Rechargeable Organic Batteries - Yongzhu Fu - E-Book

Rechargeable Organic Batteries E-Book

Yongzhu Fu

0,0
124,99 €

-100%
Sammeln Sie Punkte in unserem Gutscheinprogramm und kaufen Sie E-Books und Hörbücher mit bis zu 100% Rabatt.
Mehr erfahren.
Beschreibung

This book focuses on the next generation sustainable organic electrodes and provides a guideline for the future development of these materials.

Das E-Book Rechargeable Organic Batteries wird angeboten von Wiley-VCH GmbH und wurde mit folgenden Begriffen kategorisiert:
Akkumulator, Batterien u. Brennstoffzellen, Batteries & Fuel Cells, Chemie, Chemistry, Energie, Energy, Hydrogen, Batteries & Fuel Cells, Materialien f. Energiesysteme, Materials for Energy Systems, Materials Science, Materialwissenschaften, Wasserstoff, Wasserstoff, Batterien u. Brennstoffzellen

Sie lesen das E-Book in den Legimi-Apps auf:

Android
iOS
von Legimi
zertifizierten E-Readern

Seitenzahl: 508

Bewertungen
0,0
0
0
0
0
0
Mehr Informationen
Mehr Informationen
Legimi prüft nicht, ob Rezensionen von Nutzern stammen, die den betreffenden Titel tatsächlich gekauft oder gelesen/gehört haben. Wir entfernen aber gefälschte Rezensionen.



Table of Contents

Cover

Table of Contents

Title Page

Copyright

Preface

1 Necessity and Advantages of Developing Rechargeable Organic Batteries

1.1 Current Electrochemical Energy Storage Technologies

1.2 Rechargeable Organic Batteries

1.3 Goal, Scope, and Organization of this Book

Acknowledgments

References

2 Redox Mechanisms and Characterization Methods of Organic Electrode Materials

2.1 Introduction

2.2 Carbonyl Materials

2.3 Organosulfide Materials

2.4 Radical Materials

2.5 N-Containing Active Materials

2.6 Summary and Outlook

Acknowledgments

References

3 Carbonyl-Based Organic Cathodes

3.1 Introduction

3.2 Quinone Compounds

3.3 Imides

3.4 Anhydrides

3.5 Summary and Outlook

Acknowledgments

References

4 Sulfur-Containing Organic Cathodes

4.1 Introduction

4.2 Organodisulfide

4.3 Organopolysulfides

4.4 Heteroatom-Containing Organosulfides

4.5 Organosulfur–Inorganic Hybrid Cathodes

4.6 Other Organosulfur Cathodes

4.7 Summary and Outlooks

Acknowledgments

References

5 Radical-Based Organic Cathodes

5.1 Introduction

5.2 Radical for Metal-Ion Battery

5.3 Radicals for Redox Flow Batteries

5.4 Summary and Prospect

Acknowledgments

References

6 Organometallic Complexes-Based Electrodes

6.1 Introduction

6.2 Small Molecules

6.3 1D MOF

6.4 2D MOF

6.5 3D MOF

6.6 Summary and Outlook

Acknowledgments

References

7 Polymer-Based Organic Cathodes

7.1 Introduction

7.2 Organosulfur Polymers

7.3 Carbonyl-Derived Polymers

7.4 Covalent Organic Frameworks-Derived Polymers

7.5 Organic Radical-Derived Polymers

7.6 Other Polymers

7.7 Summary and Outlook

Acknowledgments

References

8 Organic Anode

8.1 Introduction

8.2 Conjugated Carboxylates

8.3 Schiff Bases

8.4 Azo Compounds

8.5 Covalent Organic Frameworks

8.6 Thiophene Compounds

8.7 Summary and Outlook

Acknowledgments

References

9 All-Organic Batteries

9.1 Introduction

9.2 Traditional Batteries

9.3 Flow Batteries Based on Organic Molecules

9.4 Summary and Outlook

Acknowledgments

References

10 Outlook

Acknowledgments

List of Abbreviations

Index

End User License Agreement

List of Tables

Chapter 8

Table 8.1 Performance comparison of reported conjugated dicarboxylates in re...

Chapter 9

Table 9.1 Ionic conductivities of different supporting electrolytes in aqueo...

List of Illustrations

Chapter 1

Figure 1.1 Published papers per year for the rechargeable organic batteries ...

Figure 1.2 Typical organic materials with voltage, specific capacity, and en...

Chapter 2

Figure 2.1 Three types of reaction mechanisms and some representative molecu...

Figure 2.2 Redox behavior of a carbonyl unit, R represents one organic funct...

Figure 2.3 Redox behaviors of four categories of carbonyl redox-active mater...

Figure 2.4 (a) Multielectron redox process of the BBQB electrode in lithium ...

Figure 2.5 (a) Electrochemical redox reactions of PIBN in lithium batteries....

Figure 2.6 (a) Schematic diagram of in situ NMR analysis.(b) Molecular s...

Figure 2.7 (a) Molecular structure of Na

2

C

6

O

6

and its corresponding crystal....

Figure 2.8 Redox mechanisms of three representative n-type organosulfides.

Figure 2.9 The redox reaction process of an oligomer.

Figure 2.10 Redox mechanisms of two representative p-type organosulfides.

Figure 2.11 (a) Redox mechanism of the TETD electrode. (b) UV–vis spectra of...

Figure 2.12 (a) Molecular structure of a disulfide polymer and its reduction...

Figure 2.13 Redox behaviors of nitroxyl, phenoxyl, and blatter radicals (R i...

Figure 2.14 (a) The voltage curve of lithium-TEMPO cell during charging. (b)...

Figure 2.15 The redox mechanism of an azo material, azobenzene.

Figure 2.16 Schemes 1 and 2, redox mechanisms of phenazine materials. Scheme...

Figure 2.17 The redox behavior of the dilithium 1,4-phenylenebis((methylsulf...

Figure 2.18 The redox reaction of a representative nitroaromatic material, 1...

Figure 2.19 Scheme 1, the redox reaction of a representative phenylamine mat...

Figure 2.20 (a) The redox reaction of LC in lithium batteries. (b) Ex situ F...

Figure 2.21 (a) Redox reaction of the 3Q electrode in lithium batteries. (b)...

Figure 2.22 Ex situ FTIR (a) and Raman (b) spectra of the 1,4-dinitrobenzene...

Chapter 3

Figure 3.1 The diagram of the redox reaction mechanism of carbonyl-based org...

Figure 3.2 (a) The structures of the four heteroaromatic-fused carbonyl comp...

Figure 3.3 (a) Molecular structures and electrochemical mechanism of DTT. (b...

Figure 3.4 Molecular structures of (a) DANQ and (b) IMNQ. (c) Discharge/char...

Figure 3.5 (a) Chemical structures of TT.(b) The electrochemical redox m...

Figure 3.6 (a) The synthesis and electrochemical redox mechanism of TPB....

Figure 3.7 (a) Structures of NNS, NSN, and NSS.(b) The chemical structur...

Figure 3.8 (a) The proposed electrochemical redox reactions of EV-AQ

2

cathod...

Figure 3.9 (a) Synthesis route from IDT to oIDT and the sample photographs....

Figure 3.10 (a) The synthetic route proposed electrochemical redox mechanism...

Figure 3.11 (a) Synthetic routes of QPQ-1 and QPQ-2. (b) The long-term capab...

Figure 3.12 (a) Schematic diagram of preparing C4Q. (b) Cycling performance ...

Figure 3.13 (a) Diagram illustrating the H

2

O-assisted phase transfer mechani...

Figure 3.14 (a) Illustration of the reversible reaction mechanism of the aqu...

Figure 3.15 (a) The synthetic pathway for TAPQ. (b) The calculated binding e...

Figure 3.16 (a) Preparation, structures, and theoretic reactions of designed...

Figure 3.17 (a) Potassium storage mechanism of AQDS. (b) The cycling stabili...

Figure 3.18 (a) Potassium storage mechanism and (b) CV curves of P5Q. (c) Cy...

Figure 3.19 (a) Structural formula of DMBQ.(b) Crystal structure of Na

2

C

Figure 3.20 (a) Illustration of electrochemical redox mechanism of TDK. (b) ...

Figure 3.21 (a) Schematic of cell in charge mode.(b) Illustration of mol...

Figure 3.22 (a) Schematic flow battery with three negative electrolytes (neg...

Figure 3.23 Synthesis process of AQSCH and BQDS(CH)

2

.

Figure 3.24 Molecular structure of some imides. (a) pyromellitic diimide dil...

Figure 3.25 (a) Charge/discharge curves of 3,4,9,10-perylene-bis(dicarboximi...

Figure 3.26 (a) Molecular structure of naphthalene diimide, perylene diimide...

Figure 3.27 (a) Schematic representation of the REDOX reaction of PTCDA....

Chapter 4

Figure 4.1 The development of small-molecule organosulfides for batteries. T...

Figure 4.2 Organodisulfides and their theoretical capacities.

Figure 4.3 (a) Schematic illustration of the DPDS@CNT cathode preparation....

Figure 4.4 Structural formula and theoretical capacities of (a) naphtho[1,8-

Figure 4.5 (a) Proposed lithiation mechanisms of TTT and HTP, determined by ...

Figure 4.6 (a) Schematic illustration of the synthesis of diphenyl trisulfid...

Figure 4.7 The redox reactions of (a) lenthionine and (b) lenthionine with D...

Figure 4.8 (a) Equation along with a visual representation of the phenyl pol...

Figure 4.9 Redox reactions of (a) 1,2-LBDT, (b) 1,3-LBDT, and (c) 1,4-LBDT d...

Figure 4.10 (a) Synthesis of HDBCO with 1,3-benzenedithiol and elemental sul...

Figure 4.11 (a) The comparison of the organosulfur compounds with inorganic ...

Figure 4.12 (a) Optical image of the polyethylene hexasulfide polymer in 1:1...

Figure 4.13 The molecular structure of tricyanuric acid (TTCA), polymers of ...

Figure 4.14 (a) Optical image of commercial PMTT powder along with its chemi...

Figure 4.15 (a) Chemical structures and theoretical specific capacities of D...

Figure 4.16 (a) Scheme of reaction of DPDS and DPDSe to form PhS-SePh along ...

Figure 4.17 Redox reactions of (a) DPDSe-S and (b) DPDSe-S

2

in rechargeable ...

Figure 4.18 (a) A schematic diagram of the reaction of DPDS with PhTe-TePh t...

Figure 4.19 (a) Schematic of the synthesis of BDPPTS. (b) Proposed reaction ...

Figure 4.20 (a) Schematic illustration showing the formation mechanism of po...

Figure 4.21 (a) Voltage profiles of pure DPTtS and DPTtS/TiS

2

NSs electrodes...

Figure 4.22 (a) Schematic representation of the structure of a Li/DPTtS@VSGC...

Figure 4.23 (a) The structure of TTMD.Structures and redox schemes of (b...

Figure 4.24 (a) Molecular transformation mechanism of TMTM. (b) Proposed mol...

Figure 4.25 Prospects of organosulfur-based batteries.

Chapter 5

Figure 5.1 Redox mechanisms of radicals.

Figure 5.2 TEMPO radical polymers, other nitroxyl radicals, and other radica...

Figure 5.3 Structure redox mechanism of TEMPO.

Figure 5.4 (a) Synthesis procedure of PTMA.(b) The cycle performance of ...

Figure 5.5 (a) Redox reaction of PTVE.(b) Cycle performance and charging...

Figure 5.6 (a) Discharge/charge and corresponding (b) profiles of Li-O

2

batt...

Figure 5.7 (a) Optical photo of the materials for both routes in each step a...

Figure 5.8 (a) Discharged and charged reaction of 4-OH-TEMPO and MV

2+

. (...

Figure 5.9 (a) Synthetic routes for CPL. (b) CV curves of CPL (red trace) an...

Figure 5.10 (a) Schematic of RFB assembled with TMAP-TEMPO and BTMAP-Vi an a...

Figure 5.11 (a) Structural configuration and redox processes of Pyr-TEMPO/[P...

Figure 5.12 (a) Rheological measurements of 13 mg ml

−1

PTMA

63

-b-PS

35

a...

Figure 5.13 Summary of Design and Application for the active radical materia...

Chapter 6

Figure 6.1 (a) The charge and discharge mechanism of the NiNC in Li batterie...

Figure 6.2 (a) The bipolar mechanism of the CuTAPc electrode in Li batteries...

Figure 6.3 The molecular structure of (a) poly(ethynylferrocene) and (b) pol...

Figure 6.4 (a) The scheme of the ion breathing mode of the all-organic batte...

Figure 6.5 (a) The structure and (b) the electrode reactions of the redox fl...

Figure 6.6 (a) The scheme of the charge and discharge mechanism of the Sn-BD...

Figure 6.7 (a) The synthesis route of Ni-BTA. The (b) voltage profile and th...

Figure 6.8 (a) The structure of the M-HIPA MOF. (b) The voltage profile and ...

Figure 6.9 (a) The molecular structure of the Ni-HAB. (b) The voltage profil...

Figure 6.10 The charge/discharge schematic of the (a) Zn-HHTP and (b) Cu-HHT...

Figure 6.11 Schematic of energy storage (a) and the cycling performance (b) ...

Figure 6.12 (a) The schematic of the energy storage mechanism of the Zn-PTCA...

Chapter 7

Figure 7.1 The different polymer cathodes in metal batteries.

Figure 7.2 (a) The copolymerization of S

8

with DIB. (b) The CV curves of Li–...

Figure 7.3 (a) Synthesis scheme of polySGNs.(b) The cycling performance ...

Figure 7.4 (a) One-pot synthesis of SLP from D-limonene and sulfur at heat t...

Figure 7.5 (a) Schematic diagram of the synthesis of SRVCR. (b) In-situ UV/V...

Figure 7.6 (a) The synthetic mechanism of selenium-doped PDATtSSe polymer....

Figure 7.7 (a) The chemical structure of copolymer by reaction of Ca monomer...

Figure 7.8 (a) Schematic drawings describing the synthetic procedures of sul...

Figure 7.9 (a) The synthesis scheme and corresponding visual representation ...

Figure 7.10 (a) Synthesis of SPAN at 300

o

C by Wang et al.(b) The propos...

Figure 7.11 (a) Proposed synthesis route for SPAN nanocomposite cathode mate...

Figure 7.12 (a) The reaction process scheme of proposed Li–Se

0.06

SPAN and Li...

Figure 7.13 (a) Proposed overall charge and discharge reaction of SPAN used ...

Figure 7.14 (a) Synthesis of S-CTF-1.(b) Schematic synthesis of the FCTF...

Figure 7.15 (a) Schematic synthesis of the S/P-CTFs and S/P-CTFs@rGO composi...

Figure 7.16 Schematic representation of the redox reactions of carbonyl-deri...

Figure 7.17 Some typical molecular structures of PQs.

Figure 7.18 (a) The synthesis route of PDBS. (b)

13

C solid-state NMR spectra...

Figure 7.19 (a) The synthesis route of PBQS.(b) The synthesis route of P...

Figure 7.20 (a) The electrochemical mechanism scheme of PADAQ.(b) Potent...

Figure 7.21 (a) Rate performance and (b) the corresponding voltage profiles ...

Figure 7.22 Some typical molecular structures of PIs.

Figure 7.23 (a) Voltage profiles and (b) CV curves of the five polyimide sam...

Figure 7.24 (a) Structure of different polyimides. Charge–discharge curves o...

Figure 7.25 (a) Schematic diagram of the synthesis of PPN and PTN. (b) Rate ...

Figure 7.26 (a) Synthesis of Tp-DANT-COF and Tb-DANT-COF.(b) Synthesis r...

Figure 7.27 (a) Synthesis of BQ1-COF. (b) Diagram of the single-layer conjug...

Figure 7.28 (a) Schematic diagram for the exfoliation of 2D redox-active COF...

Figure 7.29 Some typical molecular structures of TEMPO-based radical polymer...

Figure 7.30 (a) Schematic illustration of the PTMA-impregnated CNT structure...

Figure 7.31 (a) The synthesis route of the TEMPO-substituted polynorbornenes...

Figure 7.32 (a) The synthesis of CL-PTGE. (b) Schematic diagram of the charg...

Figure 7.33 (a) XPS of the PTPAn electrode at different states in the PDIB h...

Figure 7.34 (a) Long-term cycling performance of the Na/PHATN cell at 25 C f...

Figure 7.35 Summary illustration of polymer-based cathode materials in recha...

Chapter 8

Figure 8.1 (a) Redox mechanism of aromatic dicarboxylates. (b) Charge–discha...

Figure 8.2 The modified strategies of dicarboxylates.

Figure 8.3 Proposed lithiation mechanism of transition metal-tetrephthalate....

Figure 8.4 (a) Schematic illustration of the preparation of heteromat Ni-NDC...

Figure 8.5 (a) The redox mechanism of Li

2

-NDC.(b) Comparison of the dist...

Figure 8.6 (a) Redox mechanism and (b) long cycling performance of Na-PD ele...

Figure 8.7 (a) The structural evolution of Li

6

-HAT during the charge–dischar...

Figure 8.8 (a) Voltage profiles and (b) rate cyclability of Na

2

-SDC electrod...

Figure 8.9 Charge–discharge profiles of (a) Li

2

-ThDC.(b) Na

2

-PDC.(c)...

Figure 8.10 Charge–discharge profiles of (a) NH

2

-Na

2

-BDC, (b) Br-Na

2

-BDC, an...

Figure 8.11 The redox mechanism (a), (c), and cycling performance (b), (d) o...

Figure 8.12 (a) Charge–discharge profiles of Li/Li

2

C

6

H

4

O

4

cell at 0.1 C rate...

Figure 8.13 The structure of the reported conjugated dicarboxylates in recha...

Figure 8.14 (a) The redox mechanism of Schiff bases in rechargeable batterie...

Figure 8.15 (a) The synthesis of oligomeric Schiff-bases vis the condensatio...

Figure 8.16 (a) CV profiles of NBALS electrode at 0.1 mV s

−1

. (b) Cycl...

Figure 8.17 The cyclic voltage profiles of (a) ADASS and (c) ADAPS. The corr...

Figure 8.18 (a) Interaction between PBALS and LPS solid-state electrolyte. T...

Figure 8.19 The structure of the reported COFs as anode materials in recharg...

Figure 8.20 (a) Synthetic routes of PT and P33DT. Discharge-charge profiles ...

Chapter 9

Figure 9.1 Cell configurations of all-organic batteries based on n-type and ...

Figure 9.2 (a) Cell unit made of two electrodes of polyaniline. 1, Polyanili...

Figure 9.3 (a) The chemical structures of the negative electrode material PE...

Figure 9.4 (a) Cartoon (top) and photo (bottom) structure of the assembled b...

Figure 9.5 The working mechanism and molecule structure of the symmetric all...

Figure 9.6 Schematic illustration of an all-organic battery with metallic ch...

Figure 9.7 Schematic structure of an all-organic solid-state battery by Chia...

Figure 9.8 (a) The proposed redox mechanism of TAPT; (b) The optimized step-...

Figure 9.9 Illustration of corresponding redox reaction for (a) terephthalat...

Figure 9.10 (a) Proposed redox mechanism of PDB used as active bi-functional...

Figure 9.11 (a) Possible mechanism of the possible undesired O

2

evolution re...

Figure 9.12 (a) The proposed electrochemical reaction mechanism of PITN mate...

Figure 9.13 (a) The corresponding redox mechanism of terephthalate and nitro...

Figure 9.14 (a) Electrochemical redox mechanism of Na

4

C

8

H

2

O

6

, (b) Galvanosta...

Figure 9.15 (a) Illustration and (b) Charge–discharge curves of all-orgainc ...

Figure 9.16 (a) Schematic diagram of Teflon test cell based on poly (OC

10

DAS...

Figure 9.17 (a) Redox reactions mechanism and (b) Rate performance of the al...

Figure 9.18 (a) Redox reactions mechanism, and (b) Cycling performance of th...

Figure 9.19 Schematic of a typical redox flow battery.

Figure 9.20 (a) Synthesis of two-electron storage viologen molecules, (b) Sc...

Figure 9.21 (a) Redox reactions and (b) cyclic voltammetry (CV) curve of the...

Figure 9.22 (a) The reaction mechanism of the MePh|DBMMB, (b). Possible side...

Figure 9.23 Bipolar molecules for NORFBs: (a) Diagram of the symmetric flow ...

Guide

Cover

Table of Contents

Title Page

Copyright

Preface

Begin Reading

List of Abbreviations

Index

End User License Agreement

Pages

iii

iv

ix

x

1

2

3

4

5

6

7

8

9

10

11

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

273

274

275

276

277

278

279

280

281

282

283

285

286

287

288

289

290

291

292

293

Rechargeable Organic Batteries

Materials, Mechanisms, and Prospects

 

Yongzhu Fu

Xiang Li

Shuai Tang

Wei Guo

 

 

 

 

Authors

Prof. Yongzhu FuZhengzhou University100 Science AvenueZhengzhouChina 450001

Prof. Xiang LiZhengzhou UniversityCollege of Chemistry100 Science AvenueZhengzhouChina 450001

Prof. Shuai TangZhengzhou UniversityCollege of Chemistry100 Science AvenueZhengzhouChina 450001

Prof. Wei GuoZhengzhou UniversityCollege of Chemistry100 Science AvenueZhengzhouChina 450001

Cover: © Yagi Studio/Getty Images

All books published by WILEY-VCH are carefully produced. Nevertheless, authors, editors, and publisher do not warrant the information contained in these books, including this book, to be free of errors. Readers are advised to keep in mind that statements, data, illustrations, procedural details or other items may inadvertently be inaccurate.

Library of Congress Card No.: applied for

British Library Cataloguing-in-Publication DataA catalogue record for this book is available from the British Library.

Bibliographic information published by the Deutsche NationalbibliothekThe Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie; detailed bibliographic data are available on the Internet at <http://dnb.d-nb.de>.

© 2024 WILEY-VCH GmbH, Boschstraße 12, 69469 Weinheim, Germany

All rights reserved (including those of translation into other languages). No part of this book may be reproduced in any form – by photoprinting, microfilm, or any other means – nor transmitted or translated into a machine language without written permission from the publishers. Registered names, trademarks, etc. used in this book, even when not specifically marked as such, are not to be considered unprotected by law.

Print ISBN: 978-3-527-35080-3ePDF ISBN: 978-3-52783911-7ePub ISBN: 978-3-52783912-4oBook ISBN: 978-3-52783913-1

Preface

Energy plays a vital role in our society, and energy storage is important for our mobile lives and the usage of renewable energies such as solar and wind. As the state-of-the-art electrochemical energy storage technology, lithium-ion batteries (LIBs) are widely used in cell phones, laptops, and electric vehicles. LIBs were commercialized in 1991, and then their share in the energy storage market grew rapidly. However, with the rapid development of modern society, LIBs face a shortage of energy density and constraints of transition metals such as cobalt and nickel, which will become a challenge for a sustainable society. Alternative materials and systems are highly needed. Over the past years, organic batteries have attracted much attention because organic compounds are used as positive electrode materials, which are renewable and show a large difference from the traditional positive electrode materials, e.g. layered oxides, spinel oxides, and olivines. The energy storage mechanisms are well known for the ion-intercalation electrodes of LIBs. In contrast, organic materials undergo breakage/reformation of bonds, radical reactions, and anion compensation, which are quite different from those of transition metal oxides. Due to the variety of structures, organic electrode materials include carbonyl compounds, organosulfides, and N-containing organic compounds.

This book aims to provide a comprehensive “textbook” that covers most organic materials that have been studied as electrodes for rechargeable batteries. It includes various organic compounds, which can be positive and/or negative electrode materials. Beneficial to the variety of organic functional groups, the working conditions of organic batteries can be modified such as working voltage, open/sealed systems, and nonaqueous/aqueous systems. Some knowledge learned from the study of LIBs is also suitable for organic batteries such as the electrolyte design and interphase modifications between the electrode and electrolyte. Therefore, this book is suitable for readers with an electrochemistry and organic chemistry background including graduate students, postdoctoral fellows, and young faculty members. We hope this book brings new comprehension and possibilities in the field of electrochemical energy storage.

Finally, we gratefully acknowledge the substantial contribution from Dr. Qiliang Chen, Dr. Dan-Yang Wang, Dr. Qianqian Fan, Mr. Pengfei Sang, Mr. Fulong Zhu, Mr. Xiuqing Zhang, Ms. Siyuan Ren, Ms. Pei Yu, Ms. Hanshuang Song, and Ms. Qianhan Chen to this book.

   

   

Yongzhu Fu, Xiang Li, Shuai Tang, and Wei Guo

Zhengzhou, PR China

1Necessity and Advantages of Developing Rechargeable Organic Batteries

1.1 Current Electrochemical Energy Storage Technologies

Li-ion battery (LIB) is well known as one of the electrochemical energy storage (EES) technologies, which can be seen in our daily lives, such as portable equipment and electric vehicles. LIBs have made great progress in the last 30 years, which can be traced back to 1991, when the first reversible LIB was commercialized by Sony Corp. [1]. The battery is based on LiCoO2, graphite, and ester-solvents with LiPF6[2, 3]. Afterward, a series of ternary LiNixCoyMnzO2 (NCM, x + y + z = 1) and LiFePO4 spring out, considering the aspects of energy density and security [4–11]. Prior to LIBs, actually, lithium metal batteries (LMBs) were commercialized by Moli Energy Corp., based on lithium metal and metal sulfide as negative and positive electrodes, respectively [12]. However, the battery was in a tailspin after several safety incidents due to the lithium dendrites, which are easily generated after cycling [13, 14]. Note that for the discovery and development of LIBs, John B. Goodenough, M. Stanley Whittingham, and Akira Yoshino were awarded the 2019 Nobel Prize in Chemistry.

Similar to LIB, sodium-ion battery (SIB) is also one of the state-of-the-art EES technologies. Actually, SIBs have a longer history compared with LIBs, with the layered oxides discovered toward the end of the 1960s [15, 16]. Considering the limited Li resource, SIB is a suitable alternative EES due to the relatively abundant Na resource (420 times more than Li). It is noted that the oxides NaxMO2 (M = 3d element) have some special structure by regulating deficient sodium [17–20], such as O2, O3, P2, and P3 types, according to the structural packing described by Delmas [17]. Therefore, the electrochemical performance can be modified in Na-based oxides, which has an evident advantage compared with the Li-based oxides used for LIBs. The energy density of SIB could climb to 200 Wh kg−1, as reported by Hu’s group, which is a breakthrough [21]. However, the energy density is still limited with respect to LIBs, which could deliver over 300 Wh kg−1[22].

The good news is that SIBs have been commercialized by some Chinese companies such as CATL Corp. and HiNa Battery Corp. Given the energy density difference among the typical EES, the different battery systems aim for different market orientation to share the energy pressure. For example, lead-acid batteries, Ni–Cd batteries, and supercapacitors are used for devices with short mileage or low energy density, which are still required by the market. Nonetheless, both LIBs and SIBs cannot satisfy our demand in the long term, considering the resource crisis accompanied by high costs and pollution. Moreover, the traditional batteries are restricted to a sealed system and organic electrolytes (aqueous electrolytes are still facing great difficulties [23, 24]). Herein, we need a new EES without (or with mitigatory) the concerns.

A rechargeable organic battery is a good choice because the active materials are low cost, and the battery has comparable energy density when compared with LIB and SIB [25–31]. Moreover, the properties of organic materials can be controlled by different functional groups, such as the charge/discharge potential, the reaction dynamics, and the structural stability [32–37]. Furthermore, the system is unrestricted which can be used in an aqueous system, typically redox flow batteries (RFBs) [38–42]. Actually, the organic battery has been studied for over 60 years [43]. At the initial stage, the electrochemical performance of the organic material is poor with an ambiguous reaction mechanism, which impedes the development of the battery. In recent years, the corresponding published papers have a manifest rising tendency which can be seen in Figure 1.1, which partially benefits from technological advancements and several outstanding contributions made by Chen’s group and Schubert’s group since 2012 [44–69]. Now, there are many kinds of organics with active centers based on O, N, and S, enriching the family of organic batteries, which could compete with the traditional metal-ion batteries.

Figure 1.1 Published papers per year for the rechargeable organic batteries with keywords such as organic electrode, organic cathode, and organic battery. The time of the statistics is April 2023.

1.2 Rechargeable Organic Batteries

Organic batteries show different mechanisms from the typical LIBs, which mainly include insertion/extraction (LiCoO2, NCM, etc.), alloying (Al, Si, etc.), and conversion mechanism (O, P, S-based composites, etc.) [70–79]. For the organic electrodes, the mechanism usually contains the repeated breaking and bonding of a bond (carbonyl, organosulfide, and radical materials in Chapter 2). A single bond (typically S—S bond) is broken during discharge, after which the broken bond receives an electron and bonds with a metal ion for charge balance. A double bond (typically carbonyl units) shows a similar mechanism. Note that during the reaction, radical materials are usually generated which has been applied in RFBs due to the fast kinetics [80–82]. There seems to be another mechanism not involved in bond breaking, which is based on electron transfer and anion compensation (typically N-containing active materials) [83].

The first investigation of organic materials is carbonyl compounds, which can be traced back to the 1960s [43]. However, the material shows high solubility in aprotic electrolytes, restricting the application although some other carbonyl composites are constructed [32]. Afterward, the direction was turned to conductive polymers in the 1970s because of their less solubility, such as polyacetylene and polypyrrole [83–89]. Unfortunately, these electrodes suffered from limited capacity due to incomplete reaction [88, 89]. A revival emerged when Armand and Tarascon depicted a bright future for organic batteries, attracting more attention [28]. Encouragingly, molecules with popular functional groups (quinones, carboxylates, radical centers, etc.) and other redox-active centers (imines, alkenes, alkynes, azo, etc.) have been investigated [90–97].

Another typical organic material is organosulfide with S as the redox center. Visco et al. initially studied tetraethyl thiuram disulfide (TETD) in 1988 [98]. However, the electrode delivered poor electrochemical performance, which cannot be used in a battery. Actually, the research of organosulfide battery mainly focused on polysulfides from the 1980s to 2015, such as naphtho[1,8-cd][1,2]dithiol and dibenzo[c,e][1,2]dithiin, which have not attracted full attention [99, 100]. The polysulfides have a large specific capacity (over 300 mAh g−1), however, with poor stability due to the rigid framework in which the S—S bonds suffer from breaking and painful bonding, deteriorating the original structure which can be only used for lithium primary battery. Afterward, organosulfide with small molecule was investigated. However, the materials were considered hopeless because they easily dissolved into the electrolyte, leading to a shuttle effect that normally appeared in Li–S batteries. Until 2016, dimethyl trisulfide (DMTS) [101] was successfully applied in organic batteries with a reversible charge/discharge process (with 849 mAh g−1) and cycling performance (50 cycles) with the assistance of a carbon paper which was also proposed for polysulfide in 2013 [102]. The creative idea has opened a broad perspective for the research of small-molecule organosulfur [103–111]. Hereafter, molecules with more sulfur content (–Sn–) were studied to modify the specific capacity [112–116]. The performance of organosulfur can be regulated by heteroatom doping (such as Se, Te), combination with metal sulfides for good conductivity and stability, and the application of RFBs [117–126].

1.3 Goal, Scope, and Organization of this Book

It is obvious that a consequent and growing amount of literature is now easily available on organic batteries after years of silence. There is room for reversible electroactive organic systems in the future EES landscape in view of the application. However, it must be noted that there exists a certain disciplinary boundary between inorganic and organic compounds because the redox chemistry of organics is different from that of typical LIBs, making it challenging for nonspecialist readers when dealing with organic batteries. Therefore, it would be timely to provide a kind of “tutorial”-oriented book for a broader audience. Based on the latest selected and reliable data from both general and specialized scientific literature, this contribution also aims at providing the readers with a better understanding of the consecutive global demand for electrical energy sources and the evolution trends of organic batteries.

The following approach will be stepwise introduced in this book.

1.3.1 Working Principles and Fundamental Properties

First, the reader will have a concept of some prototypical organic materials, including the mechanisms and cell configurations. The organic electrodes as cathodes are the key point in this book, including the introduction of the mechanism (Chapter 2), carbonyl-based organic cathodes (Chapter 3), sulfur-containing organic cathodes (Chapter 4), radical-based organic cathodes (Chapter 5), organometallic complex-based cathodes (Chapter 6), polymer-based organic cathodes (Chapter 7), and other organic cathodes. The reader will know the basic types and reaction mechanisms of these organic electrodes. Afterward, some typical organic anodes are introduced (Chapter 8), which can be assembled as an all-organic battery counter to an organic cathode (Chapter 9).

1.3.2 A Selection of an Organic Electrode

After our introduction, the reader will know that the function of an organic electrode could be modified by applying a wealth of functional groups. Compared with the typical inorganic electrodes, organic electrodes can be considered cathodes (even working at high potential like P-type electrodes), anodes, and soluble agents for RFBs. Moreover, with the assistance of some special functional groups (such as N-based group), a battery with ultra-high-rate performance (e.g. thiuram monosulfide [TM]) and eutectic solution at low temperature as a special electrolyte (such as the interaction between 2,2′-dipyridyl disulfide (DPyDS) and lithium bis(trifluoromethanesulfonyl)imide [LiTFSI]) can be achieved [107, 127]. The organic materials hold a wide range of voltage, specific capacity, and energy density (Figure 1.2). Thus, the reader will know how to choose an organic electrode based on the demand.

Figure 1.2 Typical organic materials with voltage, specific capacity, and energy density are compared with some typical traditional electrode materials.

Note that TTF-based, DPDS, PMTT, PTS, DMTS, PPS, NTCDA, PQL, and PDA denote tetrakis(methylthio)-derivative cyclohexene-1,4-diylidenes (TTF-based) [128], diphenyl disulfide (DPDS) [119], dipentamethylenethiuram tetrasulfide (PMTT) [129], phenyl tetrasulfide (PTS) [118], dimethyltrisulfide (DMTS) [101], phenyl hexasulfide (PPS) [92], 1,4,5,8-naphthalenetetracarboxylic dianhydride (NTCDA) [95], poly(1,6-dihydropyrazino[2,3g]quinoxaline-2,3,8-triyl-7-(2H)-ylidene-7,8-dimethylidene) (PQL) [130], and polydopamine (PDA) [131], respectively.

1.3.3 EES Applications

After the introduction, the reader will understand the concept of the organic electrodes in terms of the cost, resource availability, and stability. For the cost, the typical DMTS and TM have an obvious price advantage compared with LiCoO2 and NCM. Notably, TM has an ultrahigh redox activity with long cycling performances (over 8000 cycles), making it a potential candidate [127]. As for availability, organic materials are based on plentiful elements such as C, N, O, and S. It is obvious from the price comparison that the organic electrodes have rich resources. Herein, taking the long view, the organic electrodes have a low cost. As for chemical stability, the typical inorganic electrodes always suffer from a phase transition during cycling; hereafter, the structure deteriorates until the end of the cycle life. By contrast, the solubility property of organic electrodes makes the electron transfer process faster through a new mechanism different from the insertion/extraction mechanism. Note that a moderate dissolution of the active material benefits the rate performance and mitigates the evolution of the bulk structure of the electrode.

1.3.4 Practical Applications

Although research on the organic electrodes is earlier than the typical LIBs, the development of LIBs is faster due to their successful commercialization by Sony Corp. LIBs have the advantages of high specific capacity and stable cycling lifespan. However, LIBs encounter the bottleneck at the point of long development. Herein, other types of batteries emerge as alternatives, such as SIBs and Zn-ion batteries. As a prototype post-LIB, the battery based on organic materials holds obvious advantages. On the one hand, a P-type organic electrode (e.g. ethyl viologen iodide) can deliver a capacity of 230 mAh g−1, with a high output potential of 3.7 V. On the other hand, the TM electrode can hold a robust cycling lifespan, with a high capacity retention of 70% even after 8000 cycles. As for security, the electrolyte based on DPyDS and LiTFSI consists of a eutectic solution just by grinding together, being of incombustibility, which potentially matches the demands of safety for electrolytes. Moreover, RFBs can be designed based on organic electrodes applying their solubility. Therefore, it should be declared that organic electrodes have a high potential for practical application.

1.3.5 Key Challenges

The solubility of the organic electrodes is one key challenge. Because the soluble organic material has a shuttle effect, the organic electrodes keep the silence for a long time. With the application of carbon paper for organosulfide in 2016, a reversible and long cycling performance of organic battery was achieved. A carbon paper has strong adsorption ability which restricts the dissolution of the organic electrodes, mitigating the shuttle effect to some extent. Recently, a functional MOF-based separator has been designed to hinder the shuttle of the active material for LIBs based on the organic electrode (5,5′-dimethyl-2,2′-

bis

-

p

-benzoquinone), prolonging the cycle life to 2000 cycles (capacity retention of 82.9%)

[132]

. However, the mentioned separator raises the cost. From the view of practical use, a more facile and low-cost method should be applied.

The cycling stability is another challenge for the application of organic electrodes. Although the soluble property accelerates the reaction kinetics, it also leads to the loss of the active material, resulting in a capacity loss and a short cycle life. The dissolution degree of organic electrodes should be carefully regulated. Moreover, some organic electrodes have the intrinsic properties of electrochemical inertness, resulting in a large polarization, which should be a concern.

Various mechanisms are involved in the organic reactions, which are different from the traditional mechanisms in LIBs. The reaction always involves the break and recombination of the active bond, free radical reaction, charge recombination, and so on. Herein, the detailed reactions step by step are still indistinct due to the complex electron transfer, dissolution–diffusion, element, and steric-dependent properties involved in the processes. The details also depend on the advanced equipment.

As discussed above, some organic electrodes have their advantages, such as long lifespan, high working potential, and robust stability. However, it is still hard to find one organic electrode that has its full advantages. The organic electrode should be comparable in the fields of energy density, lifespan, and cost.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant Nos. U2004214 and 21975225) and the National Key Research and Development Program of China (2022YFB2402200).

References

1

Li, M., Lu, J., Chen, Z., and Amine, K. (2018).

Adv. Mater.

30: 1800561.

2

Mizushima, K., Jones, P.C., Wiseman, P.J., and Goodenough, J.B. (1980).

Mater. Res. Bull.

15: 783–789.

3

Yazami, R. and Touzain, P. (1983).

J. Power Sources

9: 365–371.

4

Padhi, A.K., Nanjundaswamy, K.S., and Goodenough, J.B. (1997).

J. Electrochem. Soc.

144: 1188.

5

Choi, J. and Manthiram, A. (2005).

J. Electrochem. Soc.

152: A1714.

6

Jiang, J., Shi, W., Zheng, J. et al. (2014).

J. Electrochem. Soc.

161: A336.

7

Kang, B. and Ceder, G. (2009).

Nature

458: 190–193.

8

Kim, U.-H., Park, G.-T., Son, B.-K. et al. (2020).

Nat. Energy

5: 860–869.

9

Li, W., Erickson, E.M., and Manthiram, A. (2020).

Nat. Energy

5: 26–34.

10

Logan, E.R., Hebecker, H., Eldesoky, A. et al. (2020).

J. Electrochem. Soc.

167: 130543.

11

Sun, H.H., Ryu, H.-H., Kim, U.-H. et al. (2020).

ACS Energy Lett.

5: 1136–1146.

12

Whittingham, M.S. (1976).

J. Electrochem. Soc.

123: 315.

13

Liu, Y., Lin, D., Liang, Z. et al. (2016).

Nat. Commun.

7: 10992.

14

Ye, H., Xin, S., Yin, Y.-X. et al. (2017).

J. Am. Chem. Soc.

139: 5916–5922.

15

Thery, J. and Briancon, D. (1964).

Rev. Int. Hautes Temp. Refract.

1: 221.

16

Delmas, C. (2018).

Adv. Energy Mater.

8: 1703137.

17

Delmas, C., Fouassier, C., and Hagenmuller, P. (1980).

Phys. B+C

99: 81–85.

18

Parant, J.-P., Olazcuaga, R., Devalette, M. et al. (1971).

J. Solid State Chem.

3: 1–11.

19

Delmas, C., Braconnier, J.-J., Fouassier, C., and Hagenmuller, P. (1981).

Solid State Ionics

3–4: 165–169.

20

Braconnier, J.-J., Delmas, C., Fouassier, C., and Hagenmuller, P. (1980).

Mater. Res. Bull.

15: 1797–1804.

21

Li, Y., Zhou, Q., Weng, S. et al. (2022).

Nat. Energy

7: 511–519.

22

Qiao, Y., Yang, H., Chang, Z. et al. (2021).

Nat. Energy

6: 653–662.

23

Yue, J., Zhang, J., Tong, Y. et al. (2021).

Nat. Chem.

13: 1061–1069.

24

Suo, L., Borodin, O., Gao, T. et al. (2015).

Science

350: 938–943.

25

Sato, K., Ichinoi, R., Mizukami, R. et al. (2018).

J. Am. Chem. Soc.

140: 1049–1056.

26

Kim, D.J., Yoo, D.-J., Otley, M.T. et al. (2019).

Nat. Energy

4: 51–59.

27

Han, C., Li, H., Shi, R. et al. (2019).

J. Mater. Chem. A

7: 23378–23415.

28

Armand, M. and Tarascon, J.M. (2008).

Nature

451: 652–657.

29

Friebe, C., Lex-Balducci, A., and Schubert, U.S. (2019).

ChemSusChem

12: 4093–4115.

30

Wang, S., Li, F., Easley, A.D., and Lutkenhaus, J.L. (2019).

Nat. Mater.

18: 69–75.

31

Xie, J. and Zhang, Q. (2019).

Small

15: 1805061.

32

Wang, D.-Y., Guo, W., and Fu, Y. (2019).

Acc. Chem. Res.

52: 2290–2300.

33

Wang, D.-Y., Liu, R., Guo, W. et al. (2021).

Coord. Chem. Rev.

429: 213650.

34

Wang, D.-Y., Si, Y., Guo, W., and Fu, Y. (2020).

Adv. Sci.

7: 1902646.

35

Wang, D.-Y., Si, Y., Guo, W., and Fu, Y. (2021).

Nat. Commun.

12: 3220.

36

Wang, D.-Y., Si, Y., Li, J., and Fu, Y. (2019).

J. Mater. Chem. A

7: 7423–7429.

37

Wang, D.-Y., Wang, W., Li, F. et al. (2022).

J. Energy Chem.

71: 572–579.

38

Alotto, P., Guarnieri, M., and Moro, F. (2014).

Renewable Sustainable Energy Rev.

29: 325–335.

39

Sum, E. and Skyllas-Kazacos, M. (1985).

J. Power Sources

15: 179–190.

40

Sum, E., Rychcik, M., and Skyllas-kazacos, M. (1985).

J. Power Sources

16: 85–95.

41

Rychcik, M. and Skyllas-Kazacos, M. (1988).

J. Power Sources

22: 59–67.

42

Xu, Y., Wen, Y., Cheng, J. et al. (2009).

Electrochem. Commun.

11: 1422–1424.

43

Williams, D.L., Byrne, J.J., and Driscoll, J.S. (1969).

J. Electrochem. Soc.

116: 2.

44

Liang, Y., Tao, Z., and Chen, J. (2012).

Adv. Energy Mater.

2: 742–769.

45

Häupler, B., Wild, A., and Schubert, U.S. (2015).

Adv. Energy Mater.

5: 1402034.

46

Janoschka, T., Hager, M.D., and Schubert, U.S. (2012).

Adv. Mater.

24: 6397–6409.

47

Lee, S., Kwon, G., Ku, K. et al. (2018).

Adv. Mater.

30: 1704682.

48

Leung, P., Shah, A.A., Sanz, L. et al. (2017).

J. Power Sources

360: 243–283.

49

Luo, J., Hu, B., Hu, M. et al. (2019).

ACS Energy Lett.

4: 2220–2240.

50

Muench, S., Wild, A., Friebe, C. et al. (2016).

Chem. Rev.

116: 9438–9484.

51

Peng, H., Yu, Q., Wang, S. et al. (2019).

Adv. Sci.

6: 1900431.

52

Schon, T.B., McAllister, B.T., Li, P.-F., and Seferos, D.S. (2016).

Chem. Soc. Rev.

45: 6345–6404.

53

Song, Z. and Zhou, H. (2013).

Energy Environ. Sci.

6: 2280–2301.

54

Wang, H. and Zhang, X. (2018).

Chem. Eur. J

24: 18235–18245.

55

Wei, X., Pan, W., Duan, W. et al. (2017).

ACS Energy Lett.

2: 2187–2204.

56

Winsberg, J., Hagemann, T., Janoschka, T. et al. (2017).

Angew. Chem. Int. Ed.

56: 686–711.

57

Wu, Y., Zeng, R., Nan, J. et al. (2017).

Adv. Energy Mater.

7: 1700278.

58

Xie, J., Gu, P., and Zhang, Q. (2017).

ACS Energy Lett.

2: 1985–1996.

59

Xu, Y., Zhou, M., and Lei, Y. (2018).

Mater. Today

21: 60–78.

60

Zhao, Q., Lu, Y., and Chen, J. (2017).

Adv. Energy Mater.

7: 1601792.

61

Zhao, Q., Zhu, Z., and Chen, J. (2017).

Adv. Mater.

29: 1607007.

62

Zhu, L., Ding, G., Xie, L. et al. (2019).

Chem. Mater.

31: 8582–8612.

63

Zhu, Z. and Chen, J. (2015).

J. Electrochem. Soc.

162: A2393.

64

Armstrong, C.G. and Toghill, K.E. (2018).

Electrochem. Commun.

91: 19–24.

65

Casado, N., Hernández, G., Sardon, H., and Mecerreyes, D. (2016).

Prog. Polym. Sci.

52: 107–135.

66

Ding, Y., Zhang, C., Zhang, L. et al. (2018).

Chem. Soc. Rev.

47: 69–103.

67

Gong, K., Fang, Q., Gu, S. et al. (2015).

Energy Environ. Sci.

8: 3515–3530.

68

Gracia, R. and Mecerreyes, D. (2013).

Polym. Chem.

4: 2206–2214.

69

Oltean, V.A., Renault, S., Valvo, M.L., and Brandell, D. (2016).

Materials

9: 142.

70

Abel, P.R., Lin, Y.-M., de Souza, T. et al. (2013).

J. Phys. Chem. C

117: 18885–18890.

71

Gao, H., Zhou, T., Zheng, Y. et al. (2016).

Adv. Energy Mater.

6: 1601037.

72

Li, L., Seng, K.H., Li, D. et al. (2014).

Nano Res.

7: 1466–1476.

73

Li, X., Guo, S., Jiang, K. et al. (2018).

ACS Appl. Mater. Interfaces

10: 16–20.

74

Mao, J., Fan, X., Luo, C., and Wang, C. (2016).

ACS Appl. Mater. Interfaces

8: 7147–7155.

75

Wang, Y.-X., Seng, K.H., Chou, S.-L. et al. (2014).

Chem. Commun.

50: 10730–10733.

76

Wu, L., Hu, X., Qian, J. et al. (2014).

Energy Environ. Sci.

7: 323–328.

77

Xu, Y., Zhu, Y., Liu, Y., and Wang, C. (2013).

Adv. Energy Mater.

3: 128–133.

78

Zheng, Y., Zhou, T., Zhang, C. et al. (2016).

Angew. Chem. Int. Ed.

55: 3408–3413.

79

Zhou, T., Pang, W.K., Zhang, C. et al. (2014).

ACS Nano

8: 8323–8333.

80

Nakahara, K., Iwasa, S., Satoh, M. et al. (2002).

Chem. Phys. Lett.

359: 351–354.

81

Suga, T., Pu, Y.-J., Oyaizu, K., and Nishide, H. (2004).

Bull. Chem. Soc. J.

77: 2203–2204.

82

Liu, M., Visco, S.J., and De Jonghe, L.C. (1990).

J. Electrochem. Soc.

137: 750.

83

Ma, T., Liu, L., Wang, J. et al. (2020).

Angew. Chem. Int. Ed.

59: 11533–11539.

84

Shirakawa, H., Louis, E.J., MacDiarmid, A.G. et al. (1977).

J. Chem. Soc., Chem. Commun.

16: 578–580.

85

Chance, R.R., Shacklette, L.W., Miller, G.G. et al. (1980).

J. Chem. Soc., Chem. Commun.

8: 348–349.

86

Ivory, D.M., Miller, G.G., Sowa, J.M. et al. (1979).

J. Chem. Phy.

71: 1506–1507.

87

MacInnes, D., Druy, M.A., Nigrey, P.J. et al. (1981).

J. Chem. Soc., Chem. Commun.

7: 317–319.

88

Novák, P., Müller, K., Santhanam, K.S.V., and Haas, O. (1997).

Chem. Rev.

97: 207–282.

89

Jia, X., Ge, Y., Shao, L. et al. (2019).

ACS Sustainable Chem. Eng.

7: 14321–14340.

90

Lu, Y., Hou, X., Miao, L. et al. (2019).

Angew. Chem. Int. Ed.

58: 7020–7024.

91

Iordache, A., Delhorbe, V., Bardet, M. et al. (2016).

ACS Appl. Mater. Interfaces

8: 22762–22767.

92

Bhargav, A., Bell, M.E., Karty, J. et al. (2018).

ACS Appl. Mater. Interfaces

10: 21084–21090.

93

Hansen, K.-A., Nerkar, J., Thomas, K. et al. (2018).

ACS Appl. Mater. Interfaces

10: 7982–7988.

94

Matsunaga, T., Kubota, T., Sugimoto, T., and Satoh, M. (2011).

Chem. Lett.

40: 750–752.

95

Han, X., Qing, G., Sun, J., and Sun, T. (2012).

Angew. Chem. Int. Ed.

51: 5147–5151.

96

Renault, S., Oltean, V.A., Araujo, C.M. et al. (2016).

Chem. Mater.

28: 1920–1926.

97

Luo, C., Borodin, O., Ji, X. et al. (2018).

Proc. Natl. Acad. Sci. U.S.A.

115: 2004–2009.

98

Visco, S.J. and DeJonghe, L.C. (1988).

J. Electrochem. Soc.

135: 2905.

99

Liu, M., Visco, S.J., and De Jonghe, L.C. (1989).

J. Electrochem. Soc.

136: 2570.

100

Inamasu, T., Yoshitoku, D., Sumi-otorii, Y. et al. (2003).

J. Electrochem. Soc.

150: A128.

101

Wu, M., Cui, Y., Bhargav, A. et al. (2016).

Angew. Chem. Int. Ed.

55: 10027–10031.

102

Fu, Y., Su, Y.-S., and Manthiram, A. (2013).

Angew. Chem. Int. Ed.

52: 6930–6935.

103

Sun, J., Zhang, K., Fu, Y., and Guo, W. (2022).

Nano Res.

16: 3814–3822.

104

Guo, W., Zhang, W., Si, Y. et al. (2021).

Nat. Commun.

12: 3031.

105

Lian, J., Guo, W., and Fu, Y. (2021).

J. Am. Chem. Soc.

143: 11063–11071.

106

Gao, M., Lan, J., Fu, Y., and Guo, W. (2022).

ChemSusChem

15: e202200423.

107

Song, J., Si, Y., Guo, W. et al. (2021).

Angew. Chem. Int. Ed.

60: 9881–9885.

108

Tang, S., Chen, Q., Si, Y. et al. (2021).

Adv. Mater.

33: 2100824.

109

Tang, S., Guo, W., and Fu, Y. (2021).

Adv. Energy Mater.

11: 2000802.

110

Li, F., Si, Y., Liu, B. et al. (2019).

Adv. Funct. Mater.

29: 1902223.

111

Zhao, B., Si, Y., Guo, W., and Fu, Y. (2022).

Adv. Funct. Mater.

32: 2112225.

112

Li, F., Guo, W., Si, Y. et al. (2021).

Electrochim. Acta

370: 137757.

113

Zhao, J., Si, Y., Han, Z. et al. (2020).

Angew. Chem. Int. Ed.

59: 2654–2658.

114

Li, F., Si, Y., Li, Z. et al. (2020).

J. Mater. Chem. A

8: 87–90.

115

Sang, P., Si, Y., and Fu, Y. (2019).

Chem. Commun.

55: 4857–4860.

116

Wu, M., Bhargav, A., Cui, Y. et al. (2016).

ACS Energy Lett.

1: 1221–1226.

117

Chen, Q., Si, Y., Guo, W., and Fu, Y. (2022).

Chem. Commun.

58: 10993–10996.

118

Lv, X., Guo, W., Song, J., and Fu, Y. (2022).

Small

18: e2105071.

119

Ren, S., Si, Y., Guo, W., and Fu, Y. (2022).

ACS Sustainable Chem. Eng.

10: 7526–7535.

120

Xie, K., Sun, J., Lian, J. et al. (2022).

Appl. Phys. Lett.

121: 133904.

121

Cui, Y., Ackerson, J.D., Ma, Y. et al. (2018).

Adv. Funct. Mater.

28: 1801791.

122

Wang, Z., Li, X., Guo, W., and Fu, Y. (2021).

Adv. Funct. Mater.

31: 2009875.

123

Fan, Q., Guo, W., Si, Y. et al. (2021).

Adv. Funct. Mater.

31: 2001493.

124

Chen, Q., Wang, W., Li, X. et al. (2022).

Proc. Natl. Acad. Sci. U.S.A.

119: e2202449119.

125

Chen, Q., Guo, W., and Fu, Y. (2022).

Adv. Sci.

9: 2104036.

126

Chen, Q., Guo, W., Wang, D., and Fu, Y. (2021).

J. Mater. Chem. A

9: 12652–12658.

127

Chen, Q., Li, L., Wang, W. et al. (2022).

J. Am. Chem. Soc.

144: 18918–18926.

128

Kato, M., Senoo, K.-I., Yao, M., and Misaki, Y. (2014).

J. Mater. Chem. A

2: 6747–6754.

129

Bhargav, A., Ma, Y., Shashikala, K. et al. (2017).

J. Mater. Chem. A

5: 25005–25013.

130

Wu, J., Rui, X., Long, G. et al. (2015).

Angew. Chem. Int. Ed.

54: 7354–7358.

131

Sun, T., Li, Z.-J., Wang, H.-G. et al. (2016).

Angew. Chem. Int. Ed.

55: 10662–10666.

132

Bai, S., Kim, B., Kim, C. et al. (2021).

Nat. Nanotechnol.

16: 77–84.

2Redox Mechanisms and Characterization Methods of Organic Electrode Materials

2.1 Introduction

In general, organic electrode materials of rechargeable batteries are classified into n-type (e.g. carbonyls, organosulfides, azos, and imines), p-type (e.g. thioethers and phenothiazines), and bipolar-type (e.g. nitrogen radicals) according to their charge state changes of redox-active groups on the electrochemical reaction process [1, 2]. Figure 2.1 gives three types of redox mechanisms and some representative molecular structures of organic electrode materials. For n-type organic electrode materials which can change their state between a neutral state and a negatively charged state. Typically, n-type organic materials first undergo reduction from neutral states to form anions, and then compensate anions by cations, e.g., Li+, Na+, or K+, in the electrolyte. In turn, the anions experience oxidation in their original states. As the redox-active component for rechargeable batteries, n-type materials often display high specific capacity, high energy density, and low redox potential [3–5].

Conversely, p-type organic materials usually change their state between a neutral state and a positively charged state. They first lose electrons to form cations and then compensate cations via anions (e.g. TFSI− and ClO4−) from supporting electrolytes to maintain the charge neutral. Considering their high redox potentials and fast redox kinetics, p-type organic materials are usually used as the cathode component in rechargeable batteries [6, 7]. Nonetheless, the low specific capacity of p-type organic materials remains a formidable challenge [8, 9]. Bipolar-type materials can shift their state not only between a neutral state and a negatively charged state but also between a neutral state and a positively charged state. Therefore, they have characteristics of both n-type and p-type organic electrode materials.

In the following, redox mechanisms of carbonyls, organosulfides, radicals, and N-containing active materials are separately introduced in detail. Meanwhile, some typical advanced characterization techniques, including Fourier transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), Raman spectroscopy, ultraviolet–visible (UV–vis) spectroscopy, nuclear magnetic resonance (NMR), electron paramagnetic resonance (EPR) spectroscopy, and fluorescence spectroscopy, for verifying their redox mechanisms are selected and discussed, which give us a comprehensive understanding of the application of advanced characterization methods in organic rechargeable batteries and lead to building better organic rechargeable batteries.

Figure 2.1 Three types of reaction mechanisms and some representative molecular structures of organic electrode materials.

2.2 Carbonyl Materials

2.2.1 Redox Mechanisms

Carbonyl redox-active materials commonly undergo a n-type redox behavior, which includes one or more carbonyl units as redox centers to store/release electric charges [10]. Upon the electrochemical reduction, a C=O unit can accept one electron to form a monoanion of C–O− (Figure 2.2). Then, a cation, e.g., Li+ or K+, offered in the electrolyte combined with the C–O− to realize electrically neutral. This redox process is reversible in the presence of various appropriate organic functional groups. In addition, carbonyl redox-active materials are also the most widely studied in organic batteries, which have four categories, i.e., imides, quinones, anhydrides, and carboxylates till now [11–13]. Therein, quinone electrode materials consist of two adjacent carbonyl units directly connected to conjugated rings, which undergo well-known reversible two-electron redox reactions (Figure 2.3) that enable their high theoretical capacities, e.g., 1,4,5,8-naphthodiquinone affords a high theoretical value of 570 mAh g−1[14–18].

Figure 2.2 Redox behavior of a carbonyl unit, R represents one organic functional group.

Figure 2.3 Redox behaviors of four categories of carbonyl redox-active materials.

Imide electrode materials have imino moieties around the carbonyl groups. Theoretically, an imide molecule with four carbonyl units can undergo four electron transfers reversibly. However, the reversible reaction behavior only realizes the first two electron transfers [19]. If continue to accept two electrons, the charge repulsion will emerge that damages the imide structure irreversibly. Therefore, the practical capacities of imide redox-active materials are contributed by their first two electron transfers (Figure 2.3) [20]. For example, a lithiation pyromellitic diimide material undergoes two one-electron redox reactions in rechargeable lithium batteries [21]. Anhydride electrode materials carry four carbonyl units distributed around an aromatic core. They are able to undergo four electron transfers reversibly for a monomer molecule (Figure 2.3) [22–24].

Carboxylate materials have an aromatic ring connected with two/multiple metal carboxylates, which are capable of undergoing reversible two-electron conversion reactions (Figure 2.3) [4]. During the reduction, the generated dianion is stabilized via a conjugated system. The loss of resonance or aromaticity in the core enables them to have low redox potential, which is beneficial for them as the anode redox-active component for organic rechargeable batteries. For example, a sample of carboxylate material, i.e., di-lithium terephthalate, delivered a low cell voltage of 0.8 V in lithium batteries [4, 5].

2.2.2 Characterization Methods

FTIR is a powerful analytical tool to explore electrode reactions of carbonyl materials, which due to the C=O group have a strong stretching vibration peak at 1900∼1650 cm−1 of the spectrum [26–28]. For example, the reaction mechanism of 1,4-bis(p-benzoquinonyl) benzene (BBQB) in a rechargeable lithium-organic battery is analyzed by ex situ FTIR (Figure 2.4) [25]. The characteristic stretching vibration peak of C=O units in the pristine BBQB electrode appears at 1652 cm−1. Upon discharging to ∼2.6 V, the peak intensity becomes weak. It almost disappears at the end of discharge, proving that most C=O units in BBQB are reduced to –C–O−. In recharge, the characteristic peak at 1652 cm−1 shows again, suggesting that one BBQB molecule can accept four electrons and four Li+ reversibly.

Figure 2.4 (a) Multielectron redox process of the BBQB electrode in lithium batteries. (b) Discharge/charge profiles of a lithium-BBQB cell. (c) Ex situ FTIR spectra of BBQB at various states of discharge/charge as marked in (b).

Source: Reproduced with permission from Yang et al. [25]. Copyright 2020, Wiley-VCH.

The above example indicates that the ex situ FTIR spectra can effectively respond to the disappearance/appearance of the characteristic peak of C=O bonds during electrochemical oxidation/reduction. In spite of this, to gain insight into the reaction condition of electrodes at different states of charge-discharge by ex situ methods, assembling and disassembling multiple cells is always required, which may bring some inevitable experimental errors [30]. To avoid this vulnerability, in situ characterization methods emerge that only assemble a test cell to monitor electrode reactions in real time. Luo et al. used in situ FTIR spectra to explore the reaction process of poly(imide-benzoquinone)-graphene (PIBN-G) in lithium-organic batteries (Figure 2.5a)[29]. During cycling (Figure 2.5b), real-time monitoring of the PIBN-G electrode was carried via in situ FTIR spectra (Figure 2.5c). Before discharging, there are two characteristic signals of the PIBN-G electrode: one is ascribed to the =O units of imides and appears at 1720 cm−1, and another signal appears at 1570 cm−1 is the C=O units of quinones. During discharging, the characteristic peak of 1720 cm−1 gradually disappears first, and then the peak of 1570 cm−1 fades. This indicates the redox process of the PIBN-G electrode involves two steps.

Figure 2.5 (a) Electrochemical redox reactions of PIBN in lithium batteries. (b) Discharge–charge profile of the PIBN-G electrode. (c) In situ FTIR spectra collected in real time. (d) Ex situ 13C NMR spectra collected at different discharge–charge states, as marked (b).

Source: Reproduced with permission from Luo et al. [29]. Copyright 2018, Wiley-VCH.

Ex situ solid-state 13C NMR was also used to investigate the redox process of PIBN-G (Figure 2.5d) [29]. There are six signals of the pristine PIBN-G cathode. When discharging to stage II, as marked in Figure 2.5b, the signals of imide groups disappear and a new signal at 75.7 ppm (g) appears, suggesting that the C=O units of imides have been reduced to –C–O−. After fully discharging, another new signal at 92.0 ppm (e) emerges with the loss of the characteristic signals of quinones. In recharge, all signals return to the natural state. Therefore, the ex situ solid-state 13C NMR is also an effective tool for studying the redox reactions of carbonyl materials.

Organic electrode materials are also used as the redox-active component for redox flow batteries. The active species are stored in tanks and flow into an electrochemical cell for oxidation and reduction via a pump. To gain insight into the redox mechanisms of soluble organic electrode materials in flow batteries, researchers proposed in situ NMR and in situ EPR to probe changes in active species [31, 32]. Figure 2.6