMD leads to structural collapse and to cell failure72.

1.4.5. Iron

Iron is the second most abundant metal on Earth25 and is the electrode material for large-scale

aqueous rechargeable batteries. Iron is cheap and energy dense—both enviable properties for

batteries. Thomas Edison first developed the iron electrode in the early 20th century for the Ni–Fe

battery27. Later, significant interest in developing Fe–air batteries for fossil-free traction arose in the

late 1960s at NASA81. Fe–air batteries gained serious interest because of their remarkably high

theoretical energy densities, up to 9700 Wh L–1, and specific energies of more than 1200 Wh kg–1 24.

Later pioneering work by the Swedish National Development Company in the 1970s demonstrated

the feasibility of this battery, with a lifetime over 1000 cycles and an energy density of 80 Wh kg–1

82,83. Renewed interest in the iron electrode has led to recent advances in nanostructured materials

and has improved their performance further, resulting in higher energy densities84,85.

Page 22: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

9

Iron electrodes use aqueous alkaline electrolytes because of their high ionic conductivity and the

availability of reversible redox reactions24. The main discharge product, iron (II) hydroxide Fe(OH)2, is

insoluble. This favors the iron electrode because a solid-state reaction ensues and prevents the

diffusion of dissolved species86. Conventional iron-based batteries limit deep discharging and

operate upon the first two-electron reduction of iron to iron (II) hydroxide. This reaction offers a

theoretical specific capacity of 960 mAh g–1. The deep discharge regime involves further reduction to

the less electrochemically active magnetite, Fe3O4, with a theoretical specific capacity of 199 mAh g–

1. When recharged, the iron (II) hydroxide or magnetite transforms back to metallic iron. The charge

and discharge reactions can be described as87,88:

𝐹𝑒(𝑂𝐻)2 + 2𝑒− 𝐹𝑒 + 2𝑂𝐻− E° = –0.88 V vs. SHE (1.11)

𝐹𝑒3𝑂4 + 4𝐻2𝑂 + 2𝑒− 3𝐹𝑒(𝑂𝐻)2 + 2𝑂𝐻− E° = –0.76 V vs. SHE (1.12)

1.4.6. Electrochemical challenges of iron

Even though iron electrodes are robust and have been used for over a century27,82, long-term

inefficiency and unwanted side reactions have limited their large-scale use. The primary limitation is,

as for the zinc electrode, the HER (1.4)89. Unfortunately, iron is a good hydrogen evolution catalyst

leading to inadequate charging efficiencies in the range of 55–70%27. As such, significant amount of

water is lost, and superfluous hydrogen gas is evolved. Sulfide compounds are added to suppress the

HER and uphold efficient operation. Adsorbed sulfide on iron poisons the HER and is incorporated

into the electrode as FeS with Bi2O3, or Bi2S3 by itself, or in the electrolyte as Na2S or K2S89–91.

Moreover, it is critical not to operate at deep discharge regimes and encourage formation of the less

active phases Fe3O4 and Fe2O3. Utter attention is required to limit the depth-of-discharge (DOD) and

avoid passivation.

1.4.7. Oxygen electrocatalysts

Rechargeable aqueous metal–air batteries such as Zn–air92 and Fe–air24 attract research interest

because of their exceptionally high energy density. The air electrode uses bifunctional catalysts, i.e.

substances that can catalyze both the oxygen-evolution reaction (OER) and the oxygen-reduction

reaction (ORR). Metal–air batteries need an open cell design to deliver oxygen to the catalytic sites.

Oxygen is fed from an external source of air from the outer atmosphere, hence the name “air

electrode”. The air is not stored in the cell, which therefore exhibits notably high theoretical specific

energy density93.

Air electrodes require a three-phase boundary between the solid electrode in contact with the ion

conducting electrolyte and gas phase. To satisfy the three-phase interface, air electrodes are

designed with an open structure to facilitate gas diffusion while combining hydrophilic and

hydrophobic properties in separate layers. The hydrophilic layer ensures proper wetting and contact

with the aqueous electrolyte. The hydrophobic counterpart, commonly with a wet-proofing agent

such as polytetrafluoroethylene (PTFE), prevents electrolyte penetration and facilitates oxygen

diffusion to the catalytic sites92. Air electrodes can be adopted in different electrolytes, but alkaline

ones are favored because of rapid kinetics94. Alkaline electrolytes render possible the use of non-

precious-metal-based catalysts, most of which dissolve in acid. In these electrolytes, oxygen is

reduced to solvated hydroxide ions during discharge, and is regenerated upon charge. The ORR

Page 23: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

10

mechanism is complex because it involves a multistep electron transfer and follows either a four

electron- or a two electron pathway95,96. The reaction pathway varies with the catalytic material and

electronic structure97. For the direct four-electron pathway on metals, the reaction follows:

𝑂2 + 2𝐻2𝑂 + 2𝑒− → 2𝑂𝐻𝑎𝑑𝑠 + 2𝑂𝐻− (1.13)

2𝑂𝐻𝑎𝑑𝑠 + 2𝑒− → 2𝑂𝐻− (1.14)

Giving the overall reaction:

𝑂2 + 2𝐻2𝑂 + 4𝑒− → 4𝑂𝐻− E° = +0.40 V vs. SHE (1.15)

The alternate two-electron pathway proceeds through intermediate peroxide formation:

𝑂2 + 𝐻2𝑂 + 2𝑒− → 𝐻𝑂2− + 𝑂𝐻− (1.16)

𝐻𝑂2− + 𝐻2𝑂 + 2𝑒− → 3𝑂𝐻− (1.17)

Metal oxides follow the same pathways but with a different surface charge distribution. The metal

oxide cations on the surface are not completely coordinated with oxygen atoms but instead with the

oxygen of a water molecule98. The four-electron pathway is favored on precious metals, silver and

particular metal-oxide structures such as spinels and perovskites95. The two-electron peroxide

pathway dominates on carbon-based catalysts, gold and other metal-oxide structures98. The OER

mechanism is also complex. To simplify, oxygen is generally evolved from the oxide phase instead of

the metal and is followed by a release of two coordinated oxygen atoms to a metal ion on the

catalyst surface99. Transition metal oxides based on Ni, Co and Mn spinels100–103 and

perovskites97,104,105 have proven to be active bifunctional catalysts with good corrosion resistance in

alkaline electrolytes.

1.4.8. Electrochemical challenges of oxygen electrocatalysts

Bifunctional oxygen electrocatalysts are the main bottleneck of batteries that use air electrodes

because of slow kinetics and corrosion issues93. The sluggish kinetics are due to the ORR step. In

alkaline electrolyte, the competing displacement of O22–/OH– as well as hydroxide-ion conversion are

reported as the rate-limiting steps in the ORR97. The redox reaction upon operation yields large

polarization losses from high overpotentials. Overpotential is the magnitude of deviation from the

equilibrium potential and is constituted of activation, concentration, and resistance losses106:

𝜂𝑡𝑜𝑡𝑎𝑙 = 𝐸𝑐𝑒𝑙𝑙 − 𝐸𝑒𝑞 = 𝜂𝑎𝑐𝑡 + 𝜂𝑐𝑜𝑛𝑐 + 𝑖𝑅 (1.18)

where 𝐸𝑐𝑒𝑙𝑙 is the measured cell potential, 𝐸𝑒𝑞 is the equilibrium potential, 𝜂𝑎𝑐𝑡 is the activation

overpotential and defined as the required activation energy to proceed with the redox reaction,

𝜂𝑐𝑜𝑛𝑐 is the concentration overpotential and describes mass transport limitation by depletion of

charge carriers in the electrolyte at the electrode surface, and 𝑖𝑅 is the ohmic drop losses caused by

resistance in the hardware and electrolyte.

Figure 1.5 shows the total OER and ORR overpotential of an air electrode in alkaline electrolyte using

NiCo2O4 as bifunctional catalyst. Upon a full charge-discharge cycle, the total overpotential achieved

was 693 mV at a moderate current density rate of ±10 mA cm–2 and under a flow of air. This

overpotential is significant and prompts low voltaic and energy efficiencies. For instance, the energy

Page 24: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

11

efficiency of an alkaline Fe–air battery is 50% and the remaining part is unmitigated losses24. Active

oxygen electrocatalysts based on Pt, Pd, Ru and Ir encounter high total overpotentials as well; these

are typically above 500 mV92,93,107–109.

Figure 1.5. A full charge-discharge cycle at ±10 mA cm–2 of an air electrode using the bifunctional catalyst

NiCo2O4 in 6 mol dm–3 KOH. The figure highlights the total overpotential of OER and ORR. Air was used as the

oxygen feed to the air electrode. Figure reproduced from Paper V.

Polarization losses are a major challenge with oxygen electrocatalysts. The catalyst must sustain

oxidizing environments under high overpotentials. Furthermore, competition between the four- and

two-electron pathways of the ORR deteriorates the electrode, if the reaction mechanism favors the

latter. The two-electron pathway generates corrosive peroxide species, harming the electrode upon

battery operation110. Another concern with air electrodes in alkaline electrolyte is carbonate

formation when carbon dioxide reacts with hydroxide ions92:

𝐶𝑂2 + 2𝑂𝐻− → 𝐶𝑂32− + 𝐻2𝑂 (1.19)

The poorly soluble carbonates clog electrode pores and block the electrolyte channels, retarding the

electrochemical activity. To circumvent carbonate formation, it is important to purify the airflow or

use pure oxygen. Another option is to circulate the electrolyte in order to prevent the carbonate

from reaching supersaturation93.

1.5. The aim of the thesis

This thesis describes the investigation and development of rechargeable aqueous Zn–MnO2 and Fe–

air batteries to overcome hurdles in their performance. Mixed cation electrolytes containing KOH

and LiOH enhanced the cycle life and proved its potential as a drop-in electrolyte replacement for

traditional alkaline Zn–MnO2 batteries. The structural evolution and failure mechanisms were

investigated using electron microscopy and operando energy-dispersive X-ray diffraction (EDXRD)

techniques. In the analogous battery with mildly acidic electrolytes, the complex reaction

mechanism was explained to answer why perpetual access of Mn2+ ions at the electrode/electrolyte

interface enhanced the rechargeability. Differentiation of the phase evolution of cycled MnO2

electrodes used both ex situ and operando X-ray radiation analytical techniques.

ƞOER-ORR = 693 mV

Page 25: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

12

The examined hurdles for rechargeable Fe–air batteries were countered by suppressing the

evolution of hydrogen gas on the iron electrode and optimizing bifunctional catalysts for the air

electrode, in order to improve performance and cycle life. The evolution of hydrogen gas was

minimized using potassium stannate as an additive. The rationale behind the additive was that tin

metal would deposit as the iron electrode charged and serve as a barrier to hydrogen gas. Operando

neutron diffraction measurements described the phase evolution of the iron electrode with stannate

during charge-discharge cycling. Hydrogen evolution on iron, evaluated as a function of charge

current density, was studied by coupled electrochemistry and mass spectrometry. New bifunctional

catalysts for the air electrode were characterized structurally and electrochemically. The findings

presented an excellent catalyst candidate with superb activity for oxygen evolution and reduction

with excellent long-term stability. The catalyst was used in a rechargeable alkaline Fe–air battery as

a demonstration of this technology, and its effect on performance and cycle life is presented later.

This thesis aims to investigate the aforementioned battery systems and develop them as viable

technologies for large-scale energy storage. The motivation of this study emphasized rechargeable

batteries based on available resources to actualize the transition from fossil fuels to renewable

energy sources.

This thesis presents the background and motivation of the work, explains how experiments were

carried out and how analysis methods were applied, discusses the relevant results from Papers I–V,

and finally summarizes these results.

Page 26: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

13

2. Experimental 2.1. Synthesis

2.1.1. Bismuth-doped β-MnO2

The bismuth-doped β-MnO2 (MBDB) material was synthesized via the thermolysis of manganese and

bismuth nitrates. Two solutions were prepared separately before heat treatment: 1) 50 g of

Mn(NO3)2·4H2O (Sigma-Aldrich, >97%) in 80 mL deionized water and 2) 4.27 g of Bi(NO3)3·5H2O

(Sigma–Aldrich, >98%) in 18.6 mL deionized water and 6.4 mL nitric acid (HNO3, Sigma–Aldrich, 70%).

The solutions were mixed together and heated to 125 °C. Observable color changes of the solution

upon heating indicated the oxidation of manganese from 2+ to 4+, and the final solution was black.

The black solution was heated under vacuum at 125 °C for 12 h and the solid residue then baked at

325 °C for 5 h in air. Lastly, the solid material was ground into a fine powder.

2.1.2. Oxygen electrocatalysts

Three materials were synthesized as of oxygen electrocatalysts: 1) La1–xCaxMnO3 (LCMO), 2) Ni–Fe

layered double hydroxide (LDH) and NiCo2O4. For the LCMO catalyst, a solution of La, Ca and Mn

nitrates (VWR Chemicals, >99%), in the molar ratio 0.1:0.9:1, was first prepared. The solution was

then added dropwise to a heated solution of 0.2 M Na2CO3 (Sigma-Aldrich, >99.5%) at 50 °C,

resulting in a precipitate. The precipitate was washed with deionized water, filtered, and dried

before being calcined at 700 °C for 6 h in air. After calcination, the solid was quickly quenched in a

water-cooled zone of the furnace. Lastly, the solid material was washed with 10% aqueous acetic

acid, rinsed with deionized water, dried overnight at 150 °C and then ground.

Synthesis of the Ni–Fe LDH catalyst consisted of adding a reducing agent to a mother solution

containing nickel and iron nitrates. The mother solution consisted of 8 mL of 0.5 M Ni(NO3)2·6H2O

(Merck, >97%), 2 mL of 0.5 M Fe(NO3)3·9H2O (Pro Analysi, >98%) and 0.50 g of polyvinylpyrrolidone

(PVP, Fluka, K30, Mw = 40 000). This solution was mixed to ensure full dissolution of the compounds

and then transferred to a beaker with 50 mL deionized water. The reducing agent solution contained

1 g NaBH4 (Sigma-Aldrich, >98%) in 20 mL deionized water, and was added dropwise to the mother

solution. Mixing of the total solution continued overnight, resulting in a precipitate. The precipitate

was collected, filtered, and washed with both deionized water and ethanol before drying at 60 °C for

12 h.

Preparation of the NiCo2O4 catalyst comprised thermal decomposition of 3.62 g Ni(NO3)2·6H2O

(Merck, >97%) and 7.26 g Co(NO3)2·6H2O (Sigma-Aldrich, >98%) in 200 mL deionized water at 375 °C

for 2 h. After heating, the solid material was cooled to room temperature and then ground to a

powder.

2.2. Cell preparation and electrochemical characterization

2.2.1. Alkaline Zn–MnO2

After synthesis of the MBDB active material, electrodes were prepared by mixing 45 wt% MBDB, 45

wt% graphite (Timcal, KS6) and 10 wt% PTFE (Sigma-Aldrich, 60 wt% dispersion in H2O). The

electrodes were worked into pastes, dried in a vacuum oven at 125 °C for 1 h, and then pressed onto

a perforated nickel wire screen at 98 MPa. The pressed MBDB electrodes were assembled into

Page 27: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

14

planar cells against zinc foil as counter electrode and separated with one layer of polyvinyl chloride

(PVC) and two layers of non-woven cellulose (Freudenberg LLC, FV-4304). Acrylic plates with screws

held the cell under compression. After assembly, the cells were immersed in a beaker of 5 mL

electrolyte. The cycling protocol used constant current rates of 205, 31 or 147 mA g–1, with cells

connected to a battery analyzer instrument (MTI, BST8-3). The voltage range was kept between 0.4

and 1.8 V vs. Hg/HgO (+0.098 V vs. SHE) with a constant voltage step of 1.8 V during charge until the

current dropped below 10% of the input value.

2.2.2. Aqueous sulfate Zn–MnO2

Characterization of aqueous sulfate Zn–MnO2 cells used two electrolytes, 2 M ZnSO4 (Mn2+-free

electrolyte) and 2 M ZnSO4 + 0.1 M MnSO4 (added-Mn2+ electrolyte). Preparation of MnO2 electrodes

consisted of mixing 0.80 g EMD (Tronox, Ultrafine), 0.02 g graphite (Timcal, BNB90), 0.09 g carbon

black (Imerys, Super C65), and 0.09 g polyvinylidene fluoride (PVDF, Arkema, KynarFlex 2801) with 2

mL N-Methyl-2-pyrrolidone (NMP, VWR) solvent. This mixture was ball-milled (SPEX, 8000 Mill) for

20 min to form a slurry and then cast as a thin film on carbon paper (Freudenberg, H23) to a

thickness of 0.18 mm. Afterwards, the electrodes were dried in two steps, first at 60 °C for 3 h and

then at 120 °C for 12 h under vacuum. For the counter electrode, a slurry prepared in similar fashion

was casted on zinc foil (Alfa Aesar, 0.25 mm thickness) before drying at 60 °C overnight. The zinc

slurry composition consisted of 0.80 g zinc powder (Sigma–Aldrich, <10 µm, >98%), 0.10 g activated

carbon (Merck, Activated charcoal for analysis), 0.05 g graphite (Timcal, BNB90), 0.05 g PVDF

(Arkema, KynarFlex 2801) and 1.2 mL NMP. One layer of glass fiber paper (Whatman Grade GF/F)

separated the electrodes and the cell was contained between two acrylic pieces before it was

immersed in a container with 5 mL electrolyte. Cyclic voltammetry (CV) measurements were

performed on a potentiostat (BioLogic, SP-50) at potentials between 1 and 1.8 V vs. Zn/Zn2+ using a

sweep rate of 0.2 mV s–1. Charge-discharge cycling was done by connecting the cell to a current

source (Wuhan LAND Electronics, CT2001A) at a constant current rate of 60 mA g–1. Voltage cut-offs

were set at 1 and 1.8 V with a constant voltage step at 1.8 V during charge and ended when the

current dropped below 20% of the input value.

2.2.3. Iron electrode

The iron electrodes used in the stannate-additive study consisted of 80 wt% iron (Höganäs AB,

Nutrafine RS), 5 wt% Bi2S3 (Sigma–Aldrich, 99%), 8 wt% graphite (Imerys, KS6L), 2 wt% carbon black

(Imerys, Super C65) and 5 wt% PTFE (Sigma–Aldrich, 60 wt% dispersion in H2O). After mixing, the

formed paste was rolled to a thickness of 0.1 mm, dried at 110 °C for 1 h and then pressed into a

nickel wire screen (Dexmet, 100 mesh) at 30 MPa. The cell consisted of a commercial sintered nickel

electrode (Gates Energy) as counter electrode and Ag/Ag2O as reference electrode (+0.242 V vs.

Hg/HgO or +0.098 V vs. SHE)111. The measured half-cell potentials were converted to be relative to

that of Hg/HgO. One layer of non-woven cellulose (Freudenberg, 700/18F) separated the electrodes

and the cell was contained between acrylic plates before being submerged in 30 mL 6 M KOH + 1 M

LiOH electrolyte, with or without 0.1 M K2SnO3 included. The electrodes were cycled using a current

source (Wuhan LAND Electronics, CT2001A) operated at a constant current rate of 192 mA g–1.

Voltage cut offs were set to –0.458 and –1.158 V vs. Hg/HgO.

Nanostructured copper- and/or tin-doped iron materials, denominated as CuSn and Sn, were

provided by Höganäs AB. Paper IV details the structural and elemental composition analysis of these

Page 28: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

15

powders. The two iron-electrode materials were combined with single-walled carbon nanotubes

(SWCNT, OCSiAl, TUBALL™ BATT). Another sample investigated the effect of 0.65 M LiOH added to

the 6 M KOH electrolyte. In total, four samples were evaluated: 1) CuSn, 2) CuSnCNT, 3) CuSnCNTLi

and 4) SnCNT. Preparation of the electrodes involved mixing 80 wt% of the nanostructured iron

materials with 5 wt% carbon black (AkzoNobel, Ketjenblack EC-300J), 5 wt% Bi2S3 (Sigma–Aldrich,

99%) and 10 wt% PTFE (Sigma–Aldrich, 60 wt% dispersion in H2O). The mixture was homogenized in

laboratory blender (Waring, LB20ES) at 6000 RPM for 15 min in an aliphatic solvent (Shell Chemicals,

ShellSol D70). Afterwards, the filtered wet mass was rolled on a nickel wire screen (100 mesh) to a

thickness of 0.7 mm, pressed at 375 kg cm–2 and then sintered at 325 °C for 30 min. Paper IV shows

the active mass loading of iron in each sample. The nanostructured metal-doped iron material

containing copper and with SWCNT was adapted to the Fe–air prototype in Section 2.2.5.

2.2.4. Air electrode

Preparation of the air electrodes covered three parts: a catalyst layer, a current collector, and a gas

diffusion layer (GDL). Preparation of the catalyst layers were done for three sets of samples: 1) 65

wt% LCMO + 10 wt% Ni–Fe LDH, 2) 65 wt% LCMO + 10 wt% NiCo2O4 and 3) 75 wt% NiCo2O4, with 10

wt% carbon black (Imerys, Super C65) and 15 wt% PTFE (Sigma–Aldrich, 60 wt% dispersion in H2O),

constituting the rest for all three samples. The mixtures were homogenized in 20 mL ethanol per 1 g

of solids using an ultrasonic probe (Hielscher, UP200St) at 30% amplitude for 10 min. After mixing

and filtration, ethanol was added to the collected cakes to form pastes that were rolled to a

thickness of 0.4 mm. The rolled pastes were then pressed at 375 kg cm–2 onto a nickel wire screen

(Dexmet, 100 mesh) that served as current collector. Lastly, the electrodes were sintered at 340 °C

for 25 min before being cold-pressed onto the GDL, a porous PTFE foil (Guarniflon, TPF020), at 375

kg cm–2. Paper V provides the active mass loadings of the catalysts used in the prepared electrodes.

Electrochemical characterization of the OER and ORR in 6 M KOH used a specially designed cell

connected to a potentiostat (Bio-Logic, SP-50). Nickel mesh served as the counter electrode and

Ag/Ag2O as the reference electrode while the recorded half-cell potentials were converted to the

Hg/HgO reference. The cells were submerged in 50 mL of electrolyte and wetted the catalyst layer,

while the GDL layer was preserved dry with air flowing at a rate of 20 mL min–1. The active geometric

surface area was 4 cm2 and reported current densities were based on this value. Air electrodes were

preconditioned by CV over 20 cycles between –0.108 and 0.592 V at 5 mV s–1. Then, the cycle life of

these electrodes was assessed by charging and discharging for 2 h in steps at ±10 mA cm–2.

2.2.5. Fe–air prototype

Paper V details the requirements that must be considered when fabricating the Fe–air prototype to

enable stable cell operation. In short, resistant cell materials must withstand corrosive

environments, facilitate both electrolyte and gas flow, and secure the mechanical integrity and

sealing of the cell. Figure 2.1 shows the cell breakdown of the Fe–air prototype. Both stainless steel

(SS) and PVC end plates tighten the cell body, which has total dimensions of 12 10 3 cm. Inlets

and outlets for gas and electrolyte were fitted on the front. Within the cell, one iron and one air

electrode were aligned in parallel and separated by two polyethylene mesh spacers (PE, 1.35 mm

thickness, 3.4 3.2 mm mesh size). Another PE mesh spacer was placed on the GDL side of the air

electrode, facilitating gas transport to the catalytic sites. Ethylene propylene diene monomer rubber

gaskets (EPDM, Kuntze, ESO2 425-010, 0.5 mm thick) sealed the cell, and contained holes to ensure

Page 29: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

16

gas and electrolyte flows to the electrodes. The cell contained 10 mL electrolyte and had an active

geometric surface area of 49.2 cm2. Spot-welded nickel-foil tabs served as external connections.

Figure 2.1. Cell breakdown of the Fe–air prototype components. Figure reproduced from Paper V.

The flow-assisted Fe–air prototype cell was operated using copper-doped iron for the iron electrode

and NiCo2O4 as the bifunctional catalyst for the air electrode. The 6 M KOH electrolyte was circulated

through a closed system at 1 mL min–1 by a peristaltic pump (LKB Bromma, 2132 MicroPerpex).

Oxygen feed to the cell flowed at the rate of 35 mL min–1. Upon operation with electrolyte and gas

flows, the cell was connected to a current source (Wuhan LAND Electronics, CT2001B) and the

potential was recorded from the full cell. The cell operation protocol included three steps: 1) a

formation step to activate both electrodes at ±4 mA cm–2, 2) a rate-capability step at current

densities between ±5 and ±25 mA cm–2, and 3) a cycle-life-assessment step at ±10 mA cm–2. For the

first two steps, the cell was charged to the theoretical specific capacity of iron, 960 mAh g–1, whereas

in the last step, the charge capacity was optimized to maximize efficiency.

2.2.6. Coulombic efficiency

Detection of gaseous products with a specially designed electrochemical cell coupled to a mass

spectrometer (MS), as described in previous work112, enabled the quantification of hydrogen gas

evolved at iron and zinc electrodes. The measurements assumed that the HER was the only cause of

deviation from 100% Coulombic efficiency. Coulombic efficiency is defined as the ratio of total

discharge capacity output from the cell to the total charge capacity input into the cell over a full

cycle. The cell consisted of two separated chambers filled with electrolyte and with volumes of 48

cm3 each. Both chambers were purged continuously with argon to exhaust accumulated gas. From

each chamber, the exhaust was collected into a sampling point for the MS (Pfeiffer, Thermostar

GSD320-QMG220) to detect the gaseous products. Evaluated iron electrodes in alkaline electrolyte

were cycled against an oversized commercial nickel electrode (Gates Energy). The iron electrodes

were activated prior to the efficiency analysis, and then fitted into the setup to quantify the amount

of hydrogen gas. The electrodes were cycled using a potentiostat (BioLogic, SP-50) at charge current

densities of 5–15 mA cm–2 while the discharge was kept constant at 5 mA cm–2. In the case of zinc

electrodes, symmetric cells were used in mildly acidic electrolytes at current densities of 1–100 mA

cm–2. The geometric surface areas of the zinc working and counter electrodes were 1 and 9 cm2,

respectively.

Page 30: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

17

2.3. Operando techniques

2.3.1. Energy-dispersive X-ray diffraction (EDXRD)

Operando EDXRD enabled the MnO2 phase transformations that occurred during cell operation in

alkaline electrolytes to be distinguished. A 3D printed cell holder (Formlabs, Form 1 3D printer) was

prepared and from a resin that was stable in concentrated alkaline solution and highly transparent

to X-rays. The cell had an open design with no compression. It was filled with 1 M KOH + 3 M LiOH as

electrolyte and cycled against a zinc foil as the counter electrode. The cell was cycled at 147 mA g–1

for both charge and discharge, using a battery cycler (MTI, BST8-3). Data collection was conducted at

the National Synchrotron Light Source on the beamline X17B1 at Brookhaven National Laboratory.

Prior to running the cell, the incident X-ray beam and detection at a fixed angle of 2θ = 3° were

aligned. The incident beam consisted of white-beam radiation in the energy range of <20 to 200 keV

and diffracted X-rays were collected with a cryostat-cooled Ge detector (Canberra). The collected X-

ray signals were digitally processed using an 8192-channel analyzer. Data collection points occurred

every 1 min as the cell cycled. Calibration of the X-ray energy used LaB6 and CeO2 as standards.

Paper I further details the positioning of the incident beam and detector as well as the complete

scan procedure.

2.3.2. Neutron diffraction

Operando neutron diffraction measurements were carried out to distinguish the phase evolution of

an iron electrode with stannate, which used a cell design described elsewhere113. A 6 M KOD + 0.1 M

K2SnO3 electrolyte was used to minimize the incoherent scattering from H nuclei. Therefore, on the

nickel-mesh counter electrode, O2 and D2 gas evolved upon cycling. The iron electrode was prepared

in similar procedure as described in Section 2.2.3 but on a larger scale (1.25 g of iron) to achieve

satisfactory intensities of diffracted neutrons. Before data collection, the iron electrode was

activated by charge-discharge cycling until it reached a stable discharge capacity. Time-of-flight

neutron diffraction data collection was performed at the ISIS pulsed spallation neutron source using

the Polaris diffractometer114. Collection time per data point was set to 60 min to ensure good

statistical quality. The cell discharged at 96 mA g–1 for 5 h and charged at 192 mA g–1 for 4 h, with a

30-min rest step in between. Rietveld analysis of the collected data used the GSAS software and data

from the low-angle detector bank (40° < 2θ < 67°, dmax = 7 Å, Δd/d = 0.86%). The Bragg peak profiles

were described by a pseudo-Voigt function although only the Gaussian width of the function was

refined115.

2.3.3. X-ray diffraction (XRD)

Modified pouch cells, based on cells used in other reported work116, were constructed for operando

XRD measurements to identify the phase evolution of MnO2 in aqueous sulfate Zn–MnO2 cells. The

pouch cell had 13-mm holes punched on both sides to avoid contribution of the polymer-coated

aluminum foil to the XRD patterns, and these were sealed with Kapton tape. The zinc counter

electrode also contained a 5-mm hole to allow the incident X-ray beam window to be focused on the

MnO2 electrode. The incident beam passed through the following components: Kapton tape,

electrolyte, glass fiber paper and the MnO2 electrode. Figure 2.2 shows a schematic of the pouch

cell.

Page 31: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

18

Figure 2.2. Schematic of the operando XRD pouch cell components with incident X-ray beam direction

illustrated. Figure reproduced from the supplementary information for Paper III72 with permission from

Elsevier.

Data collection upon phase transformations in the MnO2 electrode used an in-house single crystal

diffractometer (Bruker AXS) with Mo source and a CCD detector (2D Apex II). XRD patterns were

collected in transmission mode every 10 min using a 0.5 or 0.8-mm beam collimator. The cell was

charged and discharged at 30 mA g–1, including a constant-voltage step at 1.8 V during charge, until

the current dropped below 10% of the input value.

2.4. Scanning electron microscopy (SEM) and energy-

dispersive X-ray spectroscopy (EDS)

SEM micrographs were captured from secondary electrons for visualizing the sample morphology

and particle size, while EDS provided information on the sample surface composition. SEM and EDS

investigations were performed for the MBDB, oxygen electrocatalysts, MnO2 electrodes after cycling

in aqueous sulfate electrolytes and iron electrodes after cycling in the stannate-additive study. SEM

micrographs of the synthesized MBDB were captured using an environmental scanning electron

microscope (ESEM, Philips XL 30) operated at 20 kV. SEM images of the oxygen electrocatalyst and

cycled MnO2 and iron electrodes were recorded using a field emission microscope (JEOL, JSM-

7001F), operated at 15 kV, and integrated with EDS capabilities to analyze the elemental

composition of the materials. Paper IV includes SEM images of the nanostructured copper- and/or

tin-doped iron materials.

2.5. XRD

XRD patterns were mainly used for phase identification and structural analysis. Ex situ XRD patterns

were collected on MBDB and the oxygen electrocatalyst materials. The XRD pattern of MBDB was

acquired using a Bruker D8 Advance diffractometer operated in Bragg–Brentano θ-2θ geometry and

Page 32: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

19

using a Cu Kɑ1 source, λ = 1.54056 Å. The XRD patterns of the oxygen electrocatalysts were acquired

using a PANalytical X´Pert Pro diffractometer, in Bragg–Brentano θ-θ geometry, and a Cu Kɑ source,

λ = 1.54056 Å. Papers II, III and IV include XRD patterns of the iron electrodes used in the stannate

additive study, MnO2 electrodes used in aqueous sulfate electrolyte, and nanostructured copper-

and/or tin-doped iron materials, respectively.

2.6. Electron energy loss spectroscopy (EELS)

EELS provided structural information of cycled MnO2 electrodes in aqueous sulfate electrolytes. EELS

data collection used a Schottky field emission electron microscope (JEOL, JEM-2100F) operated at

200 kV, with a probe size of 0.7 nm and a camera length of 2 cm. The microscope had an integrated

CCD camera (Gatan, Ultrascan 1000), post-column energy filter (Gatan Imaging Filter, Tridiem 863),

and a Gatan annular dark field detector. MnO2 electrode samples cycled in aqueous sulfate

electrolyte were washed, dried, and then scratched off the carbon paper with a diamond scriber and

collected on a holey carbon support film (SPI, 300 mesh Cu). Acquired EELS spectra covered the Mn

L2,3 edges (651 and 640 eV, respectively), the O K edge (530 eV) and the Zn L2,3 edges (1043 and 1020

eV, respectively). These spectra were recorded in the scanning transmission electron microscopy

mode with applied spatial drift correction and subpixel scanning modality. All EELS spectra were

acquired at inelastic mean free path regions between 0.3 and 0.8. Full width at half maximum of the

zero-loss peak was 1.20 eV. Multiple scattering effects in the EELS spectra were removed by Fourier

ratio deconvolution of low-loss features and the background fitted by a power-law model.

Estimation of the Mn oxidation states was based on integrated L3/L2 intensity ratios using a double

arctan continuum model117,118 as detailed in other work119,120.

2.7. X-ray photoelectron spectroscopy (XPS)

Surface sensitive XPS analysis was used to quantify tin in the cycled iron electrode with stannate. XPS

analysis used a ULVAC-PHI Versaprobe 5000 spectrometer with monochromatic Al Kɑ line (1486.6

eV) as the X-ray source. The XPS operated at a nominal power of 25 W with a beam size of 100 µm

and at a 45° angle while the sample was charge-neutralized of both low-energy electrons and Ar+

ions. Calibration used two reference points for the spectrometer energy scale: Au 4f7 (84 eV) and Cu

2p3 (932.7 eV). Detection of tin was performed on iron electrodes that had been cycled with

stannate by measuring the Sn 3d spectrum (484–498 eV) to a depth of 5 nm with an energy pass of

23.5 eV and step size of 0.1 eV. Paper II includes the deconvolution of the Bi, F, C, and O XPS spectra.

Page 33: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

20

3. Results and discussion

The results and discussion in this thesis are a synopsis of those given in Papers I–V. Paper I covers

the detailed synthesis, composition, structure characterization and electrochemical properties of

bismuth doped β-MnO2 in electrolyte mixtures containing multiple cations for rechargeable Zn–

MnO2 alkaline batteries. Paper III elucidates the reaction mechanism of rechargeable aqueous

sulfate Zn–MnO2 batteries by structural and electrochemical characterization. Paper II deals with the

effect of stannate on rechargeable alkaline iron electrodes by electrochemical, morphological and

phase evolution characterizations. Paper IV explores nanostructured doped iron materials for high-

performance rechargeable iron electrodes and gives insight into optimal cell operation. Lastly, Paper

V investigates proper bifunctional oxygen electrocatalysts for air electrodes and adapts one of them

to an Fe–air battery in order to prove the feasibility of this technology.

3.1. Cation electrolyte mixtures for rechargeable Zn–MnO2

alkaline batteries (Paper I)

3.1.1. Structural characterization of bismuth-doped β-MnO2

(MBDB)

Figure 3.1 shows the XRD pattern and SEM image of the synthesized bismuth doped β-MnO2 (MBDB)

material. The reflections at 2θ = 31–34° were from Bi2O3 (ICSD 417638). This 2θ region was excluded

in the Rietveld refinement to improve the precision of the fit. The remaining reflections matched

well with those from β-MnO2 (ICSD 73716). The stoichiometric ratio of Bi:Mn corresponding to 3.5%

of the manganese atoms in the (0,0,0) position replaced with bismuth was used in the refinement,

while oxygen atoms were located in the (0.3046,0.3046,0) position, with full occupancy, yielding Rp =

9.8% and Rwp = 15.4%. The synthesized MBDB material was produced as fine interconnected

particles in the size range of 50–100 nm, as shown in Figure 3.1b.

Figure 3.1. a) Ex situ XRD pattern of MBDB including observed data (black), calculated pattern (red), ICSD peak

positions for β-MnO2 (green), and the difference between raw data and fitted model (blue). b) SEM image of

synthesized MBDB. Figure reproduced from Paper I69 with permission from American Chemical Society.

Page 34: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

21

3.1.2. Electrochemical performance of KOH:LiOH electrolytes

Electrochemical performance of the alkaline Zn–MnO2 cells was analyzed for mixed-cation

electrolytes with varied KOH and LiOH concentrations. Figure 3.2a shows discharge capacity results

for a wide composition range of electrolyte mixtures measured at 205 mA g–1 to identify optimal

mixture concentrations. The cells containing 4:0, 3:1 and 2:2 mixtures of KOH:LiOH (total

concentration 4 M) had similar cycle trends. These cells attained high discharge capacities of 230–

370 mAh g–1 in the first cycle but these values quickly declined to 50–100 mAh g–1 in subsequent

cycles. Meanwhile the 1:3 and 0:4 KOH:LiOH mixtures sustained higher capacities over 10 cycles. The

4 M LiOH cell failed after 10 cycles, however, because of passivation of the zinc electrode as ZnO is

less soluble in the weakly basic pure LiOH electrolyte121. The 1 M KOH + 3 M LiOH cell cycled over 50

times and showed less capacity fading during these cycles. Figure 3.2b shows results obtained in

electrolyte mixtures with concentration ratios close to 1:3 and tested at slower rates, with the first

cycle at 31 mA g–1 and subsequent cycles at 147 mA g–1. Superior cell performance was achieved

within the approximate range of 1:3 to 3:7 KOH:LiOH ratios. At the higher ratio of 1:2 or the lower

ratio of 1:4, the capacity dropped significantly after the first cycle, to 110–80 mAh g–1. This quick

decay in performance can be explained by a disfavored proton insertion mechanism when the

KOH:LiOH ratio was too low and the favored formation of ZnMn2O4 when the ratio was too high.

These results indicate that an optimized KOH:LiOH ratio can improve the overall cycle life of Zn–

MnO2 cells. Paper I elucidates the electrochemical effect of bismuth doping in β-MnO2 of cells cycled

in 1 M KOH + 3 M LiOH electrolyte. In short, bismuth affected the second-electron discharge region

by complexing with dissolved manganese species of Mn(OH)23– or Mn(OH)2

4– and inhibited the

formation of Mn3O4/ZnMn2O4122. This improved the cyclability of Zn–MnO2 cells in 1 M KOH + 3 M

LiOH, where discharging occurred by simultaneous proton and lithium insertion into the MnO2

structure.

Figure 3.2. MBDB cells cycled against zinc metal in different KOH:LiOH electrolyte mixtures, at rates of a) 205

mA g–1 and b) 31 mA g–1 for the first cycle and 147 mA g–1 for subsequent ones. Figure reproduced from Paper

I69 with permission from American Chemical Society.

3.1.3. Phase evolution investigation of MBDB electrodes

The details of phase evolution in MBDB electrodes upon cycling in 1 M KOH + 3 M LiOH were studied

using both ex situ and operando techniques. Figure 3.3 shows ex situ XRD patterns of MBDB

Page 35: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

22

electrodes cycled once and stopped at the discharged (Figure 3.3c) and charged (Figure 3.3d) states.

The structural transformation that occurred upon the first full discharge reduced β-MnO2 into a

mixture of LiMn2O4 and Mn(OH)2. Observed shifts in the (111), (113) and (222) peaks of LiMn2O4 at

19.1, 33.9 and 44.9° evidenced lithium insertion. Concurrent proton insertion formed the other

product, Mn(OH)2. These reduced products demonstrated the plausible simultaneous lithium and

proton insertion into the β-MnO2 structure over the first full discharge. During the reoxidation step,

the original MBDB structure was not recovered. Instead, a new set of peaks emerged, especially the

noteworthy peak at 12°. The strongest emerging peaks can be assigned to the layered birnessite

structure (δ-MnO2)123. Bi2O3 peaks appearing at 31–34° for the pristine electrode disappeared during

discharge due to reduction of Bi2O3 to amorphous bismuth metal. The other components, PTFE,

carbon black and nickel, remained unchanged throughout cycling.

Figure 3.3. Ex situ XRD patterns of MBDB electrodes in various states of charge and discharge in 1 M KOH + 3

M LiOH: a) MBDB powder, b) pristine MBDB electrode, c) after the first full discharge and d) after the first full

recharge. Figure reproduced from Paper I69 with permission from American Chemical Society.

Figure 3.4 shows the operando EDXRD phase evolution investigation of MBDB, which revealed the

same reaction mechanism seen from the ex situ XRD measurements. MBDB was fully reduced to

Mn(OH)2 and LiMn2O4, and never recovered upon recharge. The Mn(OH)2 peaks appeared at 81.42,

95.2, 128.5 and 141.4 keV during discharge, while an increased intensity at 77.8 keV, close to the

graphite peak, corresponded to LiMn2O4. Interestingly, peaks stemming from metallic bismuth

developed during discharge and then faded away when bismuth dissolved into the electrolyte during

charge. The birnessite phase was not detected after the first cycle because of the limited energy

range of the EDXRD scan. The absence of the intermediate trivalent MnOOH phase, usually formed

during normal proton insertion71,73, may be explained by the favored lithium insertion that formed

LiMn2O4 instead. This may also be the key to the improved cyclability, as the irreversible

Mn3O4/ZnMn2O4 phases form through the trivalent intermediate. These irreversible phases cannot

be oxidized back to birnessite, but Mn(OH)2 can.

Page 36: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

23

Figure 3.4. Operando EDXRD measurements of a MBDB cell cycled against zinc metal in 1 M KOH + 3 M LiOH.

The colored vertical arrows indicate peaks associated with particular phases and the black horizontal arrows

indicate the end of charge and discharge of the MBDB electrode. Figure reproduced from Paper I69 with

permission from American Chemical Society.

Ultimately, mixed cation electrolytes containing KOH and LiOH enabled a competing proton- and

lithium-insertion mechanism in rechargeable alkaline Zn–MnO2 batteries. This improved the use of

MnO2 and upheld reversibility, while suppressing the inactive ZnMn2O4 phase from proton insertion

in favor for LiMn2O4.

3.2. Reversible aqueous sulfate Zn–MnO2 batteries with Mn2+

(Paper III)

3.2.1. Electrochemical characterization of MnO2 electrodes

Figure 3.5 depicts the electrochemical impact of Mn2+ in aqueous sulfate Zn–MnO2 cells by

comparing results for cells with 2 M ZnSO4 (Mn2+-free) and 2 M ZnSO4 + 0.1 M MnSO4 (added-Mn2+)

electrolytes. The cells used overdimensioned zinc metal counter electrodes to allow full investigation

of the MnO2 electrodes. The CV curves in a) and b) show the redox reactions that occur and the

distinction between cells. In both cases, the first reduction entailed a one-electron transfer at 1.21 V

while the subsequent ones had two peaks, at 1.25–1.28 and at 1.37–1.39 V. This shifting

electrochemistry from the first cycle to following ones was caused by the conversion of ɣ-MnO2 to a

layered hydrated MnO2 structure containing zinc. The two electron-transfer peaks after the first

cycle correlated to the insertion mechanisms of protons and zinc. The electron transfer

characteristics of both samples were analogous except that the added Mn2+ cell upheld reversibility

over 50 cycles. Figure 3.5c validates the reversibility of charge-discharge cycling in the cell that had

Mn2+ added to the electrolyte. Both samples showed increasing capacity over the first 15 cycles,

which can be explained by poor wettability of the electrode. Eventually, maximum capacities were

discharged

charged

discharged

charged

Page 37: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

24

reached and the added-Mn2+ cell maintained a stable capacity of 220 mAh g–1 over 100 cycles while

the Mn2+-free cell deteriorated. The Coulombic efficiencies were unaffected by the declining capacity

and remained over 98%. Figure 3.5d and e shows the charge and discharge profiles from these

measurements. The figures depict a significant capacity fade when no Mn2+ is added to the

electrolyte. The discharge over the last cycles in d) showed a declining capacity for the second

plateau; this was related to zinc insertion and altered the reversibility. The added-Mn2+ cell however

showed a stable performance that implied that Mn2+ restrained zinc insertion. The rate capability of

the added-Mn2+ cell shown in Figure 3.5f achieved discharge capacities of 233, 216, 188, 141, 97 and

76 mAh g–1 at 30, 60, 120, 300, 750 and 1500 mA g–1, respectively. Overall, the Zn–MnO2 cell offered

satisfactory rate capability.

Figure 3.5. Electrochemical performance of aqueous sulfate Zn–MnO2 cells in 2 M ZnSO4 (Mn2+ free) and 2 M

ZnSO4 + 0.1 M MnSO4 (added Mn2+). Cyclic voltammetry over 50 cycles at 0.2 mV s–1 of a) Mn2+-free and b)

added-Mn2+ electrolytes. c) Charge-discharge cycle summary of the two samples over 100 cycles at 60 mA g–1,

where the left y-axis corresponds to specific discharge capacity (blue markers) and the right y-axis to the

Coulombic efficiency (red markers). Charge and discharge profiles of the MnO2 electrodes in d) Mn2+-free and

e) added-Mn2+ electrolytes at 60 mA g–1. f) Rate capability analysis of the MnO2 electrode in the added Mn2+

electrolyte. All cells were cycled against zinc metal and the current and capacity were normalized based on the

active mass of MnO2. Figure reproduced from Paper III72 with permission from Elsevier.

3.2.2. Characterization of cycled MnO2 electrodes

SEM images and elemental analysis of the surface in Figure 3.6 show structural changes to the MnO2

electrodes after cycling. The pristine electrode in Figure 3.6a consisted of EMD, carbon black,

graphite and PVDF. After cycling in Mn2+-free electrolyte, and at a fully charged state, the MnO2

electrode shown in Figure 3.6b had a densified surface and contained a significant amount of zinc.

This indicated that zinc insertion occurred during cycling, and in a weight amount close to that of

manganese. Moreover, the dense surface decreased the porosity of the electrode structure and

inhibited electrolyte contact with active sites. The electrode cycled in an electrolyte with added Mn2+

had a fibrous morphology and contained less zinc than manganese at its surface. As will be explained

in a later section, zinc insertion into the MnO2 structure formed Zn1–xMnO2·nH2O. The availability of

Page 38: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

25

Mn2+ in the electrolyte affected the amount of zinc inserted. In Figure 3.6c and d, the detected B and

Si on the surface originate from the borosilicate glass paper. The glass paper tended to stick to the

surface because of the precipitate, Zn4SO4(OH)6·5H2O, which formed during discharge. The

precipitate adsorbed large amount of H2O and dehydrated the surface. Figure 3.6d shows the

precipitate as large hexagonal layered flakes 10–30 µm in size. The precipitate completely covered

the electrode surface and has low electrical conductivity. It was indeed surprising that the cell even

managed to cycle with such coverage, indicating facile dissolution of the precipitate during charge

via decreasing pH61.

Figure 3.6. SEM images and elemental analysis results (inset) for MnO2 electrodes in aqueous sulfate

electrolytes. a) Pristine MnO2 electrode composed of EMD, carbon black, graphite and PVDF. The MnO2

electrodes were cycled 100 times and stopped at a charge state in b) Mn2+-free and c) added-Mn2+

electrolytes. d) MnO2 electrode cycled 100 times and stopped at a discharged state in added-Mn2+ electrolyte.

In c), the highlighted blue color area corresponds to glass fiber paper and the red color area to carbon paper.

Figure reproduced from Paper III72 with permission from Elsevier.

EELS provided further structural analysis of cycled MnO2 electrodes at fully charged states. The

intensities in EELS spectra reflect the density of states and enabled a determination of the oxidation

state of manganese. Figure 3.7a shows representative EELS spectra of the O K edge, Mn L2,3 edges

and Zn L2,3 edges from the pristine MnO2 electrode and electrodes cycled in electrolytes with and

without Mn2+. All samples showed strong white line features at Mn L edges from the non-filled 3d

Page 39: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

26

conduction band117,118,124. The energy position and characteristics of Mn L2,3 edges were similar for

the uncycled electrode and the electrode cycled in electrolyte with added Mn2+; the former had an

L3/L2 intensity ratio of 2.31 ± 0.15 and the latter 2.42 ± 0.19. The intensity ratio of the pristine MnO2

deviated from the theoretical value 2 ± 0.1117,125 but it has been reported elsewhere that high values

are obtained for nanosized MnO2126. The slight increase in intensity ratio for the added Mn2+ sample

indicates a lower oxidation state of Mn, which correlates with the addition of zinc to the structure,

as detected in the EELS spectra shown in Figure 3.7b. By comparison, the sample that was cycled in

Mn2+-free electrolyte sample showed a small shift of the Mn L2,3 edges to lower energies. The

intensity ratio of the Mn L2,3 edges increased here to 3.05 ± 0.56, indicating an even lower oxidation

state of Mn than in the added Mn2+ sample. The O K edge characteristics of the samples were more

distinct. The “a” and “b” peaks for the uncycled electrode and the one cycled in added Mn2+ had

similar features. The “a” peak was located at 531.40 eV for both samples and the “b” peak was

found at 556.60 and 543.8 eV for the respective samples. However, another peak appeared between

these features after cycling in added Mn2+ electrolyte and is attributed to an additional state given by

zinc in the structure127. In the sample cycled Mn2+-free electrolyte, the features are in contrast with

those of the other samples and are labeled as “a*” and “c*”. These two peaks were located at 538.4

and 555.8 eV, respectively, and did not match with typical fine-structure features of manganese

oxides. Instead, the broad energy width and range covered by the “a*” peak suggested that it may

originate from mixed manganese oxides with different valence states. This sample contained a

significant amount of zinc, consistent with the EDS results shown above.

Page 40: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

27

Figure 3.7. a) EELS spectra covering the Mn L2,3 and O K-edges of the pristine and cycled MnO2 electrodes at

charge states in both electrolytes. The intensity of Mn L2,3 and O K-edges are normalized to the maximum

intensity of Mn L3. b) EELS spectra covering the Zn L2,3 spectra for the cycled MnO2 electrodes at charge states

in both electrolytes. The intensity of Zn L2,3 edge from the sample cycled in added-Mn2+ electrolyte is amplified

for visualization and not relative to the Mn2+-free sample. Figure reproduced from Paper III72 with permission

from Elsevier.

3.2.3. Progression of the MnO2 charge and discharge mechanism

Operando measurements of the phase evolution of MnO2 electrodes used a suitable designed beam

window in the modified pouch cell to collect XRD patterns in transmission mode every 10 min.

Reference patterns were measured for the individual cell components, EMD, carbon paper, glass

fiber paper and electrolyte, before running the cell. Figure 3.8 shows the reference patterns of these

components. The electrolyte had the highest background contribution and shifted during cycling

because of varying H2O volumes. The background was set to be the same when evaluating different

state-of-charge (SOC) patterns upon cycling.

Page 41: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

28

Figure 3.8. Reference XRD patterns of the electrolyte, MnO2 electrode, carbon paper, and glass fiber paper

from operando XRD data collection. The background was set to be the same throughout different SOC patterns

upon cycling. The patterns of the three given SOC at 0, 50 and 100% show the significant background shift.

Figure reproduced from the supplementary information for Paper III72 with permission from Elsevier.

Figure 3.9 shows the XRD patterns and charge and discharge profiles of MnO2 cells in the Mn2+-free

and added-Mn2+ electrolytes. The significant peak arising at 2θ below 10° during charge can be

assigned to the (001) reflection of the hydrated phyllomanganate, Zn-buserite. Zn-buserite, a layered

structure similar to that of birnessite (δ-MnO2), but with an interlayer spacing of 10 Å containing two

H2O molecules instead of one and Zn2+ balancing the negative charge in the octahedral layers128. The

intensities of the (001)b and (002)b reflections belonging to the Zn-buserite phase and their peak

positions differed in the cells containing electrolyte without and with Mn2+. The former had 2θ

values of 8.65 and 18.16° (d = 10.22 and 4.88 Å, respectively) while the latter had values of 9.61 and

18.43° (d = 9.20 and 4.81 Å, respectively). The larger d-spacing values can be explained by the

increased quantity of Zn2+ inserted into the interlayer region, which expands the unit cell22. The

remaining reflections highlighted correspond to Zn1–xMnO2·nH2O, which is present in two intergrown

phases: hexagonal pyrolusite (β-MnO2) and orthorhombic ramsdellite. In both cells, the evolved

phases during charge represented Zn1–xMnO2·nH2O and Zn-buserite, whereas ZnMnO2·nH2O and

Zn4SO4(OH)6·5H2O formed during discharge. In Figure 3.9a, the unit cell of Zn1–xMnO2·nH2O expanded

and contracted during cycling, as evidenced from the peak shifts to lower and higher 2θ,

respectively. The peak shifts suggested Zn2+ insertion to form ZnMnO2·nH2O and extraction back to

Zn1–xMnO2·nH2O. Similar operando phase evolution of zinc insertion into MnO2 has been reported

elsewhere45. In contrast, the cell containing added Mn2+ had smaller peak shifts, which indicated that

the crystal structure was more stable to minor unit cell volume changes than the cell without Mn2+.

For comparison, the (102)h and (110)h peak positions of Zn1–xMnO2·nH2O from fully discharged to

charged states of the Mn2+-free cell increased by 1.8 and 2.7%, respectively, whereas for the added-

Mn2+ cell they increased by 0.4 and 1.1%, respectively. The Zn-buserite peaks faded upon discharge,

suggesting dissolution of the compound. From these observations, the impact of added Mn2+ is

correlated to the quantity of Zn2+ insertion. The accessible Mn2+ suppressed dissolution and inhibited

Mn vacancies for Zn2+ insertion. Higher Zn2+ insertion may destabilize the crystal structure and

inevitably result in deteriorating capacity.

Page 42: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

29

The peaks that arose during discharge included those of the crystalline Zn4SO4(OH)6·5H2O. This

compound precipitated due to increased pH (lowered concentration of H+ during proton insertion) at

pH around 5.561,62 without any exchange of electrons. The initial pH values of the Mn2+-free and

added-Mn2+ electrolytes were 4.2 and 3.9, respectively. The precipitate directly influenced the cell

reversibility by buffering the pH around 5.5 and inhibited formation of inactive ZnMn2O4 at pH > 7129.

However, the precipitate has low electrical conductivity and could impede diffusion of ions to active

sites by blocking the surface and plugging the pores. Limiting the amount of this precipitate is

important to maintain the rechargeability of the aqueous sulfate Zn–MnO2 cell.

Figure 3.9. Operando phase evolution analysis by XRD of MnO2 electrodes upon charge and discharge.

Acquired XRD patterns in samples a) Mn2+-free and c) added-Mn2+ electrolyte at 30 mA g–1 with their

corresponding charge and discharge profiles in b) and d), respectively. Assigned diffraction indices with

subscripts correspond to the following phases: b = Zn-buserite, o = orthorhombic and h = hexagonal part of

Zn1–xMnO2·nH2O. In the XRD patterns, the blue line represents the fully charged state and red the fully

discharged state. Straight dashed orange lines display peak position shifts of the hexagonal reflections. Figure

reproduced from Paper III72 with permission from Elsevier.

To summarize, the MnO2 reaction mechanism in aqueous sulfate Zn–MnO2 cells can be described as:

𝑀𝑛𝑂2 + 4𝐻+ + 2𝑒− 𝑀𝑛2+ + 2𝐻2𝑂 (3.1)

𝑍𝑛1−𝑥𝑀𝑛𝑂2 ⋅ 𝑛𝐻2𝑂 + 𝑥𝑍𝑛2+ + 2𝑥𝑒− 𝑍𝑛𝑀𝑛𝑂2 ⋅ 𝑛𝐻2𝑂 (3.2)

Page 43: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

30

4𝑍𝑛2+ + 𝑆𝑂42− + 6𝑂𝐻− + 5𝐻2𝑂 𝑍𝑛4𝑆𝑂4(𝑂𝐻)6 ⋅ 5𝐻2𝑂 (3.3)

The reaction during the first discharge proceeded via proton insertion into EMD and dissolution to

Mn2+ (3.1). In subsequent cycles, the same mechanism took place, but with Zn-buserite dissolving

instead. From these reactions, the concentration of H+ decreased upon proton insertion and lead to

the precipitation of Zn4SO4(OH)6·5H2O (3.3). Simultaneously, Zn2+ insertion occurred where x denotes

the number of reversibly inserted cations (0 < x < 1) and formed ZnMnO2·nH2O (3.2). Paper III

describes a hypothesis for the intermediate Mn3+ reaction that is excluded here.

3.3. Effect of stannate on rechargeable iron electrodes (Paper

II)

3.3.1. Electrochemical and structural characterization of iron

electrodes

Charge and discharge profiles of iron electrodes cycled in alkaline electrolytes, shown in Figure 3.10,

illustrate the impact of stannate in the electrolyte. The cells were cycled over Fe0/Fe2.67+ and

included deep discharge regimes and the formation of Fe3O4. This regime is avoided in commercial

alkaline iron electrodes, where the discharge reaction is limited to Fe(OH)2 with an Fe oxidation state

of 2+. Both electrodes contained Bi2S3 as an additive to suppress the HER; this compound reduced

during charging130:

𝐵𝑖2𝑆3 + 6𝑒− 2𝐵𝑖 + 3𝑆2− E° = –0.92 V vs. Hg/HgO (3.4)

Metallic bismuth is deposited on the electrode surface while sulfide is adsorbed on iron and poisons

its activity towards hydrogen90. The discharge profile of the iron electrode in an electrolyte without

stannate (Figure 3.10a) had two plateaus. The first involved a two-electron oxidation to Fe(OH)2 at –

0.85 V and the second further oxidation to Fe3O4 at –0.70 V. The capacity increased successively over

30 cycles and stabilized thereafter around 400 mAh g–1. This increase in discharge capacity is a

known formation step that occurs as the electrolyte gradually penetrates to the core of the iron

particles89. During charge, metallic iron recovered and overcharged to 500 mAh g–1; the remaining

capacity was attributed to the HER. In the cell with stannate added, different charge and discharge

trends evolved as shown in Figure 3.10b. K2SnO3 dissolved to give stannate ions, Sn(OH)62–, that were

reduced to tin metal, which increased the HER overpotential during charge131:

𝑆𝑛(𝑂𝐻)62− + 4𝑒− 𝑆𝑛 + 6𝑂𝐻− E° = –1.02 V vs. Hg/HgO (3.5)

The redox potentials with stannate shifted, and after formation, the discharge profile had one main

plateau at –0.84 V that stabilized around 250 mAh g–1. This shift in redox potentials is attributable to

the formation of an alloy of iron and tin. The voltage cut-off during charge was set to –1.158 V to

avoid the excessive reduction of stannate ions in the electrolyte to metallic tin. Interestingly, the

HER overpotential shifted to more negative potentials, which allowed the Coulombic efficiency to be

determined. The efficiency obtained after formation was 85 ± 5%. If the two cells were compared

over the first discharge plateau, where commercial iron electrodes typically operate, the cell with

stannate (250 mAh g–1) outperformed the one without (150 mAh g–1).

Page 44: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

31

Figure 3.10. Charge and discharge profiles of iron electrodes cycled against nickel electrodes over 150 cycles at

192 mA g–1 in a) 6 M KOH + 1 M LiOH and b) + 0.1 M K2SnO3. Figure reproduced from Paper II111 with

permission from Journal of the Electrochemical Society.

Figure 3.11 shows SEM images of charged iron electrodes after cycling in alkaline electrolytes, as

well as of a pristine electrode. The pristine electrode in Figure 3.11a consisted of large graphite

flakes, together with smaller carbon black particles in the size range 40–60 nm, and iron powder

with an average particle size of 10 µm. After cycling the electrode, the particle size of iron dropped

to 0.2–1 µm, as shown in Figure 3.11b. This decrease in particle size occurred as iron dissolving and

was redeposited during cycling. In the electrode that was used in an electrolyte with stannate

added, octahedron-like particles 0.6–1 µm in size were deposited on the surface, as shown in Figure

3.11c. Spot elemental analysis with EDS showed an accumulation of tin in these particles. The

surface in Figure 3.11b consisted of 45wt% Fe and 4 wt% Bi, whereas that in Figure 3.11c was 22

wt% Fe, 2 wt% Bi, and 3 wt% Sn.

B

A

Page 45: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

32

Figure 3.11. SEM images of a) pristine iron electrode, b) charged iron electrode cycled in 6 M KOH + 1 M LiOH

and c) charged iron electrode cycled in 6 M KOH + 1 M LiOH + 0.1 M K2SnO3. Figure reproduced from Paper II111

with permission from Journal of the Electrochemical Society.

XPS analysis confirmed the presence of tin in the iron electrode cycled in the presence of stannate.

Figure 3.12 depicts the Sn 3d spectrum of the electrode to a depth of 5 nm. The fitted data matched

that of SnO2 formed from tin oxidized under exposure to air. Detection of tin here agrees with the

EDS results described earlier. Other elements detected on the surface included C, O, Bi, F and Fe.

However, the signal of Fe was weak and indicated only a minor amount on the surface. Paper II

includes the XPS spectra of the other elements.

Page 46: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

33

Figure 3.12. Deconvolution of the Sn 3d XPS spectrum measured to a depth of 5 nm depth on the iron

electrode that was cycled 150 times in 6 M KOH + 1 M LiOH + 0.1 M K2SnO3, at charge state. Figure reproduced

from Paper II111 with permission from Journal of the Electrochemical Society.

3.3.2. Phase evolution characterization and structure refinement

Operando neutron diffraction measurements used a cell design reported in other work for the

Polaris instrument at ISIS113,132. Prior to the measurement, the iron electrode was activated by

cycling in deuterated 6 M KOD + 1 M K2SnO3 electrolyte. After activation, the discharge capacity

reached 420 mAh g–1. Figure 3.13 depicts a representative neutron diffraction pattern and Rietveld

refinement of the activated iron electrode in the cell, which also contained nickel (current collector

and counter electrode). The background from the amorphous and liquid components was fitted by a

shifted Chebyschev function with 12 coefficients. Initially, the refinement included the following

phases: Ni(s) (PDF 00-004-0850), Fe(s) (PDF 00-006-0696), Fe3O4(s) (both nuclear and magnetic

structures, PDF 00-019-0629), graphite (“C”, PDF 00-056-0159), and Bi(s) (PDF 00-044-1246). Ni

dominated the pattern and Bi was found to have a very small contribution to the refinement and

was hence excluded.

A set of unidentified peaks whose intensity changed during cycling indicated a new phase taking part

in the electrochemical reaction. The peaks of this new phase did not shift in position and were

therefore not the result of any intercalation reaction, which would induce changes in the volume of

the unit cell. The peaks could be indexed to a tetragonal unit cell with the dimensions a = 3.4049(7)

and c = 3.255(2) Å. This unit cell was reminiscent of a high-pressure body-centered Sn modification

with a = 3.70(1) and c = 3.37(1) Å133. Fe, Bi and C have also been reported to form isostructural solid

solutions with Sn134–136. The observations that Sn metal was not detected in Figure 3.13, but that the

presence of Sn was confirmed ex situ by EDS and XPS, suggest the presence of a new solid solution

phase, composed of Sn in combination with other available elements in the electrode. Indeed, Fe, Bi

and C have been found to form related structures. A good fit to the neutron diffraction data was

accomplished using the space group P4/mmm with Fe, Sn, Bi and C located at (0,0,0) and (½,½,½)

and D atoms located in (½,½,0). Thus, this structure model contains an octahedral interstitial site in

the metal atom framework, similar to ones found for metal hydride structures. The presence of D

obstructs a determination of the element ratios in the solid solution, in particular considering the

correlation between site occupancies and thermal parameters, and the similar neutron scattering

cross-sections of the elements present. D and C have almost equal scattering lengths, but the 1.63 Å

distance between the interstitial sites and the metal atoms is in better agreement with D being the

Page 47: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

34

interstitial atom. Thus, the new phase is identified as an intermediate hydride phase (IHP). A good

agreement between observed and calculated data was achieved, shown in Figure 3.13, with the

refinement converging with a goodness of fit value χ2 = 2.56. The corresponding weight fractions of

phases were: Ni (59.2 wt% and RF = 1.1%), C (1.1 wt% and RF = 4.2%), (Fe,Sn,Bi,C)2D (15.3 wt% and RF

= 7.3%), Fe (14.4 wt% and RF = 4.7%), Fe3O4 (9.0 wt% and RF = 4.5%), and Bi (1.0 wt% and RF = 8.8%).

Figure 3.13. Fitted neutron diffraction pattern at a charge state of the iron electrode in 6 M KOD + 0.1 M

K2SnO3. All observed diffraction peaks could be assigned with their respective phase and diffraction indices.

Figure reproduced from Paper II111 with permission from Journal of the Electrochemical Society.

Figure 3.14a shows a series of neutron diffraction patterns collected over a full charge-discharge

cycle. Specific reflections of the Fe, Fe3O4, and IHP phases underwent the most noteworthy changes

during cycling. Figure 3.14b compares the relative weight fractions of these phases over time.

Refinement of each neutron diffraction pattern included all phases. Prior to the operando analysis,

the cell was fully charged in fresh 6 M KOD + 0.1 M K2SnO3 electrolyte, to ensure a full reduction of

the iron electrode. This explained the high initial weight fraction of IHP as Sn(OD)62– is expected to

undergo reduction upon D2 evolution. The discharge capacity achieved from this cell corresponded

to 380 mAh g–1, consistent with the performance of half-cells. The weight fraction trends of Fe and

Fe3O4 followed the expected phase evolution during charge and discharge. The amount of IHP,

however, correlated inversely with that of Fe. When Fe formed, IHP was depleted and vice versa.

Neither Fe(OH)2 nor Sn metal were detected operando. XRD analysis of the fully charged iron

electrode recovered after these measurements showed the presence of Sn (see supplementary

information of Paper II). This corroborated that Sn was part of the solid solution in the IHP and

reacted during cycling. The IHP phase was metastable and decomposed to Sn among other species.

Page 48: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

35

Figure 3.14. a) Operando neutron diffraction patterns depicting the phase evolution of Fe, Fe3O4 and the

intermediate hydride phase (IHP) in the iron electrode during charge and discharge cycling in 6 M KOD + 0.1 M

K2SnO3. Each neutron diffraction pattern was collected over 1 h. b) Weight fraction evolution of representative

phases (markers with error bars) and cell potential vs. O2/D2 (dashed green line) as a function of time. Figure

reproduced from Paper II111 with permission from Journal of the Electrochemical Society.

In summary, the evolution of hydrogen gas was minimized using potassium stannate as an additive.

Tin prolonged the discharge profile of iron electrodes and increased the overpotential required for

hydrogen evolution, both desirable attributes for batteries. Operando neutron diffraction

measurements showed a correlation of these attributes to the formation of a novel intermediate

hydride phase during charge-discharge cycling.

3.4. Hydrogen evolution on iron and zinc electrodes (Papers III

and IV)

The HER (1.4) contributes to deviation from 100% Coulombic efficiency in rechargeable alkaline iron

electrodes and zinc electrodes in aqueous sulfate electrolytes. Hydrogen evolution on both

b

a

Discharge Charge Rest

Page 49: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

36

electrodes was quantified using electrochemistry coupled with mass spectrometry. The design of a

special electrochemical cell connected to the MS allowed the detection of hydrogen gas

accumulated in the system. The measurements assumed that the HER was the sole contributor to

the deviation from 100% efficiency and were based on Faraday’s law:

𝑛 =𝐼𝑡

𝑧𝐹 (3.6)

Where n = number of moles (mol), I = current (A), t = time (s), z = electrons transferred per ion (z = 2

for H2) and F = 96485 C mol–1. Calibration used SS as a standard to quantify hydrogen gas. This

included linear sweep voltammetry over 0.3–0.55 V vs. Ag/AgCl at 0.2 mV s–1 upon accumulating

hydrogen gas. The charge passed (Coulombs) was correlated to the recorded counts of H2 on the MS.

From calibration, quantification factors were determined:

𝑞𝑢𝑎𝑛𝑡. 𝑓𝑎𝑐𝑡𝑜𝑟 =∫ 𝐼𝐸𝐶𝑡

∫ 𝐼𝑀𝑆𝑡 (3.7)

Where IEC = current recorded from the potentiostat (A) and IMS = extracted ion signal from the MS

(A). Calculated quantification factors corresponded to (12.1 ± 1) 106. The number of moles of H2

detected per second based on (3.6) and (3.7) was:

𝑛𝐻2

𝑡= 𝑞𝑢𝑎𝑛𝑡. 𝑓𝑎𝑐𝑡𝑜𝑟 ⋅

𝐼𝑀𝑆

𝑧𝐹 (3.8)

This was further derived to give the cumulative capacity of H2:

𝑄𝐻2= 𝑞𝑢𝑎𝑛𝑡. 𝑓𝑎𝑐𝑡𝑜𝑟 ⋅ 𝐼𝑀𝑆𝑡 (3.9)

Where QH2 = charge passed of H2 (As). The charge passed values of H2 were converted to mAh for

comparison with the charge capacity of the iron electrode. Finally, the charge efficiency was defined

as:

𝜂𝑐ℎ𝑎𝑟𝑔𝑒 =𝑄𝑡𝑜𝑡−𝑄𝐻2

𝑄𝑡𝑜𝑡 (3.10)

Where Qtot = total capacity obtained upon charging the iron electrode.

Figure 3.15 shows the evolution of charge efficiency for the four iron electrodes: a) CuSn, b)

CuSnCNT, c) CuSnCNTLi, and d) SnCNT. For each figure, the top plot shows the cell potential and

moles of H2 formed per second (3.8) as a function of time, while the bottom plot shows the charging

efficiency (3.10) and the cumulative H2 capacity (3.9) as a function of time. The extended H2

detection window from the MS accounted for all accumulated hydrogen gas. As such, the cumulative

H2 capacity and charge efficiency remained relatively constant close to the end of charge. The

highest efficiencies were obtained for the CuSnCNT sample, followed by the CuSn sample. Charging

did not improve with LiOH added to the electrolyte, as this decreased the efficiency. This can be

explained by the increased dissolution of iron, which facilitates electrolyte access to active sites and

stimulates hydrogen evolution137. The sample containing Sn as the sole dopant performed more

poorly because of its lower electrical conductivity and had a lower charge acceptance of iron. The

efficiency evolution showed three distinct regions during charge. In the first region, below 1.5 V,

nucleation of iron occurred, and the hydrogen gas evolution was lowest. In the second region,

between 1.5 and 1.6 V, iron reduction proceeded as hydrogen gas evolved at a constant rate. In the

Page 50: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

37

last region, above 1.6 V, the iron became fully reduced, and the hydrogen evolution rate increased

dramatically and overcharged the cell. From these results, the general trend showed that iron

electrodes achieved higher charge efficiencies at current densities above 10 mA cm–2 (C/5 or 192 mA

g–1 charge rates). The efficiency values were validated by comparing the calculated charge

efficiencies (3.10) acquired from the MS with Coulombic efficiencies from the potentiostat, and only

small deviations, ±3%, were observed.

Figure 3.15. Charge efficiency analysis at 5, 10 and 15 mA cm–2 of nanostructured copper and/or tin doped iron

electrode materials: a) CuSn, b) CuSnCNT, c) CuSnCNTLi and d) SnCNT. In each figure, the top plot shows the

cell potential (left y-axis) and number of H2 moles evolved per second (right y-axis) vs. time. The bottom plot

shows the charge efficiency (left y-axis) and cumulative H2 capacity (right y-axis) vs. time. The total charge

capacity of the iron electrode is annotated in the bottom plot in each sample. Figure reproduced from Paper

IV138 with permission from MDPI.

Figure 3.16 shows hydrogen gas detected upon plating and stripping zinc in aqueous sulfate

electrolytes. The setup used symmetric Zn vs. Zn cells coupled to the MS and operated at current

density rates of 1–100 mA cm–2. Results showed no noteworthy overpotential difference between

Page 51: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

38

the two cells, and neither evolved significant quantities of hydrogen gas; the detection of hydrogen

remained close to zero. This was consistent with the Coulombic efficiency values obtained from the

potentiostat, which were 99% and higher at all current densities in both cells. The zinc electrodes

cycled performed particularly efficient and had excellent electrochemical reversibility.

Figure 3.16. Zinc electrodeposition and dissolution in Zn vs. Zn cells in a) Mn2+ free and b) added Mn2+

electrolytes at 1, 10, 25 and 100 mA cm–2. The blue line corresponds to the voltage profile and red to the ion

current of H2 from the mass spectrometer. Figure reproduced from the supplementary information for Paper

III72 with permission from Elsevier.

Conclusively, hydrogen evolution on iron and zinc, evaluated as a function of charge current density,

was studied by coupled electrochemistry and mass spectrometry. From this, it was concluded that

the iron electrode benefited from high-rate operation, which lowered the rate of hydrogen

evolution, whereas hydrogen gas was not detected on zinc electrodes in aqueous sulfate

electrolytes.

3.5. Flow-assisted rechargeable Fe–air batteries (Paper V)

3.5.1. Structural characterization of oxygen electrocatalysts

New bifunctional catalysts for the air electrode were characterized structurally and

electrochemically. The structures of the synthesized oxygen electrocatalysts were determined prior

to the electrochemical performance analysis. Figure 3.17 shows the XRD patterns of the LCMO, Ni–

Fe LDH and NiCo2O4 catalysts. LCMO consisted of two phases with different La:Ca ratios: 1)

La0.1Ca0.9MnO3 and 2) La0.5Ca0.5MnO3. The peaks of the first phase can be indexed to an

orthorhombic unit cell with a = 5.3063(3) Å, b = 7.4923(5) Å and c = 5.2988(3) Å in the space group

Pnma139. The second phase, also orthorhombic, can be indexed with unit cell parameters a =

5.4164(5) Å, b = 5.4454(7) Å and c = 7.7099(8) Å in the space group Pbnm140. The structure of the Ni–

Fe LDH material is related to that of β-Ni(OH)2, but with larger interlayer spacing, and incorporated

Fe3+ replacing of Ni2+ 141. The pattern for of this material showed poor crystallinity and matched with

the reported LDH with an 8:2 ratio of Ni:Fe142. The NiCo2O4 XRD pattern is ascribed to a face-

centered cubic structure with the unit cell parameter a = 8.1164(5) Å, space group Fd-3m143.

Page 52: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

39

Figure 3.17. XRD patterns and assigned diffraction indices for synthesized oxygen electrocatalyst materials:

LCMO, Ni–Fe LDH and NiCo2O4. For the LCMO catalyst, the subscripts belong to the compounds: 1)

La0.1Ca0.9MnO3 and 2) La0.5Ca0.5MnO3. Figure reproduced from Paper V.

Figure 3.18 shows SEM micrographs of the a) LCMO, b) Ni–Fe LDH, and c) NiCo2O4 catalysts. The

synthesis of LCMO included a quenching step to break up agglomerates and produced fine

interconnected particles 100–300 nm in size. The Ni–Fe LDH material was composed of large

amorphous nanosheets, with a needle-like structure, and showed poor crystallinity as seen in the

XRD pattern in Figure 3.17. The single-phase NiCo2O4 catalyst consisted of varied crystallite

morphologies, with irregularly shaped spheres and rod-like structures. This material exhibited a

broad particle size distribution ranging from 50 to 500 nm.

003

006 012

015 110 113

111 220

311

222 400

422 511 440

1011

0021

2201 202

1 020

2 022

2 220

2

2042

Page 53: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

40

Figure 3.18. SEM images of synthesized oxygen electrocatalysts: A) LCMO, B) Ni–Fe LDH and C) NiCo2O4. Figure

reproduced from Paper V.

3.5.2. Electrochemical performance of air electrodes

The electrochemical performance of air electrodes was analyzed first by CV, then by charge-

discharge cycling until cell failure. The prepared air electrodes were highly hydrophobic and

therefore activated with CV before cycling. This established a three-phase boundary between

A

500 nm

B

C

500 nm

500 nm

Page 54: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

41

electrolyte, oxygen, and catalyst. In each measurement, nickel served as the counter electrode to

produce oxygen or hydrogen gas during operation, and an airflow supplied as oxygen to the

electrode. Three air electrode combinations were examined: 1) LCMO + Ni–Fe LDH, 2) LCMO +

NiCo2O4 and 3) NiCo2O4. The rationale for the two first air electrodes was to use combinations of

materials to merge active OER and ORR catalysts40,144. LCMO has excellent ORR performance but

lacks OER activity, hence it was mixed with superior OER catalysts such as Ni–Fe LDH and

NiCo2O497,105. On the contrary, Ni–Fe LDH catalysts perform poorly in reduction but are highly active

for oxidation40. NiCo2O4 has been investigated both as an OER enhancer but also as a bifunctional

catalyst39,145,146. Figure 3.19a shows the first CV cycle of the examined air electrodes during OER and

ORR. The maximum current densities of OER/ORR for the three samples were: 1) +4/–8, 2) +4/–11

and 3) +9/–16 mA cm–2. The air electrode with NiCo2O4 as single catalyst attained the highest current

densities in both directions and also had earlier onset potentials for OER (+5 mA cm–2 and +479 mV)

and ORR (–4.8 mA cm–2 and +80 mV).

Figure 3.19b shows the assessed cycle life of the air electrodes subjected to 2 h charge and discharge

steps until they failed. Mixtures of LCMO with Ni-Fe LDH or NiCo2O4, proved less activity and less

stable than the single NiCo2O4 catalyst. The first sample had significant overpotentials in ORR and

failed after 110 h of operation. This is ascribed to the limited ORR activity of the Ni–Fe LDH, which

caused rapid cell failure147. In the second sample, a large current spike in ORR occurred after 205 h

because of an unintended stop of airflow. The airflow was recovered shortly thereafter but the cell

deteriorated in subsequent cycles. Nevertheless, the air electrode containing only NiCo2O4 exhibited

lower total overpotentials of OER and ORR and sustained 440 h of operation. This electrode had a

total overpotential of 693 mV after 175 h operation at ±10 mA cm–2 (shown in Figure 1.5). Based on

these results, the NiCo2O4 air electrode was as the best candidate for an Fe–air prototype.

Page 55: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

42

Figure 3.19. A) Cyclic voltammetry analysis (1st cycle) at 5 mV s–1 for the three sample catalysts: 1) LCMO + Ni–

Fe LDH, 2) LCMO + NiCo2O4 and 3) NiCo2O4. The maximum current densities for the OER and ORR are

highlighted. B) Charge and discharge cycle life performance of the three samples at ±10 mA cm–2 with 2 h per

step. Figure reproduced from Paper V.

3.5.3. Electrochemical performance of the Fe–air prototype

Electrochemical cell testing of the Fe–air prototype operated with assisted flow of both electrolyte

and oxygen. The electrolyte flow dissipated heat and removed trapped bubbles, while oxygen gas

inhibited carbonate formation from carbon dioxide in air. The cell-adapted electrodes were aligned

in parallel with identical geometric surface areas of 49.2 cm2. The active mass of iron in the

electrode was 1.41 g while the air electrode used 1.04 g NiCo2O4. Charge and discharge were based

on the capacity of the iron electrode and the total cell had a theoretical capacity of 1355 mAhFe. For

example, ±10 mA cm–2 corresponded to a current rate of 349 mA g–1Fe or C/2.75 based on iron. The

reactions taking place in the cell during charge involved Fe(OH)2 reduction to iron and hydroxide ions

oxidized to oxygen and water (OER). Upon discharge, the opposite reactions followed. Running the

cell incorporated a preconditioning step of both electrodes at ±4 mA cm–2 until the discharge

capacity stabilized. Paper V includes the preconditioning data. During this step, the cell was

overcharged to the theoretical capacity of iron. After 300 h the discharge capacity stabilized and

output 450 mAh g–1Fe with an energy efficiency of 51%. The low energy efficiency achieved is

attributed to the sluggish OER and ORR kinetics.

After preconditioning, the Fe–air rate capability testing, shown in Figure 3.20, used charge and

discharge rates between ±5 to ±25 mA cm–2. In a), the discharge profiles show that specific capacities

-16 mA cm-2

-11 mA cm

-2

-8 mA cm-2

+9 mA cm-2

+4 mA cm-2

B

A

Page 56: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

43

of 510 to 333 mAh g–1 were obtained, depending on the varying rate used. Discharge potentials

decreased with increased current densities, induced by a significant increase of ohmic and

polarization losses. The limited kinetics of the cell reaction resulted in impractical cell potentials at

current densities above ±15 mA cm–2 (below 0.6 V). Figure 3.20b summarizes the discharge

capacities and calculates the energy densities. The highest energy density of 377 Wh kg–1Fe, achieved

at 10 mA cm–2, represented a deviation from the expected trend of decreasing with increasing

current densities. The charge efficiency evolution of iron electrodes, seen in Figure 3.15, showed the

highest charge acceptance at this point (10 mA cm–2) and might explain the trend here. When

accounting for the total mass of the electrodes (4.22 g for the iron electrode and 3.78 g for the air

electrode), the energy density corresponded to 90 Wh kg–1electrodes. Figure 3.20c depicts the power

density curves based on the geometric surface area and mass of iron. The highest power output

reached 11.3 mW cm–2 / 555 mW g–1Fe.

Figure 3.20. A) Discharge performance of the Fe–air prototype at 5–25 mA cm–2. B) Discharge capacities and

calculated energy densities based on the nominal voltages of respective current densities. C) Power density

curve of the prototype cell. Figure reproduced from Paper V.

C

B

A

Page 57: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

44

The investigated durability of the Fe–air cell upon cycling at ±10 mA cm–2, limited the charge capacity

to 440 mAh g–1Fe and set the discharge cutoff at 0.58 V. Figure 3.21a shows the charge and discharge

curves over 76 cycles with an average discharge potential of 0.73 V, discharge capacity of 355 mAh

g–1Fe, and energy efficiency of 47%. The charge and discharge characteristics of the cell followed the

electron-exchange character of the iron electrode, with the OER/ORR plateaus remaining flat as

shown in Figure 1.5. The cell cycled over 200 h and exhibited stable performance until failure. In the

last charge cycles, another plateau evolved at 1.8 V, corresponding to the HER. The increased

hydrogen evolution rate indicated poor charge acceptance of iron caused by passivation. Figure

3.21b summarizes the measured nominal cell discharge voltages, specific discharge capacities, and

calculated energy densities. The cell averaged an energy density of 259 Wh kg–1Fe and Coulombic

efficiency of 79%. This compares to other reported work on Fe–air cells, which have achieved 290

Wh kg–1Fe

85. Figure 3.21c shows the ohmic resistance (IR) over all cycles, with stable values around

0.37 Ω. This corresponds to a resistivity of 0.68 Ω m from a geometric surface area of 49.2 cm2 and

interlayer spacing between the electrodes of 0.27 cm.

Figure 3.21. A) Charge and discharge profiles of Fe–air prototype cell over 76 cycles at ±10 mA cm–2 and with

oxygen flow. B) Measured and calculated cell discharge potentials, specific discharge capacities, and energy

densities over 76 cycles. C) Measured Coulombic efficiency and IR drop over 76 cycles. Figure reproduced from

Paper V.

C

B

A

Page 58: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

45

In summary, new bifunctional catalysts for the air electrode were characterized structurally and

electrochemically. The findings presented an excellent catalyst candidate in the non-precious-metal

phase NiCo2O4. This catalyst showed superb activity for oxygen evolution and reduction with

excellent long-term stability. The catalyst was used in a rechargeable alkaline Fe–air battery as a

demonstration of this technology, and its effect on performance and cycle life was presented. The

results proved a robust bifunctional performance of the Fe–air cell and demonstrated the obstacles

and opportunities of Fe–air batteries. The main limitations of poor rate capability and low energy

and Coulombic efficiencies are hard to overcome. The outstanding advantage of this battery is the

low cost of the materials it uses. Further development of this prototype should use lightweight

materials, shorten the interlayering spacing between the electrodes, and eliminate the pump to flow

electrolyte. Upon these improvements, the cell needs to show minimal unwanted side reactions

such as the HER, which hurts the cell efficiency.

Page 59: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

46

4. Conclusions

This thesis undertook to outline the limitations of rechargeable aqueous Zn–MnO2 and Fe–air

batteries and address these issues using new and modified electrode and electrolyte materials. The

impact of LiOH addition to alkaline electrolytes for Zn–MnO2 cells depend on the concentration of

the KOH:LiOH mixtures due to the competing reaction mechanisms of proton and lithium insertion.

When lithium insertion or lithium–proton insertion dominated, the reaction mechanism inhibited

zinc poisoning i.e. by forming ZnMn2O4. This allowed the cell to cycle up to 60 cycles with minimal

capacity fade, and average around 250 mAh g–1. The optimal electrolyte composition consisted of 1

M KOH and 3 M LiOH and enabled both lithium and proton insertion. This electrolyte composition,

demonstrated to be optimal using operando EDXRD and ex situ XRD, suppressed the intermediate

Mn3+ phases from proton insertion in favor of lithiated Mn spinel phases, and upheld reversibility.

Aqueous sulfate Zn–MnO2 cells offered better reversibility and performance than their alkaline

equivalents. Cells with Mn2+ added to the zinc sulfate electrolyte cycled over 100 times and had

stable capacities of 220 mAh g–1 at 1.35 V. The findings showed that accessible Mn2+ in the

electrolyte limited Mn dissolution. This limited Mn vacancies for Zn2+ insertion, as demonstrated by

EDS, which showed a lower Zn/Mn ratio than the cell cycled without Mn2+ in the electrolyte.

Furthermore, operando XRD showed smaller unit cell volume expansion upon zinc insertion when

Mn2+ was available. These observations correlated to the enhanced cell rechargeability. Operando

phase evolution analysis revealed that layered Zn-buserite was formed during the first recharge. This

explained the transition in electrochemical redox reaction characteristics from the first cycle to

subsequent cycles. Zinc sulfate hydroxide precipitated during discharge due to increased pH upon

proton insertion but did not involve the exchange of electrons. This precipitate showed significant

coverage on the electrode surface after 100 cycles but still managed to dissolve easily during charge.

Rechargeable iron electrodes with stannate added to the electrolyte presented stable performance

over 150 cycles with 85 ± 5% in Coulombic efficiencies. The additive prolonged the first reduction

plateau and increased this capacity by 100 mAh g–1. Post-mortem SEM and XPS analysis confirmed

that tin was deposited on the surface of the iron electrode, while operando neutron diffraction

analysis did not. Instead, operando phase analysis unveiled a suggested intermediate hydride phase

composed of a solid solution of the participating elements. The joint effect of iron and the new

phase both increased the hydrogen evolution overpotential and prolonged the discharge state for

practical batteries.

Quantification of hydrogen evolution on rechargeable iron electrodes showed three distinct rate

domains for hydrogen evolution during charge. Low hydrogen rates developed during charge until

the cell overcharged, which significantly increased the hydrogen evolution rate. The higher rates of

hydrogen evolution deteriorated the electrodes and capacity declined over cycles. The highest

charge acceptance of iron occurred at charge rates higher than 192 mA g–1. This finding should

promote suitable cycling protocols for commercial iron electrodes. On zinc electrodes in aqueous

sulfate electrolytes, hydrogen gas was not detected, Coulombic efficiencies above 99% were

achieved for rates of 1–100 mA cm–2.

The optimized charge protocol for iron electrodes was used to demonstrate a proof-of-concept flow-

assisted Fe–air cell. The cell performed with long cycle life and output an energy density of 377 Wh

kg–1Fe. The cell operated over 588 h with oxygen flow while preventing the HER and Fe3O4 passivation

on the iron electrode. The findings presented NiCo2O4 as a highly active bifunctional catalyst with

total overpotential (OER/ORR) of 693 mV at ±10 mA cm–2. This performance is encouraging, and if

Page 60: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

47

the weight of the cell can be lowered in future, it may increase the energy density while refining the

cumbersome limitations of low efficiencies. This system shows potential for large-scale energy

storage solutions because of the materials used.

These results are a stepping-stone toward the use of rechargeable Zn–MnO2 and Fe–air batteries for

large-scale energy storage as they motivate low-cost and safe operation. These batteries show

compelling potential for sustainable solutions in the electric grid.

Page 61: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

48

5. Future perspectives

Rising concerns on the dependence of Li-ion batteries limited cobalt supply encourage research on

alternative batteries using available materials. Current studies focused on improving the

performance and cycle life of rechargeable aqueous Zn–MnO2 and Fe–air batteries, both of which

are environmentally friendly and not resource restricted.

Improvements in the cyclability of alkaline Zn–MnO2 batteries were made by promoting the lithium

insertion mechanism on the manganese dioxide electrode. This formed lithiated Mn spinels instead

of the known irreversible phase, ZnMn2O4. Investigations to further prolong the cycle life of these

batteries are necessary. One possibility is to limit the DOD of the Zn–MnO2 battery. Limiting the DOD

to the one-electron exchange of manganese dioxide (Mn4+/Mn3+) may inhibit further reduction to

the inactive spinels. In addition to validating the electrochemical reversibility, this would need

accurate structural analysis to confirm that the desired redox reaction of the manganese dioxide

electrode occurs. The downside with limiting the DOD is clearly that the battery would not fully

utilize the capacity potential of the manganese dioxide electrodes. Thus, replacing alkaline

electrolytes with aqueous sulfate ones is exceptionally promising for reversible and high-

performance Zn–MnO2 batteries. These electrolytes, in which the redox reactions occur at lower pH,

overcome the main obstacles of alkaline ones. However, actualizing aqueous sulfate Zn–MnO2

batteries for commercial use will require a transitioning to practical battery conditions, in order to

scale up. This entails not using flooded cells, using low active mass loadings of the manganese

dioxide and carbon paper as current collector. Focus on the manganese dioxide electrode, which is

significantly less energy-dense than the zinc electrode, is essential to reach high energy densities in

this battery. Moreover, morphological investigations of zinc plating are recommended to minimize

dendrite formation.

Overcoming the hurdles associated with Fe–air batteries is challenging considering the cumbersome

nature of the redox reactions involved. The HER on the iron electrode and the large overpotentials

of the air electrode impair the cell efficiency. Further development to suppress the HER on the iron

electrode is required to reach higher Coulombic efficiencies. Reported efficiencies around 90% are

not enough for practical batteries; research must aim for efficiencies above 99%. This may be

possible with more efficient hydrogen-evolution suppressors than tin or sulfide. Another possibility

is to deposit a few monolayers of tin or sulfide compounds on each individual iron particle to

efficiently increase the hydrogen overpotential. Advanced engineering of active bifunctional

catalysts that lower the OER/ORR overpotentials of the air electrode is vital to enhance efficient

operation of the Fe–air battery. An interesting approach is to directly synthesize active catalysts on

conductive metal backbones without carbon additives. This would minimize degradation of the air

electrode performance due to carbon corrosion. Further opportunities should incorporate

lightweight cell materials, optimize the cell design, and eliminate the use of a pump for electrolyte

flow, in order to enhance the energy density of the cell.

Conclusively, the suggested studies would be of interest for rechargeable Fe–air and Zn–MnO2

batteries.

Page 62: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

49

6. Sammanfattning

Globala energisystem övergår alltmer ifrån fossila bränslen till förnybara energikällor för att

upprätthålla en hållbar lösning för människans framtida energibehov. Incitamentet är att ersätta

stora traditionella kraftverk med mindre förnybara energikällor. Paradigmskiftet till förnybar energi

behöver emellertid hantera den ökande globala energikonsumtionen som uppskattas att fördubblas

år 2050. Att fullända övergången till förnybara energikällor som vindkraft, solkraft, vattenkraft och

geotermisk energi är en komplex uppgift. Deras fluktuerande karaktär är beroende av årstid och tid

på dygnet, som i sig resulterar i oförutsägbara energigenerationsprofiler. Därför behövs effektiva

energilagringssystem för att hantera dessa fluktuationer och säkerställa tillförlitlig energiförsörjning

vid behov. Inom energilagringteknologier är batterier den snabbast växande teknologin på grund av

dess mångsidighet av att möta uppsatta energi- och effektkrav samt dess enkla förmåga att skala

upp energilagringskapacitet från kWh till MWh. Dessutom är batterier inte beroende av tid, väder

eller geografiska förutsättningar vilket möjliggör snabb installation och expansion. För storskalig

implementering av batterier krävs energieffektiva teknologier som är miljövänliga, baserade på

lättillgängliga råvarumaterial, har låga installationskostnader samt lång livslängd. Dagens

dominerade batteriteknologier, exempelvis litiumjon och blysyra, uppfyller inte alla dessa krav och

därför motiveras forskning av alternativa batterilösningar.

Batterier består av elektrokemiska celler som innehåller två elektroder där de

elektrokemiskareaktionerna sker. En elektrolyt separerar elektroderna och innehåller joner som

vandrar fritt från en elektrod till den andra. Elektroderna ansluts externt för att påbörja sina

reaktioner medan elektroner överförs i en yttre krets som genererar ström. Dessa elektrokemiska

celler kan sedan kopplas i serie eller parallellt för att uppnå den önskade spänningen respektive

energikapaciteten. Viktiga egenskaper hos batterier är energitäthet (Wh L–1), specifik energi (Wh kg–

1) och specifik effekt (W kg–1), dvs mängden energi de kan lagra per massa eller volym samt

tidslängden av energitillförseln. I storskalig energilagring är de primära faktorerna inte energitäthet

eller effekt, utan istället låga installationskostnader, lång livslängd, hög energieffektivitet och

enkelheten med att skala upp lagringskapaciteten.

Batterier som använder vattenbaserade elektrolyter möjliggör låga tillverkningskostnader och är i sig

säkrare än litiumjonbatterier. Zink-, mangandioxid-, järn- och luftbaseradebatterier har hög

energirelevans, de är inte resursbegränsade och kan bidra till storskaliga energilagringslösningar.

Zink har länge använts som elektrodmaterial för primära alkaliska zink-mangandioxidbatterier.

Historiskt har zinkelektroder begränsats för uppladdningsbara batterier på grund av morfologiska

osäkerheter och passiveringseffekter. Mangandioxidelektroder är ineffektiva som uppladdningsbara

elektroder på grund av felmekanismer associerade med fastransformationer under battericykling.

Irreversibiliteten av mangandioxid är starkt korrelerad med bildningen av de elektrokemiskt inaktiva

spinellfaserna, Mn3O4/ZnMn2O4. Utvecklingen av järnelektroden för järnluftbatterier initierades

redan i slutet på 1960-talet men begränsas än idag av undermålig laddningseffektivitet. Dessutom är

luftelektrodens livslängd begränsad vid långvarig drift på grund av den långsamma syrgasutvecklings-

och reduktionskinetiken. Dessa begränsningar av järnluftbatterier medför stora energiförluster vid

battericykling.

I detta avhandlingsarbete motverkades begränsningarna för uppladdningsbara zink-mangandioxid-

och järnluftbatterier genom att syntetisera elektrodmaterial och modifiera elektrolytkompositioner.

Olika katjonelektrolytblandningar av kaliumhydroxid och litiumhydroxid för uppladdningsbara

alkaliska zink-mangandioxidbatterier medförde konkurrerande proton- och

Page 63: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

50

litiuminsättningsmekanismer. Med den optimala elektrolytkompositionen av 1 molar kaliumhydroxid

och 3 molar litiumhydroxid, visade elektromikroskopologiska och röntgenkristallografiska resultaten

att den inaktiva fasen ZnMn2O4 undertrycktes till förmån för litiuminnehållande

manganspinellfaser, som i sig förbättrade livslängden av uppladdningsbara mangandioxidelektroder.

I det analoga batteriet med samma elektrodmaterial men istället med ett milt surt zinksulfat

elektrolytlösning undersöktes. Med hjälp av operando röntgendiffraktionsmätningar klargjordes

både den komplexa reaktionsmekanismen och varför tillsättningen av Mn2+ joner i elektrolyten

förbättrade uppladdningsförmågan. Den föreslagna reaktionsmekanismmodellen föreslog insättning

av både protoner och zinkjoner i mangandioxiden. Vid urladdning detekterades en

zinksulfathydroxid fällning, utlöst av en lokal pH-ökning. Denna fällning involverade inga

elektronöverföringar och visade sig ha en signifikant påverkan på den elektrokemiska prestandan.

Vidare visade studien att tillgängliga Mn2+ joner vid elektrod-elektrolytskiktet begränsade

mangandioxidupplösningen. Detta i sig begränsande zinkinsättning och förbättrade

uppladdningsbarheten av mangandioxidelektroden.

De undersökta hindren för uppladdningsbara järnluftbatterier motverkades genom att undertrycka

vätegasutveckling på järnelektroden och optimering av bifunktionella syrgaskatalysatorer för

luftelektroden. Minimering av vätgasutveckling utnyttjade kaliumstannat som tillsatsmedel.

Bakgrunden till tillsatsen var att tennmetall elektropläterades under uppladdning av järnelektroden

och agerade som en vätgasutvecklingsbarriär. Tillsatsen av kaliumstannat förlängde

urladdningsprofilen av järnelektroderna och ökade överpotentialen av vätgasutvecklingen, båda

önskvärda attribut för batterier. Operando neutrondiffraktionsmätningar visade en korrelation av

dessa attribut till en bildad intermediär fas vid uppladdning. Vidare utvärderades vätgasutvecklingen

av nanostrukturerade järnmaterial dopade med koppar och tenn, genom kopplad elektrokemi och

masspektrometri. Studien visade att järnelektroden gynnades av höga uppladdningsströmmar

eftersom vätgasutvecklingen reducerades. För luftelektroden syntetiserades bifunktionella

katalysatorer som karakteriserades strukturellt och elektrokemiskt. Resultaten visade en utmärkt

katalysatorkandidat i NiCo2O4 med hög katalysatoraktivet för både syrgasutveckling och

syrgasreduktion. Katalysatorn anpassades till ett uppladdningsbart alkaliskt järnluftbatteri koncept.

Konceptet omfattade assisterande flöden av både elektrolyt och syrgas. Batteriets prestanda och

livslängd påvisades samt diskuterades möjligheterna och begränsningarna av järnluftsbatterier.

Mikro-, makro-, nano- och atomskala studier av elektrodmaterialen genomfördes för att öka

förståelsen för dessa materials växelverkan och natur. Detta inkluderade både operando och ex situ

karaktärisering. Röntgen- och neutronstrålning, analytiska och elektrokemiska metoder gav insikt på

förbättringar av batteriernas prestanda och livslängd. Resultaten rapporterade i detta

avhandlingsarbete är en språngbräda för att realisera uppladdningsbara zink-mangandioxid- och

järnluftbatterier för storskalig energilagring. Dessa batterier motiverar låg kostnad och säker drift

samt visar potential för hållbara lösningar i framtidens energisystem.

Page 64: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

51

Acknowledgements

Foremost, I would like to express my sincere gratitude to my supervisor, Prof. Dag Noréus, for

supporting me throughout my PhD studies, guidance, and invaluable knowledge. I highly value your

trust, our discussions, and your encouragement to strive for perfection.

I want to thank my co-supervisors, Prof. Gunnar Svensson and Dr. Jekabs Grins, for taking interest in

my work, for fruitful discussions, and supporting me at times when I needed it the most.

I want to thank my former PhD fellow, Dr. Yang Shen, for the support and amazing times shared

together. Further thanks to Erik Grape, Mirva Eriksson, Mohammed Naeem Iqbal, and Ghislaine

Robert-Nicoud for the enjoyable atmosphere in the office.

Special thanks to my collaborators in the FAIR project, Alagar Raj Paulraj (KTH), Prof. Yohannes Kiros

(KTH), Dr. Björn Skårman (Höganäs AB) and Dr. Hilmar Vidarsson (Höganäs AB), as well as Dr. Gunder

Karlsson (SiteTel Sweden AB). Thank you for excellent collaborations and for the achievements we

accomplished together. Thanks to the Swedish Energy Agency for supporting this work.

Thanks to my collaborators, Dr. Cheuk-Wai Tai, Dr. Henrik Svengren, Prof. Mats Johnsson, Dr.

William Brant (Uppsala University), Dr. Darius Milcius (LEI) and Dr. Martynas Lelis (LEI). Your

engagement and contributions are much appreciated.

I would also like to thank my former supervisor, Prof. Dan Steingart (Princeton University), former

colleagues, Dr. Benjamin Hertzberg (Princeton University), Dr. Andrew Hsieh (Princeton University),

Dr. Greg Davies (Princeton University), Dr. Can Erdonmez (BNL), and collaborators, Dr. An Huang

(UCSD), Dr. Joon Kyo Seo (UCSD), Dr. Zhong Zhong (BNL), Prof. Mark Croft (Rutgers University), Prof.

Ying Shirley Meng (UCSD).

My appreciation goes to Jordi Jacas Biendicho for preparing the neutron beam time proposal

provided by the UK Science and Technology Facilities Council, as well as Dr. Ron Smith and Dr.

Stephen Hull for your technical support and measurements at ISIS, UK.

Many thanks to Dr. Lars Eriksson, Dr. Kjell Jansson and Dr. Zoltán Bascik for user training and

instructions on various instruments.

I want to thank Dr. Tamara Church and Prof. Xiaodong Zou for proofreading my thesis and for your

helpful feedback.

I offer my appreciation to the entire administrative, technical, safety, and workshop staff at MMK:

Tatiana Bulavina, Helmi Frejman, Ann Loftsjö, Daniel Emanuelsson, Camilla Berg, Elisabeth Pernbom,

Rolf Eriksson, Anne Ertan, Kadir Abdul Karim, Hans-Erik Ekström, Per Jansson and Jakob Paulin.

To all my friends and colleagues at MMK (past and present), I thank you for making my time as a PhD

student enjoyable: Andrew Kentaro Inge, Stef Smeets, Yunchen Wang, Laura Samperisi, Ning Yuan,

Viktor Bengtsson, Alexandra Neagu, Dickson Ojwang, Bin Wang, Jingjing Zhao, German Salazar

Page 65: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

52

Alvarez, Thomas Thersleff, Hani Nasser Abdelhamid, Fei Peng, Eleni Mitoudi-Vagourdi, Ahmed

Etman, Magdalena Cichocka, István-Zoltan Jenei , Tom Willhammar, Hongyi Xu, Elina Kapaca, Jonas

Ångström, Dariusz Wardecki, Paulo Barros, Fredrik Björnerbäck, Aditya Dharanipragada, Daniel Eklöf,

Chris Celania, Alisa Gordeeva, Korneliya Gordeyeva, Yulia Trushkina, Valentina Guccini, Zhehao

Huang, Martin Kupuscinski, Konstantin Kriechbaum, Molly Lightowler, Yi Luo, Taimin Yang, Pierre

Munier, Przemyslaw Rzepka, Jon Olsén, Hugo Voisin, Varvara Apostolopoulou-Kalkavoura, Sara

Skoglund, Xia Wang, Anne Willert, Shun Yu and Yuan Zhong.

Last but not least, I am grateful to my wonderful family; my parents (Marin and Nabil), brothers

(Ninus and Tony) and my life partner (Eloisa) for their constant support, love, and care. They have

always been by my side throughout this journey, both in happy and difficult moments. I am also

grateful to my friends and relatives for the encouragement that helped me reach this moment.

Page 66: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

53

References

1. Larcher, D. & Tarascon, J.-M. Towards greener and more sustainable batteries for electrical

energy storage. Nat. Chem. 7, 19–29 (2015).

2. Castillo, A. & Gayme, D. F. Grid-scale energy storage applications in renewable energy

integration: A survey. Energy Convers. Manage. 87, 885–894 (2014).

3. Soloveichik, G. L. Battery technologies for large-scale stationary energy storage. Annu. Rev.

Chem. Biomol. Eng. 2, 503–527 (2011).

4. Ibrahim, H. & Ilinc, A. Techno-Economic Analysis of Different Energy Storage Technologies.

Energy Storage - Technologies and Applications (ed. Zobaa, A.) (InTech, 2013).

5. Dunn, B., Kamath, H. & Tarascon, J.-M. Electrical energy storage for the grid: a battery of

choices. Science 334, 928–935 (2011).

6. Energilagring - Teknik för lagring av el. (Kungliga Ingenjörsvetenskapsakademien (IVA), 2015).

7. Goodenough, J. B. Electrochemical energy storage in a sustainable modern society. Energy

Environ. Sci. 7, 14–18 (2014).

8. Tarascon, J. M. & Armand, M. Issues and challenges facing rechargeable lithium batteries.

Nature 414, 359–367 (2001).

9. Ragone, D. V. Review of battery systems for electrically powered vehicles. (SAE Technical Paper,

1968).

10. 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.

Renewable Sustainable Energy Rev. 68, 1174–1182 (2017).

11. Goodenough, J. B. & Kim, Y. Challenges for Rechargeable Li Batteries. Chem. Mater. 22, 587–

603 (2010).

12. Wanger, T. C. The Lithium future—resources, recycling, and the environment. Conservation

Letters 4, 202–206 (2011).

13. Roth, E. P. & Orendorff, C. J. How electrolytes influence battery safety. Electrochem. Soc.

Page 67: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

54

Interface, 45–49 (2012).

14. Yabuuchi, N. & Ohzuku, T. Novel lithium insertion material of LiCo1/3Ni1/3Mn1/3O2 for advanced

lithium-ion batteries. J. Power Sources 119-121, 171–174 (2003).

15. European Commission. 2017 list of Critical Raw Materials for the EU. (2017).

16. Massé, R. C., Uchaker, E. & Cao, G. Beyond Li-ion: electrode materials for sodium- and

magnesium-ion batteries. Science China Materials 58, 715–766 (2015).

17. Thackeray, M. M., Wolverton, C. & Isaacs, E. D. Electrical energy storage for transportation—

approaching the limits of, and going beyond, lithium-ion batteries. Energy Environ. Sci. 5, 7854–

7863 (2012).

18. Goodenough, J. B. & Park, K.-S. The Li-ion rechargeable battery: a perspective. J. Am. Chem. Soc.

135, 1167–1176 (2013).

19. Luntz, A. Beyond Lithium Ion Batteries. J. Phys. Chem. Lett. 6, 300–301 (2015).

20. Kundu, D., Vajargah, S. H., Wan, L., Adams, B., Prendergast, D. & Nazar, L. F. Aqueous vs.

nonaqueous Zn-ion batteries: consequences of the desolvation penalty at the interface. Energy

Environ. Sci 11. (2018).

21. White, C. D. & Zhang, K. M. Using vehicle-to-grid technology for frequency regulation and peak-

load reduction. J. Power Sources 196, 3972–3980 (2011).

22. Zhang, N., Cheng, F., Liu, J., Wang, L., Long, X., Liu, X., Li, F. & Chen, J. Rechargeable aqueous

zinc-manganese dioxide batteries with high energy and power densities. Nat. Commun. 8, 405

(2017).

23. Ingale, N. D., Gallaway, J. W., Nyce, M., Couzis, A. & Banerjee, S. Rechargeability and economic

aspects of alkaline zinc–manganese dioxide cells for electrical storage and load leveling. J.

Power Sources 276, 7–18 (2015).

24. McKerracher, R. D., Ponce de Leon, C., Wills, R. G. A., Shah, A. A. & Walsh, F. C. A Review of the

Iron–Air Secondary Battery for Energy Storage. ChemPlusChem 80, 1-14 (2014).

25. Reich, M. & Vasconcelos, P. M. Geological and Economic Significance of Supergene Metal

Page 68: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

55

Deposits. Elements 11, 305–310 (2015).

26. McBreen, J. Zinc Electrode Shape Change in Secondary Cells. J. Electrochem. Soc. 119, 1620–

1628 (1972).

27. Falk, S. U. & Salkind, A. J. Alkaline storage batteries. (John Wiley & Sons, 1969).

28. Chakkaravarthy, C., Waheed, A. K. A. & Udupa, H. V. K. Zinc—air alkaline batteries — A review.

J. Power Sources 6, 203–228 (1981).

29. Shoji, T., Hishinuma, M. & Yamamoto, T. Zinc-manganese dioxide galvanic cell using zinc

sulphate as electrolyte. Rechargeability of the cell. J. Appl. Electrochem. 18, 521–526 (1988).

30. Kalu, E. E. & White, R. E. Zn/Br2 cell: Effects of plated zinc and complexing organic phase. AIChE

J. 37, 1164–1174 (1991).

31. Tomazic, G. Zinc-bromine battery with circulating electrolytes. J. Power Sources 70, 168–168

(1998).

32. Lim, H. S., Lackner, A. M. & Knechtli, R. C. Zinc‐Bromine Secondary Battery. J. Electrochem. Soc.

124, 1154–1157 (1977).

33. Zhang, X. G. Corrosion and electrochemistry of zinc. (Springer Science & Business Media, 2013).

34. McLarnon, F. R. & Cairns, E. J. The Secondary Alkaline Zinc Electrode. J. Electrochem. Soc. 138,

645–656 (1991).

35. Reddy, T. Linden’s Handbook of Batteries, 4th Edition. (McGraw Hill Professional, 2010).

36. Parker, J. F., Chervin, C. N., Pala, I. R., Machler, M., Burz, M. F., Long, J. W. & Rolison, D. R.

Rechargeable nickel–3D zinc batteries: An energy-dense, safer alternative to lithium-ion.

Science 356, 415–418 (2017).

37. Parker, J. F., Pala, I. R., Chervin, C. N., Long, J. W. & Rolison, D. R. Minimizing Shape Change at

Zn Sponge Anodes in Rechargeable Ni–Zn Cells: Impact of Electrolyte Formulation. J.

Electrochem. Soc. 163, A351–A355 (2016).

38. Mainar, A. R., Iruin, E., Colmenares, L. C., Kvasha, A., de Meatza, I., Bengoechea, M., Leonet, O.,

Boyano, I., Zhang, Z. & Blazquez, J. An overview of progress in electrolytes for secondary zinc-air

Page 69: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

56

batteries and other storage systems based on zinc. Journal of Energy Storage 15, 304–328

(2018).

39. Pichler, B., Weinberger, S., Reščec, L., Grimmer, I., Gebetsroither, F., Bitschnau, B. & Hacker, V.

Bifunctional electrode performance for zinc-air flow cells with pulse charging. Electrochim. Acta

251, 488–497 (2017).

40. Li, Y., Gong, M., Liang, Y., Feng, J., Kim, J.-E., Wang, H., Hong, G., Zhang, B. & Dai, H. Advanced

zinc-air batteries based on high-performance hybrid electrocatalysts. Nat. Commun. 4, 1805

(2013).

41. Mehta, S. Investigation of capacity fade in flat-plate rechargeable alkaline MnO₂/Zn cells.

(University of British Columbia, 2016).

42. Tang, L., Wang, J.-W., Meng, L.-R. & Jin, C.-C. Influence of the EMD on the high-rate alkaline Zn-

MnO2 battery. Chinese Battery Industry 4, 006 (2012).

43. Fang, G., Zhou, J., Pan, A. & Liang, S. Recent Advances in Aqueous Zinc-Ion Batteries. ACS Energy

Lett. 3, 2480–2501 (2018).

44. Xu, C., Li, B., Du, H. & Kang, F. Energetic zinc ion chemistry: the rechargeable zinc ion battery.

Angew. Chem. Int. Ed Engl. 51, 933–935 (2012).

45. Zhang, N., Cheng, F., Liu, Y., Zhao, Q., Lei, K., Chen, C., Liu, X. & Chen, J. Cation-Deficient Spinel

ZnMn2O4 Cathode in Zn(CF3SO3)2 Electrolyte for Rechargeable Aqueous Zn-Ion Battery. J. Am.

Chem. Soc. 138, 12894–12901 (2016).

46. Gupta, T., Kim, A., Phadke, S., Biswas, S., Luong, T., Hertzberg, B. J., Chamoun, M., Evans-

Lutterodt, K. & Steingart, D. A. Improving the cycle life of a high-rate, high-potential aqueous

dual-ion battery using hyper-dendritic zinc and copper hexacyanoferrate. J. Power Sources 305,

22–29 (2016).

47. Yang, H. S., Park, J. H., Ra, H. W., Jin, C.-S. & Yang, J. H. Critical rate of electrolyte circulation for

preventing zinc dendrite formation in a zinc–bromine redox flow battery. J. Power Sources 325,

446–452 (2016).

Page 70: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

57

48. Wu, M. C., Zhao, T. S., Jiang, H. R., Zeng, Y. K. & Ren, Y. X. High-performance zinc bromine flow

battery via improved design of electrolyte and electrode. J. Power Sources 355, 62–68 (2017).

49. Li, B., Nie, Z., Vijayakumar, M., Li, G., Liu, J., Sprenkle, V. & Wang, W. Ambipolar zinc-polyiodide

electrolyte for a high-energy density aqueous redox flow battery. Nat. Commun. 6, 6303 (2015).

50. Xie, C., Duan, Y., Xu, W., Zhang, H. & Li, X. A Low-Cost Neutral Zinc-Iron Flow Battery with High

Energy Density for Stationary Energy Storage. Angew. Chem. Int. Ed Engl. 56, 14953–14957

(2017).

51. Gong, K., Ma, X., Conforti, K. M., Kuttler, K. J., Grunewald, J. B., Yeager, K. L., Bazant, M. Z., Gu,

S. & Yan, Y. A zinc–iron redox-flow battery under $100 per kW h of system capital cost. Energy

Environ. Sci. 8, 2941–2945 (2015).

52. Davies, G., Hsieh, A. G., Hultmark, M., Mueller, M. E. & Steingart, D. A. Utilization of Hyper-

Dendritic Zinc during High Rate Discharge in Alkaline Electrolytes. J. Electrochem. Soc. 163,

A1340–A1347 (2016).

53. Bass, K., Mitchell, P. J., Wilcox, G. D. & Smith, J. Methods for the reduction of shape change and

dendritic growth in zinc-based secondary cells. J. Power Sources 35, 333–351 (1991).

54. Parker, J. F., Nelson, E. S., Wattendorf, M. D., Chervin, C. N., Long, J. W. & Rolison, D. R.

Retaining the 3D framework of zinc sponge anodes upon deep discharge in Zn-air cells. ACS

Appl. Mater. Interfaces 6, pp. 19471-19476 (2014).

55. Ko, J. S., Sassin, M. B., Parker, J. F., Rolison, D. R. & Long, J. W. Combining battery-like and

pseudocapacitive charge storage in 3D MnOx@ carbon electrode architectures for zinc-ion cells.

Sustainable Energy & Fuels 2, 626-636 (2018).

56. Chamoun, M., Hertzberg, B. J., Gupta, T., Davies, D., Bhadra, S., Van Tassell, B., Erdonmez, C. &

Steingart, D. A. Hyper-dendritic nanoporous zinc foam anodes. NPG Asia Materials 7, e178

(2015).

57. Hilder, M., Winther-Jensen, O., Winther-Jensen, B. & MacFarlane, D. R. Graphene/zinc nano-

composites by electrochemical co-deposition. Phys. Chem. Chem. Phys. 14, 14034–14040

Page 71: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

58

(2012).

58. Parker, J. F., Chervin, C. N., Nelson, E. S., Rolison, D. R. & Long, J. W. Wiring zinc in three

dimensions re-writes battery performance—dendrite-free cycling. Energy Environ. Sci. 7, 1117–

1124 (2014).

59. McKubre, M. C. H. & Macdonald, D. D. The Dissolution and Passivation of Zinc in Concentrated

Aqueous Hydroxide. J. Electrochem. Soc. 128, 524–530 (1981).

60. Pan, H., Shao, Y., Yan, P., Cheng, Y., Han, K. S., Nie, Z., Wang, C., Yang, J., Li, X., Bhattacharya, P.,

Mueller, K. T. & Liu, J. Reversible aqueous zinc/manganese oxide energy storage from

conversion reactions. Nature Energy 1, Article number: 16039 (2016).

61. Lee, B., Seo, H. R., Lee, H. R., Yoon, C. S., Kim, J. H., Chung, K. Y., Cho, B. W. & Oh, S. H. Critical

Role of pH Evolution of Electrolyte in the Reaction Mechanism for Rechargeable Zinc Batteries.

ChemSusChem 9, 2948–2956 (2016).

62. Kim, S. H. & Oh, S. M. Degradation mechanism of layered MnO2 cathodes in Zn/ZnSO4/MnO2

rechargeable cells. J. Power Sources 72, 150–158 (1998).

63. Xu, D., Li, B., Wei, C., He, Y.-B., Du, H., Chu, X., Qin, X., Yang, Q.-H. & Kang, F. Preparation and

Characterization of MnO2/acid-treated CNT Nanocomposites for Energy Storage with Zinc Ions.

Electrochim. Acta 133, 254–261 (2014).

64. Sun, W., Wang, F., Hou, S., Yang, C., Fan, X., Ma, Z., Gao, T., Han, F., Hu, R., Zhu, M. & Wang, C.

Zn/MnO2 Battery Chemistry With H(+) and Zn(2+) Coinsertion. J. Am. Chem. Soc. 139, 9775–

9778 (2017).

65. Biswal, A., Tripathy, B. C., Sanjay, K., Subbaiah, T. & Minakshi, M. Electrolytic manganese

dioxide (EMD): a perspective on worldwide production, reserves and its role in

electrochemistry. RSC Adv. 5, 58255–58283 (2015).

66. Dose, W. M. & Donne, S. W. Optimising heat treatment environment and atmosphere of

electrolytic manganese dioxide for primary Li/MnO2 batteries. J. Power Sources 247, 852–857

(2014).

Page 72: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

59

67. Ortiz-Vitoriano, N., Drewett, N. E., Gonzalo, E. & Rojo, T. High performance manganese-based

layered oxide cathodes: overcoming the challenges of sodium ion batteries. Energy Environ. Sci.

10, 1051–1074 (2017).

68. Kim, J.-S., Chang, W.-S., Kim, R.-H., Kim, D.-Y., Han, D.-W., Lee, K.-H., Lee, S.-S. & Doo, S.-G. High-

capacity nanostructured manganese dioxide cathode for rechargeable magnesium ion

batteries. J. Power Sources 273, 210–215 (2015).

69. Hertzberg, B. J., Huang, A., Hsieh, A., Chamoun, M., Davies, G., Seo, J. K., Zhong, Z., Croft, M.,

Erdonmez, C., Meng, Y. S. & Steingart, D. Effect of Multiple Cation Electrolyte Mixtures on

Rechargeable Zn–MnO2 Alkaline Battery. Chem. Mater. 28, 4536–4545 (2016).

70. Julien, C. M., Massot, M. & Poinsignon, C. Lattice vibrations of manganese oxides. Part I.

Periodic structures. Spectrochim. Acta A Mol. Biomol. Spectrosc. 60, 689–700 (2004).

71. Gallaway, J. W., Menard, M., Hertzberg, B., Zhong, Z., Croft, M., Sviridov, L. A., Turney, D. E.,

Banerjee, S., Steingart, D. A. & Erdonmez, C. K. Hetaerolite Profiles in Alkaline Batteries

Measured by High Energy EDXRD. J. Electrochem. Soc. 162, A162–A168 (2015).

72. Chamoun, M., Brant, W. R., Tai, C.-W., Karlsson, G. & Noréus, D. Rechargeability of aqueous

sulfate Zn/MnO2 batteries enhanced by accessible Mn2+ ions. Energy Storage Materials 15, 351–

360 (2018).

73. Gallaway, J. W., Hertzberg, B. J., Zhong, Z., Croft, M., Turney, D. E., Yadav, G. G., Steingart, D. A.,

Erdonmez, C. K. & Banerjee, S. Operando identification of the point of [Mn2]O4 spinel formation

during γ-MnO2 discharge within batteries. J. Power Sources 321, 135–142 (2016).

74. Boden, D., Venuto, C. J., Wisler, D. & Wylie, R. B. The Alkaline Manganese Dioxide Electrode: I .

The Discharge Process. J. Electrochem. Soc. 114, 415–417 (1967).

75. Devaraj, S. & Munichandraiah, N. Effect of Crystallographic Structure of MnO2 on Its

Electrochemical Capacitance Properties. J. Phys. Chem. C 112, 4406–4417 (2008).

76. McBreen, J. The electrochemistry of β-MnO2 and γ-MnO2 in alkaline electrolyte. Electrochim.

Acta 20, 221–225 (1975).

Page 73: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

60

77. Maskell, W. C., Shaw, J. E. A. & Tye, F. L. Manganese dioxide electrode—IV. Chemical and

electrochemical reduction of an electrolytic γ-MnO2. Electrochim. Acta 26, 1403–1410 (1981).

78. Mondoloni, C., Laborde, M., Rioux, J., Andoni, E. & Lévy‐Clément, C. Rechargeable Alkaline

Manganese Dioxide Batteries: I . In Situ X‐Ray Diffraction Investigation of the (EMD‐Type)

Insertion System. J. Electrochem. Soc. 139, 954–959 (1992).

79. Zhao, S., Han, B., Zhang, D., Huang, Q., Xiao, L., Chen, L., Ivey, D. G., Deng, Y. & Wei, W.

Unravelling the reaction chemistry and degradation mechanism in aqueous Zn/MnO2

rechargeable batteries. J. Mater. Chem. A Mater. Energy Sustain. 6, 5733–5739 (2018).

80. Huang, J., Wang, Z., Hou, M., Dong, X., Liu, Y., Wang, Y. & Xia, Y. Polyaniline-intercalated

manganese dioxide nanolayers as a high-performance cathode material for an aqueous zinc-ion

battery. Nat. Commun. 9, 2906 (2018).

81. Wagner, O. C. Secondary iron-air batteries. (1968).

82. Öjefors, L. Self-discharge of the alkaline iron electrode. Electrochim. Acta 21, 263–266 (1976).

83. Öjefors, L. & Carlsson, L. An iron—air vehicle battery. J. Power Sources 2, 287–296 (1978).

84. Paulraj, A. R., Kiros, Y., Skårman, B. & Vidarsson, H. Core/Shell Structure Nano-Iron/Iron Carbide

Electrodes for Rechargeable Alkaline Iron Batteries. J. Electrochem. Soc. 164, A1665–A1672

(2017).

85. Figueredo-Rodríguez, H. A., McKerracher, R. D., Insausti, M., Luis, A. G., de Leόn, C. P., Alegre,

C., Baglio, V., Aricò, A. S. & Walsh, F. C. A Rechargeable, Aqueous Iron Air Battery with

Nanostructured Electrodes Capable of High Energy Density Operation. J. Electrochem. Soc. 164,

A1148–A1157 (2017).

86. Hang, B. T., Eashira, M., Watanabe, I., Okada, S., Yamaki, J.-I., Yoon, S.-H. & Mochida, I. The

effect of carbon species on the properties of Fe/C composite for metal–air battery anode. J.

Power Sources 143, 256–264 (2005).

87. Hang, B. T., Watanabe, T., Egashira, M., Watanabe, I., Okada, S. & Yamaki, J.-I. The effect of

additives on the electrochemical properties of Fe/C composite for Fe/air battery anode. J.

Page 74: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

61

Power Sources 155, 461–469 (2006).

88. Sundar Rajan, A., Ravikumar, M. K., Priolkar, K. R., Sampath, S. & Shukla, A. K. Carbonyl-Iron

Electrodes for Rechargeable-Iron Batteries. Electrochemical Energy Technology 1, (2014).

89. Manohar, A. K., Malkhandi, S., Yang, B., Yang, C., Surya Prakash, G. K. & Narayanan, S. R. A High-

Performance Rechargeable Iron Electrode for Large-Scale Battery-Based Energy Storage. J.

Electrochem. Soc. 159, A1209–A1214 (2012).

90. Manohar, A. K., Yang, C. & Narayanan, S. R. The Role of Sulfide Additives in Achieving Long Cycle

Life Rechargeable Iron Electrodes in Alkaline Batteries. J. Electrochem. Soc. 162, A1864–A1872

(2015).

91. Yang, B., Malkhandi, S., Manohar, A. K., Prakash, G. & Narayanan, S. R. Organo-sulfur molecules

enable iron-based battery electrodes to meet the challenges of large-scale electrical energy

storage. Energy Environ. Sci. 7, 2753-2763 (2014).

92. Li, Y. & Dai, H. Recent advances in zinc–air batteries. Chem. Soc. Rev. 43, 5257–5275 (2014).

93. Wang, Z.-L., Xu, D., Xu, J.-J. & Zhang, X.-B. Oxygen electrocatalysts in metal–air batteries: from

aqueous to nonaqueous electrolytes. Chem. Soc. Rev. 43, 7746–7786 (2014).

94. Kinoshita, K. Electrochemical oxygen technology. 30, (John Wiley & Sons, 1992).

95. Jörissen, L. Bifunctional oxygen/air electrodes. J. Power Sources 155, 23–32 (2006).

96. Christensen, P. A., Hamnett, A. & Linares-Moya, D. Oxygen reduction and fuel oxidation in

alkaline solution. Phys. Chem. Chem. Phys. 13, 5206–5214 (2011).

97. Suntivich, J., Gasteiger, H. A., Yabuuchi, N., Nakanishi, H., Goodenough, J. B. & Shao-Horn, Y.

Design principles for oxygen-reduction activity on perovskite oxide catalysts for fuel cells and

metal–air batteries. Nat. Chem. 3, 546–550 (2011).

98. Vielstich, W., Gasteiger, H. A. & Yokokawa, H. Handbook of Fuel Cells: Fundamentals Technology

and Applications: Advances in Electrocatalysis, Materials, Diagnostics and Durability. 5, (John

Wiley & Sons, 2009).

99. Mattioli, G., Giannozzi, P., Amore Bonapasta, A. & Guidoni, L. Reaction pathways for oxygen

Page 75: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

62

evolution promoted by cobalt catalyst. J. Am. Chem. Soc. 135, 15353–15363 (2013).

100. Hamdani, M., Singh, R. N. & Chartier, P. Co3O4 and Co-based spinel oxides bifunctional oxygen

electrodes. Int. J. Electrochem. Sci. 5, 556–577 (2010).

101. Hu, L., Wu, L., Liao, M., Hu, X. & Fang, X. Electrical Transport Properties of Large, Individual

NiCo2O4 Nanoplates. Adv. Funct. Mater. 22, 998–1004 (2012).

102. Zhao, Q., Yan, Z., Chen, C. & Chen, J. Spinels: Controlled Preparation, Oxygen

Reduction/Evolution Reaction Application, and Beyond. Chem. Rev. 117, 10121–10211 (2017).

103. Yuan, C., Wu, H. B., Xie, Y. & Lou, X. W. Mixed transition-metal oxides: design, synthesis, and

energy-related applications. Angew. Chem. Int. Ed. 53, 1488–1504 (2014).

104. Karlsson, G. Perovskite catalysts for air electrodes. Electrochim. Acta 30, 1555–1561 (1985).

105. Suntivich, J., May, K. J., Gasteiger, H. A., Goodenough, J. B. & Shao-Horn, Y. A perovskite oxide

optimized for oxygen evolution catalysis from molecular orbital principles. Science 334, 1383–

1385 (2011).

106. Hamann, C. H., Hamnett, A. & Vielstich, W. Electrochemistry. 2nd. Completely Revised and

Updated Edition, New York (2007).

107. Chen, J., Lim, B., Lee, E. P. & Xia, Y. Shape-controlled synthesis of platinum nanocrystals for

catalytic and electrocatalytic applications. Nano Today 4, 81–95 (2009).

108. Lee, Y., Suntivich, J., May, K. J., Perry, E. E. & Shao-Horn, Y. Synthesis and Activities of Rutile IrO2

and RuO2 Nanoparticles for Oxygen Evolution in Acid and Alkaline Solutions. J. Phys. Chem. Lett.

3, 399–404 (2012).

109. McKerracher, R. D., Alegre, C., Baglio, V., Aricò, A. S., Ponce de León, C., Mornaghini, F., Rodlert,

M. & Walsh, F. C. A nanostructured bifunctional Pd/C gas-diffusion electrode for metal-air

batteries. Electrochim. Acta 174, 508–515 (2015).

110. Cheng, F. & Chen, J. Metal–air batteries: from oxygen reduction electrochemistry to cathode

catalysts. Chem. Soc. Rev. 41, 2172–2192 (2012).

111. Chamoun, M., Skårman, B., Vidarsson, H., Smith, R. I., Hull, S., Lelis, M., Milcius, D. & Noréus, D.

Page 76: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

63

Stannate Increases Hydrogen Evolution Overpotential on Rechargeable Alkaline Iron Electrodes.

J. Electrochem. Soc. 164, A1251–A1257 (2017).

112. Svengren, H., Chamoun, M. & Grins, J. Water Splitting Catalysis Studied by using Real‐Time

Faradaic Efficiency Obtained through Coupled Electrolysis and Mass Spectrometry. (2018).

113. Biendicho, J. J., Roberts, M., Offer, C., Noréus, D., Widenkvist, E., Smith, R. I., Svensson, G.,

Edström, K., Norberg, S. T., Eriksson, S. G. & Hull, S. New in-situ neutron diffraction cell for

electrode materials. J. Power Sources 248, 900–904 (2014).

114. Hull, S., Smith, R. I., David, W. I. F., Hannon, A. C., Mayers, J. & Cywinski, R. The Polaris powder

diffractometer at ISIS. Physica B Condens. Matter 180, 1000–1002 (1992).

115. Larson, A. C. & Von Dreele, R. B. Gsas. General Structure Analysis System. LANSCE, MS-H805, Los

Alamos, New Mexico (1994).

116. Blidberg, A., Gustafsson, T., Tengstedt, C., Björefors, F. & Brant, W. R. Monitoring LixFeSO4F (x=

1, 0.5, 0) Phase Distributions In Operando to Determine Reaction Homogeneity in Porous

Battery Electrodes. Chemistry of Materials (2017).

117. Leapman, R. D., Grunes, L. A. & Fejes, P. L. Study of the L2,3 edges in the 3d transition metals and

their oxides by electron-energy-loss spectroscopy with comparisons to theory. Phys. Rev. B

Condens. Matter 26, 614–635 (1982).

118. van Aken, P. A. & Liebscher, B. Quantification of ferrous/ferric ratios in minerals: new

evaluation schemes of Fe L2,3 electron energy-loss near-edge spectra. Phys. Chem. Miner. 29,

188–200 (2002).

119. Mayence, A., Wéry, M., Tran, D. T., Wetterskog, E., Svedlindh, P., Tai, C.-W. & Bergström, L.

Interfacial strain and defects in asymmetric Fe–Mn oxide hybrid nanoparticles. Nanoscale 8,

14171–14177 (2016).

120. Ibrahem, I., Iqbal, M. N., Verho, O., Eivazihollagh, A., Olsén, P., Edlund, H., Tai, C.-W., Norgren,

M. & Johnston, E. V. Copper Nanoparticles on Controlled Pore Glass and TEMPO for the Aerobic

Oxidation of Alcohols. ChemNanoMat 1, 1–6 (2017).

Page 77: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

64

121. Shen, Y. & Kordesch, K. The mechanism of capacity fade of rechargeable alkaline manganese

dioxide zinc cells. J. Power Sources 87, 162–166 (2000).

122. Im, D., Manthiram, A. & Coffey, B. Manganese(III) Chemistry in KOH Solutions in the Presence of

Bi-or Ba-Containing Compounds and Its Implications on the Rechargeability of γ-MnO2 in

Alkaline Cells. J. Electrochem. Soc. 150, A1651–A1659 (2003).

123. Musil, M., Choi, B. & Tsutsumi, A. Morphology and Electrochemical Properties of α-, β-, γ-, and

δ-MnO2 Synthesized by Redox Method. J. Electrochem. Soc. 162, A2058–A2065 (2015).

124. Garvie, L. A. J., Craven, A. J. & Brydson, R. Use of electron-energy loss near-edge fine structure

in the study of minerals. Am. Mineral. 79, 411–425 (1994).

125. Schmid, H. K. & Mader, W. Oxidation states of Mn and Fe in various compound oxide systems.

Micron 37, 426–432 (2006).

126. Laffont, L. & Gibot, P. High resolution electron energy loss spectroscopy of manganese oxides:

Application to Mn3O4 nanoparticles. Mater. Charact. 61, 1268–1273 (2010).

127. Tan, H., Verbeeck, J., Abakumov, A. & Van Tendeloo, G. Oxidation state and chemical shift

investigation in transition metal oxides by EELS. Ultramicroscopy 116, 24–33 (2012).

128. Lee, B., Lee, H. R., Kim, H., Chung, K. Y., Cho, B. W. & Oh, S. H. Elucidating the intercalation

mechanism of zinc ions into α-MnO2 for rechargeable zinc batteries. Chem. Commun. 51, 9265–

9268 (2015).

129. Massé, R. C. & Gerken, J. B. Assembly of a Robust and Economical MnO2-Based Reference

Electrode. J. Chem. Educ. 92, 110–115 (2015).

130. Rajan, A. S., Sampath, S. & Shukla, A. K. An in situ carbon-grafted alkaline iron electrode for

iron-based accumulators. Energy Environ. Sci. 7, 1110–1116 (2013).

131. Soler, L., Candela, A. M., Macanás, J., Muñoz, M. & Casado, J. Hydrogen generation from water

and aluminum promoted by sodium stannate. Int. J. Hydrogen Energy 35, 1038–1048 (2010).

132. Biendicho, J. J., Roberts, M., Noréus, D., Lagerqvist, U., Smith, R. I., Svensson, G., Norberg, S. T.,

Eriksson, S. G. & Hull, S. In situ investigation of commercial Ni (OH)2 and LaNi5-based electrodes

Page 78: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

65

by neutron powder diffraction. J. Mater. Res. 30, 407–416 (2015).

133. Liu, M. & Liu, L.-G. Compressions and phase transitions of tin to half a megabar. High

Temperatures. High Pressures 18, 79–85 (1986).

134. Dobersek, M., Kosovinc, I. & Schubert, K. Metallographie und Konstitution der Fe-Ecke des

Systems Fe-C-Sn. Archiv für das Eisenhüttenwesen 263–266 (1984).

135. Degtyareva, V. F., Degtyareva, O. & Allan, D. R. Ordered Si-VI-type crystal structure in BiSn alloy

under high pressure. Phys. Rev. B Condens. Matter 67, 212105 (2003).

136. Savidan, J.-C., Joubert, J.-M. & Toffolon-Masclet, C. An experimental study of the Fe--Sn--Zr

ternary system at 900° C. Intermetallics 18, 2224–2228 (2010).

137. Lei, D., Lee, D.-C., Magasinski, A., Zhao, E., Steingart, D. A. & Yushin, G. Performance

Enhancement and Side Reactions in Rechargeable Nickel-Iron Batteries with Nanostructured

Electrodes. ACS Appl. Mater. Interfaces 8, 2088-2096 (2016).

138. Paulraj, A. R., Kiros, Y., Chamoun, M., Svengren, H., Noréus, D., Göthelid, M., Skårman, B.,

Vidarsson, H. & Johansson, M. Electrochemical Performance and in Operando Charge Efficiency

Measurements of Cu/Sn-Doped Nano Iron Electrodes, Batteries 5, 1-15 (2019).

139. Taguchi, H. Relationship between Crystal Structure and Electrical Properties of the Ca-Rich

Region in (La1- xCax) MnO2. 97. J. Solid State Chem. 124, 360–365 (1996).

140. Zhu, J. L., Yu, R. C., Li, F. Y., Jin, C. Q. & Zhang, Z. The structure and properties of the manganate

with nominal composition La1.0Ca2.0Mn2O7. Mater. Sci. Eng. B 95, 19–23 (2002).

141. Tang, D., Liu, J., Wu, X., Liu, R., Han, X., Han, Y., Huang, H., Liu, Y. & Kang, Z. Carbon quantum

dot/NiFe layered double-hydroxide composite as a highly efficient electrocatalyst for water

oxidation. ACS Appl. Mater. Interfaces 6, 7918–7925 (2014).

142. Zhang, K., Wang, W., Kuai, L. & Geng, B. A facile and efficient strategy to gram-scale preparation

of composition-controllable Ni-Fe LDHs nanosheets for superior OER catalysis. Electrochim. Acta

225, 303–309 (2017).

143. Lapham, D. P. & Tseung, A. C. C. The effect of firing temperature, preparation technique and

Page 79: DiVA portal1271235/FULLTEXT01.pdf · SHE Standard hydrogen electrode SMES Superconducting magnetic energy storage SO C State -of-charge SS Stainless steel XPS X-ray photoelectron

66

composition on the electrical properties of the nickel cobalt oxide series NixCo1 − xOy. J. Mater.

Sci. 39, 251–264 (2004).

144. Paulraj, A. R. & Kiros, Y. La0.1Ca0.9MnO3/Co3O4 for oxygen reduction and evolution reactions

(ORER) in alkaline electrolyte. J. Solid State Electrochem. 22, 1697–1710 (2018).

145. Flegler, A., Hartmann, S., Settelein, J., Mandel, K. & Sextl, G. Screen printed bifunctional gas

diffusion electrodes for aqueous metal-air batteries: Combining the best of the catalyst and

binder world. Electrochim. Acta 258, 495–503 (2017).

146. Price, S. W. T., Thompson, S. J., Li, X., Gorman, S. F., Pletcher, D., Russell, A. E., Walsh, F. C. &

Wills, R. G. A. The fabrication of a bifunctional oxygen electrode without carbon components

for alkaline secondary batteries. J. Power Sources 259, 43–49 (2014).

147. Paulraj, A., Kiros, Y., Göthelid, M. & Johansson, M. NiFeOx as a Bifunctional Electrocatalyst for

Oxygen Reduction (OR) and Evolution (OE) Reaction in Alkaline Media. Catalysts 8, 328 (2018).