You are on page 1of 17

REVIEWS

Organic batteries for a greener


rechargeable world
Jihyeon Kim1,6, Youngsu Kim1,6, Jaekyun Yoo1,6, Giyun Kwon1, Youngmin Ko1,2,5
and Kisuk Kang   1,2,3,4 ✉
The emergence of electric mobility has placed high demands on lithium-ion batteries, inevitably
requiring a substantial consumption of transition-metal resources. The use of this resource raises
concerns about the limited supply of transition metals along with the associated environmental
footprint. Organic rechargeable batteries, which are transition-metal-free, eco-friendly and
cost-effective, are promising alternatives to current lithium-ion batteries that could alleviate
these mounting concerns. In this Review, we present an overview of the efforts to implement
transition-metal-free organic materials as the redox-active component in diverse types of
organic rechargeable batteries. In addition, we critically evaluate the current status of organic
rechargeable batteries from a practical viewpoint and assess the feasibility of their use in various
energy-storage applications with respect to environmental and economic aspects. We believe
this Review provides a timely evaluation of organic rechargeable batteries from a real-world
perspective, and we hope it will spur more intensive efforts towards a greener energy future.

Global efforts to lessen our carbon footprint have electrochemical performances on par with those of con-
prompted a transition to renewable energy and the ventional transition-metal-based materials. Given these
increased adoption of electric mobility. Because recharge- potential merits, remarkable efforts have been made over
able batteries are a key enabler in these endeavours, a the past few decades to introduce these redox-active
1
Department of Materials
substantial rise in battery production is foreseeable in organic materials as the key electrode component6–11.
Science and Engineering,
Research Institute of the coming years. Conventional lithium-ion batteries More recently, extensive investigations have aimed to
Advanced Materials (RIAM), rely on transition-metal-oxide-based materials — such implement these redox-active organic electrodes in
Seoul National University, as cobalt and nickel oxides — for their positive elec- practical battery systems and obtain performance lev-
Seoul, Republic of Korea. trodes, as they offer high energy density and long cycle els comparable to those of transition-metal-containing
2
Institute of Engineering life. As the demands for lithium-ion batteries are pro- commercial batteries. Moreover, some general disadvan-
Research, College of
Engineering, Seoul National
jected to soar, these transition metals are likely to see tages of organic materials, such as their low electronic
University, Seoul, Republic heavy consumption, which presents a potential supply conductivity and high dissolution properties, can be
of Korea. problem. The availability of transition-metal resources effectively mitigated through advanced electrode design
3
Center for Nanoparticle is limited, resulting in high production costs, and their or converted into technical merits by revising the battery
Research, Institute of Basic large-scale use is neither sustainable nor environmen- configuration.
Science, Seoul National tally benign. Moreover, transition-metal-based materi- In this Review, we highlight the recent progress in
University, Seoul, Republic
of Korea.
als are currently responsible for nearly 30–50% of the organic rechargeable battery technologies, focusing
global warming impact from battery-manufacturing mainly on practical aspects. In addition to providing a
4
School of Chemical and
Bioengineering, College of processes1,2. Various alternative approaches have thus comparative assessment of the individual redox-active
Engineering, Seoul National been suggested to enable us to break away from con- materials 6,9, we propose a more pragmatic evalua-
University, Seoul, Republic ventional transition-metal-based active materials and tion considering the entire electrode composition as
of Korea. proceed towards sustainable battery chemistry3–5. a component for organic batteries. We then discuss
5
Present address: Lawrence Organic rechargeable batteries have emerged as a alternative platforms for redox-active materials in
Berkeley National Laboratory,
promising alternative for sustainable energy storage post-lithium-ion-battery systems, covering the inclusion
Berkeley, CA, USA.
as they exploit transition-metal-free active materials, of Earth-abundant metal ions and the development of
6
These authors contributed
equally: Jihyeon Kim,
namely redox-active organic materials mostly com- aqueous organic batteries and organic redox-flow batter-
Youngsu Kim, Jaekyun Yoo. prising earth-abundant carbon, oxygen, hydrogen ies. This Review aims to offer a comprehensive and prac-
✉e-mail: matlgen1@snu.ac.kr and nitrogen6–9. Fundamental research into the redox tical assessment of the current challenges and solutions,
https://doi.org/10.1038/ activity in organic compounds has led to the discovery thus aiding in the further development of sustainable
s41578-022-00478-1 of several important groups of materials that exhibit and affordable organic rechargeable batteries.

Nature Reviews | Materials

0123456789();:
Reviews

Assessment of redox-active organics GWP100 for each additional process. Even though com-
Although the use of transition-metal-free active mate- plex industrial organic chemicals generally require two
rials is attractive from a sustainability point of view, to four steps of additional chemical processes from the
the precise assessment of redox-active organic materi- precursors16,17, currently mass-produced simple organic
als is challenging compared with that of commercially compounds generate less than one-fifth of the GWP100
mass-produced transition-metal-based electrode mate- of transition-metal-containing cathode materials1,14,15.
rials, especially in terms of practical cost per perfor- This value suggests that the estimated GWP100 of most
mance and environmental merits. These issues stem redox-active organic materials still remains substantially
from lack of mass production and related economies lower than that of transition-metal-based materials, as
of scale. Nonetheless, rough estimations are possible, depicted in Fig. 1a. Moreover, this projection is conserv-
using some relevant indexes such as the current market ative, meaning that some redox-active organic materials
price of typical active materials and the global warm- requiring simpler synthetic conditions would have a far
ing potential (GWP). The GWP is an index referring to lower GWP100 value.
the amount of carbon dioxide (CO2), or gas equivalent Figure 1b illustrates the changes in prices of repre-
to CO2, generated per the production of one kilogram sentative commercial cathode materials over the past
of a specific product over a given time12,13. Figure 1a 3 years. The costs clearly fluctuate even over a short
compares the GWP100 values (GWP over 100 years) period, varying more than two times in price during
of representative cathode materials — here LiCoO2 this period, which closely correlates with the raw-metal
(LCO) and LiNi0.6Co0.2Mn0.2O2 (NCM622)1,14 — with costs such as that of cobalt (blue line in Fig. 1b). The
that of several mass-produced organic compounds — transition-metal-based cathode materials usually make
methanol (CH3OH), olefins (ethylene (C2H4), propylene up 40% of the total cell cost, and so their fluctuating
(C3H6)) and common aromatic compounds (benzene, price poses an unpredictable risk factor in the mass
toluene and mixed xylenes)15. The commercialized production of lithium-ion batteries18. More importantly,
transition-metal-containing cathode materials clearly the long-term supply of the major transition metals
have a substantial effect on greenhouse gas emissions, remains vulnerable and may not be able to cope with
generating nearly 20 kg of CO2 for every kilogram of cath- the rapid growth of the market. According to recent
ode production. Although the GWP100 values for pro- reports19, the lithium-ion battery market is expected to
spective redox-active organic materials are not known, reach a capacity of 3 TWh by 2030, which is approxi-
they can be estimated from the GWP100 values for mately 20 times greater than the current capacity (about
common precursors and the corresponding increase in 140 GWh in 2020). The 3-TWh capacity of lithium-ion

a b
80
Cobalt metal
LCO
20 70 NCM622
NCM523

60 Price of conventional electrode materials


GWP100 (kg CO2-eq kg–1)

15
50
Price (US$ kg–1)

40
10
30

20
5

10 Estimated price of redox-active organic materials

0 0
LCO NCM622 Methanol Olefins Common Redox-active July 2018 July 2019 July 2020 July 2021
aromatic materials
compounds (estimated) Date

Fig. 1 | Comparative analysis of environmental and economic merits of redox-active organic materials are synthesized in two to four steps from
redox-active organic materials. a | Comparison of the global warming simple organic precursors and that little additional energy is needed during
potential (GWP) between common transition-metal-based active materials the synthesis process. The error bar denotes the range of estimated GWP
(blue bars) and organic materials (green bars). GWP100 denotes the relevant values in accordance with expected two to four synthesis steps. b | Daily
global warming impact of the specific product compared with the impact price of cobalt metal (blue), LCO (green), NCM622 (yellow) and NCM523
of CO2 gas emissions over 100 years. The GWP100a method from the IPCC (red) over a recent 3-year period. Data sources: Shanghai metal market,
2007 (ref. 236) report is used for LiCoO2 (LCO), whereas the GWP100a 2021 (for LCO, NCM622 and NCM523); London metal exchange, 2021
method from the IPCC 2013 (ref.237) report is used for NCM622 and organic (for cobalt metal). The olefins are ethylene (C2H4) and propylene (C3H6); the
materials. We note that the life cycle assessment for the calculation of common aromatic compounds are benzene, toluene and mixed xylene.
GWP100 excludes the end-of-life stage. The GWP of redox-active organic NCM523 and NCM622 refer to LiNi0.5Co0.2Mn0.3O2 and LiNi0.6Mn0.2Co0.2O2,
materials (orange bar) is an estimated value with the assumption that respectively.

www.nature.com/natrevmats

0123456789();:
Reviews

Oxygen-containing motif Nitrogen-containing motif

–e,+A
– – – e–, + A– + e–, + M+
O O + A– N N + A– N C N C + M+
+ e–, – A– + e–, – A– – e–, – M+

Ether motif Tertiary amine motif Imine motif


O O
+ e–, + M+ + 2e–, + 2M+ + e–, + M+
C C
+ M+ N N N N + 2M+ C C N C C N + M+
– e–, – M+ – 2e–, – 2M+ – e–, – M+

Carbonyl motif Azo motif Nitrile motif

Sulfur-containing motif
S S
– e–, + A– + e–, + M+ + 2e–, + 2M+
S S C C
+ A – + M+ S S 2 S + 2M+
+ e–, – A– – e–, – M+ – 2e–, + 2M+

Thioether motif Thiocarbonyl motif Disulfide motif

Other types of redox motifs

+ e–, + M+ x + xe–, – xA– + ye–, + yM+ y


+ e–, – A–
N O + A– N O N O + M+ + xA– + yM+
n
– e–, + A– – e–, – M+ n – xe–, + xA– n – ye–, – yM+

Nitroxide radical motif Motifs in conducting polymer

M+ = Li+, Na+, K+, other cations


A– = ClO4–, PF6–, TFSI–, other anions

Fig. 2 | redox motifs in representative redox-active organic electrodes. Representative redox motifs that are
commonly observed in redox-active organic materials in the literature. The blue and red colours indicate p-type and
n-type motifs, respectively. During p-type and n-type redox reactions, charge compensation by anions and cations,
respectively, in the electrolyte is generally required.

batteries would require approximately 2.2 million tons typically undergo reduction from their neutral state,
of transition metal (either cobalt or nickel) based on the forming negatively charged molecular states that can be
current cathode chemistry. Considering that the annual reversed during oxidation. Conversely, the electrochem-
production of cobalt is at present 0.14 million tons and ical reaction of p-type organic materials involves oxida-
that of nickel is 2.5 million tons, the demand for transi- tion from their neutral state, forming positively charged
tion metals to achieve this 3-TWh capacity cannot eas- molecular states reversibly. Some organic compounds
ily be met. According to the United States Geological are known to undergo a bipolar-type reaction, exploiting
Survey (USGS), this demand will be a serious issue in redox moieties capable of both n-type and p-type behavi­
the near future because competition with other indus- ours. The basic redox nature of redox-active organic
tries is to be expected. In particular, major demands for materials depends on the redox motif, a functional
nickel from the competing stainless-steel manufactur- group in the molecule, which confers redox activity.
ing industry could substantially raise the price of nickel In the following, we briefly introduce the major redox
during battery-market growth. Unlike these projections motifs constituting representative redox-active organic
on transition-metal availability and associated costs materials, as depicted in Fig. 2.
(Fig. 1b), previous studies unequivocally suggest that the
cost of redox-active organic materials would be less than Oxygen-containing motifs
US$15 per kilogram when mass-produced through con- The most common redox motifs found in redox-active
ventional organic synthetic routes9,20. We note that this organic materials are oxygen-containing units, which
price was estimated in 2017. This analysis implies that include two major types of functional groups: car-
redox-active organic materials have a greater economic bonyl and ether groups. Since the first report on
potential than do cobalt-based (London metal exchange, carbonyl-based organic materials as an active material
2021) and other transition-metal-based materials in 1969 (ref.21), various carbonyl redox-active organic
(Shanghai metal market, 2021), for TWh-level battery materials have been investigated, including quinone
mass production. derivatives22–26, carboxylate salts27,28, imides29,30 and
anhydrides31. The carbonyl motif (C=O) in these mate-
Redox-active organic materials rials enables an n-type redox reaction, accompanied by
Redox-active organic materials are usually classified as the reversible formation of a radical monoanion (C–O−)
n-type, p-type or bipolar-type according to their capabil- through the enolization reaction. Quinone derivatives
ities to release electrons (oxidation) or receive electrons such as vitamin K, ubiquinone and plastoquinone have
(reduction) in their neutral state during the electro- been the most widely studied among carbonyls because
chemical reaction. n-type redox-active organic materials they are abundantly found in nature and play a key part

Nature Reviews | Materials

0123456789();:
Reviews

in biochemical electron transfer23,32,33. Organic materi- redox potential than the nitrile groups because of elec-
als with carbonyl motifs generally supply a discharge tron delocalization on the sulfonyl group attached next
potential of approximately 2.5 V (versus Li+/Li), which to the nitrogen centre55. Redox-active organic materi-
is relatively low for cathodes and too high for anode als with the sulfonamide motif are usually synthesized
applications. However, the carboxylate compounds, in a reduced form with a negative charge in its pris-
which have analogous C=O bonds in their molecular tine state (that is, bound with a counter-cation such
structures, can deliver a low potential (approximately 1 V as Li+ or Na+); thus, the initial electrochemical process
versus Li+/Li)27, comparable to that of commercialized involves the extraction of the counter-cations to allow
anode materials such as Li4Ti5O12. an n-type reaction.
The ether motif (C–O–C) promotes a high-potential
p-type redox reaction, upon which a radical cation is Sulfur-containing motifs
reversibly formed. Organic electrodes with ether motifs Sulfur-containing redox motifs have similar redox
often provide the highest redox potential (over 4 V ver- characteristics to oxygen-containing redox units, pos-
sus Li+/Li)34 because of their strong electronegativity. sibly because sulfur and oxygen are in the same group
Nonetheless, the exploitation of ether-based materials, in the periodic table. The representative thioether motif
such as phenoxazine derivatives, as high potential elec- (C–S–C), which consists of two single bonds between
trodes is usually hindered by the limited electrochemical sulfur and carbon atoms, typically triggers a p-type reac-
window of conventional electrolytes and the complica- tion with a redox mechanism similar to that of the ether
tions involved with anion incorporation during p-type motif 56–58. The thiocarbonyl motif (C=S) is comparable
redox reactions35,36. to the carbonyl motif. However, since sulfur is less elec-
tronegative than oxygen, it induces an increased redox
Nitrogen-containing motifs potential and enhanced electronic conductivity 59,60.
Nitrogen-containing redox motifs include the tertiary The disulfide motif (C–S–S–C), which contains a
amine-, azo-, imine-, and nitrile-redox centres (Fig. 2). single bond between two sulfur atoms, has a distinc-
The ability of nitrogen atoms to form three covalent tive n-type redox mechanism; the S–S bond is broken
bonds allows for much more diverse redox moieties after accepting two electrons, and each R–S− anion is
than in their oxygen-containing counterparts. Among stabilized by a counter-cation at a relatively low redox
these, the tertiary amine motif permits p-type reactions potential (approximately 2.2 V versus Li+/Li). Because
similar to the ether motif, thus allowing a high redox of the multi-electron redox per motif, these compounds
potential over 3 V (versus Li+/Li) in electrode materi- can display high specific capacities and have thus been
als based on dihydrophenazine37–40, triphenylamine41,42, intensively studied61–63. However, the intrinsically slug-
phenothiazine 35,36 and phenoxazine 36,43 derivatives. gish kinetics originating from S–S bond breakage or
In particular, dihydrophenazine-based electrodes are formation during reduction and oxidation, respectively,
capable of undergoing a two-electron redox reaction lead to a high overpotential64. The solubility of the small
per molecule, offering a high specific capacity and thus molecules generated by the bond-breakage reaction also
a specific energy comparable to that of commercial leads to premature capacity decay caused by the loss of
LiFePO4 (LFP) cathodes37. active materials65.
The redox motifs of azo compounds (N=N)44–46,
imines (C=N)38,47–51, nitriles (C≡N)34,52–54 and sulfo- Other types of redox motifs
namides (C–N(SO2)–)55 are based on double or triple Other types of redox motifs appear in a radical form
bonds between nitrogen and carbon and typically display or in conducting polymers. One of the most repre-
n-type activity through the change in the bond order sentative examples in this class is the 2,2,6,6-tetra
upon redox reaction. Given their relatively simple struc- methylpiperidin-1-yl)oxyl (TEMPO)-based compound
ture and ability to reversibly react with two alkali ions, containing a nitroxide radical motif (N–O·). They
the family of azo compounds features a high capacity44. are intrinsically bipolar in nature but are more often
Imine compounds are often found in biological employed as cathode materials selectively utilizing
charge-transfer systems such as in pteridine and pyrid- p-type reactions to exploit associated high redox poten-
inium derivatives. They generally consist of a hetero­ tials (approximately 3.6 V versus Li+/Li). TEMPO-based
aromatic ring, with nitrogen atoms on the conjugated compounds are usually reported in the form of polymers
structure, allowing effective simplification or modifica- owing to their highly soluble nature66–69 and can exhibit
tion of the molecular structure through various means, a high rate capability resulting from their self-exchange
including by cutting out the conjugated moiety and reaction, which is a fast intermolecular electron trans-
keeping only the nitrogen-containing conjugated ring50. fer process69. However, the large redox-inactive parts
For the nitrile redox motif, 7,7,8,8-tetracyanoquinod- in the polymer typically yield a low specific capacity.
imethane and its derivatives are the most studied because Similarly, conducting polymers based on polyacetylene70,
they possess the highest redox potential among the polythiophene71 or polyaniline72 have been studied as
n-type redox-active organic materials (2.4–3.2 V versus redox-active materials. Doping these polymer elec-
Li+/Li)34,52–54. The high redox potential of nitrile compounds trodes with counter-ions should enable redox reac-
is attributable to the strong electron-accepting character- tions, leading to theoretically high capacities. However,
istics originating from the four electron-withdrawing to achieve these predicted capacities, high doping
nitrile groups. The sulfonamide motif — the newest of levels are necessary, which limit the cycle stability because
this group to be discovered — displays an even higher of the occurrence of side reactions with the dopants61.

www.nature.com/natrevmats

0123456789();:
Reviews

Applications in rechargeable batteries Strategies to increase the specific energy of redox-active


To date, 56 redox-active organic materials have been organic materials. As shown in Fig. 4, the high spe-
reported to possess the aforementioned redox motifs cific capacities of organic electrodes generally come at
(Fig.  3) . Here, we refer to them as organics 1 to 56 . the cost of a low redox potential; thus, the increase in the
When incorporated in organic electrodes, these mate- specific energy is not substantial in most cases. Specific
rials deliver the most remarkable electrochemical per- capacities over 300 mAh g−1 are routinely obtainable for
formances of the redox-active organics. Regarding the some organic electrodes; however, their redox poten-
suitability of these candidates for practical electrodes, tial mostly lies substantially below 3 V (versus Li+/Li).
which consist of not only active material but several Accordingly, extensive efforts have been placed in iden-
other components, it should be noted that their per- tifying high-redox-potential organic electrodes55,59,60,82–86.
formance in the literature has been evaluated mostly in For example, a few p-type organic materials exhibit a
the framework of conventional transition-metal-based high redox potential over 3.5 V (versus Li +/Li) 9,61,
battery assessment. Moreover, comparative evaluations as indicated by the triangular data points in Fig. 4a.
are not straightforward given the different electrode Moreover, chemical modifications of n-type materi-
fabrication procedures employed for each reported als were performed to enhance the redox potentials to
material. These issues imply that the intrinsic capabil- values comparable to those of transition-metal-based
ities of redox-active organic materials might not have materials61. For organic 732,87, the redox potential of
been properly appraised. The following section dis- lithium (2,5-dilithium-oxy)-terephthalate could be ele-
cusses the distinctive characteristics of organic materi- vated by as much as 800 mV by including spectator ions
als (organics 1 to 45 in Fig. 3) with respect to the major with higher ionic potential than lithium, such as Ca2+,
battery performance metrics — specific energy, specific Ba2+ and Mg2+. A comprehensive overview of the volt-
power and cycle stability — comprehensively considering age engineering of organic electrode materials has been
their practical applicability. Approaches that exploit the provided in other review papers6,9,61,88,89.
unique characteristics of organic electrodes that could
aid them to surpass conventional lithium-ion battery Electrode-level specific energy assessment of organic
electrochemical performances are also introduced. redox-active materials. Since the discovery of organic
redox-active materials with enhanced energy density, their
Comparative analysis of specific energy use in practical battery systems has been arduous. When it
The specific energy is determined by the numerical comes to practical energy density, the weight and volume
product of the specific capacity and the redox potential of the redox-inactive components must also be consid-
of the active material. The specific capacity and the volt- ered. These components, such as the conducting agent
age (versus Li+/Li) of state-of-the-art organic electrode and binder, are needed for electrode fabrication to secure
materials as well as conventional transition-metal-based mechanical stability and electrical conductivity. Because
electrode materials are compared in Fig.  4a (and in of the insulating nature of organic materials, organic
Supplementary Fig. 1a, where references for each mate- electrodes usually require high conducting agent and
rial are noted). The presented values of the specific binder contents. According to our literature survey, most
capacities and voltages are practical values obtained organic electrodes reported thus far contain 30–70 wt%
from the discharge curves in the literature. We note of conducting agent to make up for the low electronic
that the specific capacities in Figs.  4 and 5 are esti- conductivity so that organic electrodes can achieve a spe-
mated on the basis of the weight of each material in the cific energy comparable to that of transition-metal-based
ready-to-charge state (that is, the discharged state) for electrodes. This is in contrast with commercial electrodes
fair comparison with those of transition-metal-based where these additives make up less than 5 wt% of the total
inorganic materials. Since the accompanying anions electrode90,91. This issue has become an impediment for
are not taken into account in the specific capacities of organic electrodes92. Although many previous reviews
the presented p-type materials, the deliverable capacity have evaluated the specific capacity, redox potential or
may vary depending on the electrolyte systems in some specific energy of organic electrodes, a critical evalua-
cases, such as in dual-ion systems. Because of their light tion of the electrode-level specific energy, covering all
constituent elements and multi-electron transfer capa- the electrode components, has rarely been performed.
bilities, some organic electrode materials are capable Therefore, we compare the specific energy of the reported
of delivering specific capacities that are two to three redox-active organic materials at the electrode and mate-
times greater than those of transition-metal-based elec- rial level, highlighting the general deviation between the
trode materials such as LCO73,74, NCM75,76 and LFP77. two (Fig. 4b and Supplementary Fig. 1b, where references
Electrodes based on rhodizonate (organic 1)28,30,78 or for each material are noted). The x axis is the specific
cyclohexanehexone (organic 2)79 derivatives allow four energy calculated based on the weight of the redox-active
or six electron transfers per molecule, respectively, material, as suggested in the literature, and the y axis
resulting in capacities of approximately 503 mAh g−1 is the specific energy calculated based on the weight
or 723 mAh g−1 and energy densities of 1,117 Wh kg−1 or of the entire electrode including the redox-active material,
1,229 Wh kg−1. This is a clear benefit of organic electrode conducting agent and binder. The dashed line in Fig. 4b
materials over conventional layered transition-metal indicates the point at which the material-level specific
oxide cathodes, which are rapidly reaching their the- energy coincides with the electrode-level specific energy.
oretical limit of 270 mAh g −1 owing to their heavy The analysis indicates that conventional transition-metal-
transition-metal-element content32,80,81. based electrodes deviate only slightly from this line.

Nature Reviews | Materials

0123456789();:
Reviews

O O O 2
N N O O
O O M O O Li S O O Li O
Na S S
N N M2
N N
O
O O M O O O S S S Na
Li O
O O N N S S
3 O 4 5
M = Li, Na, K 2 6 O Li O
1 O O Li N M = Mg, Ca, Ba, Mn
Li O
N O O 7
M O S S S
O S N
N N N O O
O O M S S 12
8 M = Li, Na O Li O O
Li O O
11 13
10
9 O O
O n
O O
O O N N O O
Li O O Li
N N
Li O O Li S
O N O O
O O N N O O
O O
14 15 17 18
16

O O
O O O O O
N N H
O N N N N
N N O
O N N
N 20 N N N
O O H
O O N N O n
19 21 22 N 23
O

N
O O HO OH O O O O O
O O S

N N O O
S
NH NH O
O O
27
n 29
28
n O O
24 25 30
O 26 n
O Li O Li O
O O HO n
N S
O O O O
O O S
O N n
N S N
O O
O Li O Li n 34 N HO N
O 31 O 33 36
32 O
O O O
S 35 37
O Li O
N O
S Li S O
I M O O M
N N N
O O N N N
O S Li N
O I M O O M
S N
O 41 O
43
n S Li
39 40 M = Li, Na
38 n O O 42
O
O
O O O Na O3S
O O
O O
H
N N N SO3 Na
O O O S
O
O O
44 O 45 n n 48 O O
46 47

O O
OH OH
M O HO N N OH
N O
S O N
HO N N OH
O O O M
49 n M = Li, Na, K OH OH O O
51
50
O O O O
O O 52
O O 4
S
N N
4Cl
S N N
O O n O O
O O 54 55 56
53

www.nature.com/natrevmats

0123456789();:
Reviews

◀ Fig. 3 | Molecular structures of representative redox-active organic materials. organic materials may be another strategy. The distinc-
Representative redox-active organic materials that exhibit high battery performances. tive properties of organic materials may offer unexplored
Their redox characteristics are usually determined from their redox motif. benefits in electrode manufacturing. For instance, some
molecules, such as organic 14, exhibit a strong π–π inter-
By contrast, most of the organic electrodes display con- action with graphitic carbon owing to their polycyclic
siderably lower electrode-level specific energy than would aromatic hydrocarbon-bearing molecular structure.
be expected from the material level. It is noteworthy that Molecular interactions with the carbon-based conduct-
organic 1 appears to retain a high specific energy at the ing agent lead to an electrically well connected electrode
electrode level. Nevertheless, this specific energy rapidly structure even with a small amount of carbon additive95.
degrades within a few cycles, implying that the content Moreover, carbon nanotube additives further facilitate
of the redox-inactive components (conducting agent or electrical connections between the organic particles
binder) is too low to enable the components to effectively in the electrode because they have higher aspect ratios
perform their intended roles24. than spherical carbon additives95. Through the combi-
nation of the abovementioned approaches, an electrode
Strategies to increase the specific energy of redox-active with 90 wt% of active materials was fabricated from
organic materials at the electrode level. Unlocking the the microporous redox-active organic polymer 15 and
potential high specific energy of organic materials at single-walled carbon nanotubes, delivering an enhanced
the electrode level is an imperative step for their prac- electrode-level specific energy96. Usually, fabricating a
tical applicability. Certain strategies can close the gap thick electrode using the conventional slurry-casting
and achieve a high electrode-level specific energy, as method results in cracking and delamination owing
schematically illustrated in Fig. 4c–e. For instance, the to the mechanical instability of the electrode. However,
formation of a charge-transfer complex (Fig. 4c), when in the interaction between carbon additives, such as car-
a particular intermolecular arrangement facilitating elec- bon nanotubes, and microporous polymer leads to well
tron conduction, may reduce the amount of conducting connected and high-mass-loading electrodes (Fig. 4e).
agent needed in the electrode owing to the increased elec- Further efforts should optimize organic electrode fabri-
tronic conductivity of the complex. The charge-transfer cation processes to bring electrode-level specific energy
complexes consisting of organics 3 and 10 (depicted on par with material-level specific energy (as represented
as 3•10 in ref.34, Fig. 4b) or organics 3 and 11 (depicted as by the dashed line in Fig. 4b).
3•11 in ref.52, Fig. 4b) exhibit an electronic conductivity
markedly enhanced beyond the intrinsic conductivities Comparative analysis of specific power
of individual organics 3, 10 or 11. Thus, the electrode High power capability is needed for batteries to generate a
comprising one of these complexes can be fabricated large amount of energy in a short time and to make the
with more than 90 wt% of active material, enhancing the charge process faster97–99. Although the most intuitive
electrode-level specific energy. This type of approach, if approach to realize high power is to employ electrode
further developed, would allow high electronic conduc- materials with fast ion- and electron-transfer capabili-
tivity while retaining the intrinsic redox activities from ties, it remains challenging for most organic electrodes
the two precursor organic materials. Similarly, connect- owing to their generally low electronic conductivity.
ing redox-active organic materials (organic 13) into a In Fig. 5a (and in Supplementary Fig. 2a, where refer-
covalent organic framework structure provided a high ences for each material are noted), we selected repre-
conductivity by π–π interactions through the eclipsed sentative high-power organic electrodes and plotted the
stacking structure of the framework45. Consequently, normal­ized rate capabilities in comparison with those of
even with a small amount of conducting agent (less than conventional inorganic electrodes.
20 wt%), a high specific energy at the electrode level can Our analysis indicates that radical polymer electrodes
be achieved, retaining nearly the same specific energy as such as the ones with (2,2,6,6-tetramethylpiperidin-1-yl)
at the material level. oxyl units (organic 35) exhibit a competitive rate capa-
Using a multifunctional redox-active conducting bility up to nearly 20C, which is comparable to that
agent with a high electronic conductivity93,94 (Fig. 4d) also of transition-metal-based materials67,100,101. Moreover,
achieves a high electrode-level specific energy. For exam- a radical polymer electrode based on organic 16 proved
ple, titanium disulfide (TiS2) not only displays a high to have a superior power capability by delivering 46%
electronic conductivity of 74.4 S cm−1 but can addition- of the theoretical capacity within less than 1 min of
ally contribute to the specific capacity through reversible charging and discharging (100C)35, as indicated by the
intercalation. When the redox-inactive conducting agent circular data points in Fig. 5a. This power capability is
in an electrode fabricated from organic 13 is partially attribu­table to the high electronic conductivity of the
replaced with TiS2 (ref.94), the resulting hybrid electrode radical polymers stemming from the electron-hopping
can achieve a specific energy of 286 Wh per kilogram mechanism through electron self-exchange69,102–104.
of electrode, outperforming the electrode made only
with the conventional conducting agent (167 Wh per Strategies to increase the rate capability of redox-active
kilogram of electrode). organic materials. To improve performance, a charge-
Considering that the conventional electrode fabrica- transfer complex composed of two redox-active organic
tion process has been optimized for commercial electrode materials, 3 and 25, (3•25 in Fig. 5a) was produced and
materials (Fig. 4e), engineering new electrode configura- operated at 965 mA g−1, showing a remark­able capac-
tions that take advantage of the distinctive properties of ity utilization corresponding to 76% retention of the

Nature Reviews | Materials

0123456789();:
Reviews

capacity delivered at 48 mA g−1 (ref.34). This performance 3•25 charge-transfer complex was increased by more
stems from the enhanced electronic conductivity of the than 10,000 times compared with that of the pristine
charge-transfer complex, where the charge transport constituting materials (3 and 25)34,52.
is facilitated in the rearranged molecular structure34,52, A sulfur-substituted dicarboxylate compound such as
as discussed earlier. The electronic conductivity of the organic 5 can also deliver a respectable specific power

a 34 LCOb n-type redox-active organics


4 LCOa p-type redox-active organics
20 NCMa Transition-metal-based inorganics
16 21 LFP NCMb
23
7 9
3⋅11 3a
3⋅25 24
Voltage (V versus Li+/Li)

3 3⋅10
3b
6 29 33
28 32 30a 31 19a
13a 1c 17 22 1b Energy density (Wh kg–1)
2 15 4 18a
26a 1,500
19b 2
12 1a
14 1,000
8b 5
1 700
8a
500 27
200
0
0 100 200 300 400 500 600 700 800 1,200 1,300
Specific capacity (mAh g–1)
b1,000 c Molecular rearrangement
n-type redox-active organics
p-type redox-active organics 1b Single redox-active Organic
Transition-metal-based inorganics organic electrode complex
Electron donor
Electron acceptor
800 Organic complex
Low intrinsic High intrinsic Conducting agent
NCMa
Specific energy of electrode (Wh kg–1)

electronic electronic
conductivity conductivity
LCOa
d Hybrid electrode
2
27 Conventional Hybrid
600
31 electrode electrode
LCOb
NCMb Redox-active
19a organic materials
3⋅11 3a
24 Conducting agent
22 Conductive inorganic
400 23 32 33 Low density High density cathode materials
LFP Low capacity High capacity
19b
1a 9
13a 5 e New electrode engineering
15 26a Redox-active
3⋅10 21 30a Conventional Buckypaper organic materials
1 3⋅25 slurry-casting approach
14 7 4 c 18a Single-walled
200 17 Low mass High mass carbon nanotube
12 6 20 29 loading loading
28 Reduced graphene
16 8 34 oxide
b
8a 3b
Current
collector Electrically connected network:
0 high capacity, high rate capability
0 200 400 600 800 1,000 1,200 1,400
Specific energy of material (Wh kg–1)

Fig. 4 | Comparative analysis of specific energy of representative redox-active organic materials. a | Specific capacity,
voltage and corresponding specific energy (dashed line) of representative redox-active organic materials. The specific
capacity and voltage are the values from the discharge process. In particular, the discharge voltage was obtained by dividing
the capacity from the integrated value of the discharge curves. b | Correlation between the specific energy at the material
and electrode level of selected redox-active organic materials. The dashed line indicates the point at which the specific
energy at the material level coincides with that at the electrode level. c–e | Approaches that can lead to high-specific-energy
redox-active organic materials at the electrode level (grey shading in Fig. 4b). The formation of organic complexes, such as
charge-transfer complexes, by molecular rearrangement can enhance the electronic conductivity of organic compounds,
leading to an increase of the active material content (part c). By replacing the conducting agent with inorganic materials
with reversible redox activity as well as high conductivity, both the gravimetric and volumetric energy density of
the electrode can be improved (part d). A novel electrode engineering method exploiting the advantages of organic
materials, combined with the use of carbon nanotubes, allows slurry casting to be bypassed, enabling the fabrication
of a high-mass-loaded electrode with high energy density (part e). LCO, LiCoO2; LFP, LiFePO4.

www.nature.com/natrevmats

0123456789();:
Reviews

a b 300
3⋅25 36
100 LCOa Radical
Charge-transfer complex
NCMb Sulfur-substituted dicarboxylate compound
Capacity retention (%)

80 Morphology tailoring
35b
35a Nanohybrid with carbon
250
Transition-metal-based inorganics
60
LFP
Radical
40 Charge-transfer complex 5

Specific capacity of electrode (mAh g–1)


16
Sulfur-substituted
dicarboxylate compound 200
20 Morphology tailoring
Nanohybrid with carbon 6
LCOb NCMb
Transition-metal-based inorganics
0
0.1 1 10 100 32
C-rate 150
c
100 41 LFP 12
40 37 38a 16
7 18a 16
13a
13b 35b NCMb 24 23
28 14 33 17
31 LFP
Capacity retention (%)

80 LCOb 39 45 NCMa
LCOa 100
43 3 44
37
3⋅10 30b 42 18b 17
20 21 26a 28
6
30a 19a
60
1b 6 22 35a
15
19b 50 3⋅10
36
40 Non-treated organics 16
Polymerization Separator engineering
Nanohybrid with carbon Multiple engineering 35b
Electrolyte engineering Transition-metal-based inorganics
20 0
10 100 1,000 10,000 0.1 1 10 100
Cycle number C-rate

Fig. 5 | Comparative analysis of rate capability and cycle performance. a | Capacity retention of selected redox-active
organic materials, where representative strategies were applied to improve the rate capability with increasing current
rate (blue data points represent redox-active organic materials, red data points represent transition-metal-based
materials). Each curve shows how much capacity was maintained compared with that at the lowest current in each
report. A gentler slope indicates better rate capability, the redox-active organic materials performing just as well as the
transition-metal-based electrodes. b | Reversible capacity of electrodes containing redox-active organic materials as a
function of increasing current rate. c | Capacity retention after the number of charge/discharge cycles reported in each
paper. Multiple engineering signifies redox-active organic materials to which two or more engineering strategies were
applied. All the capacities are the discharge capacities in a Li half-cell. LCO, LiCoO2; LFP, LiFePO4.

of 909 W per kilogram of material with a capacity of in an increase of the contact area with the conducting
176 mAh g−1 at 4.2C. Although this performance is not agent. This enhanced contact area enables efficient
remarkable, it represents a substantial enhancement over uptake of electrons from the conducting agent as well
that of the pristine dicarboxylate motif of 225 W per kilo- as a decrease of the ion-diffusion path, thereby success-
gram of material. It has been speculated that replacing fully enhancing the rate capability105,106, as illustrated by
oxygen with sulfur in the dicarboxylate-based motif could the triangular data points in Fig. 5a. Moreover, hybrid
increase the electronic conductivity to about 10−5 S cm−1 structures employing advanced carbon materials can
because of the higher electron polarizability of sulfur59,60, achieve high power by exploiting the exceptionally high
enhancing power capability by a factor of almost ten. conductivity (higher than 104 S cm−1) of graphene or car-
Although most redox-active organic materials exhibit bon nanotubes107,108. Through the path connected by the
low electronic conductivity (less than 10−10 S cm−1, interaction between the π systems in the organic skel-
Supplementary Table 1) relative to that of transition- eton and the carbon material, the electronic transport
metal-based materials (approximately 10−5 S cm−1), to and from the redox-active organic material can be
structural or chemical modifications can result in tailored, resulting in improved rate capability. Using this
higher power capability. The rate capabilities of com- strategy, various organic electrodes including electrodes
mon redox-active organic materials for which the spe- with organics 17, 28, 32, 35 and 37 could successfully
cific power was enhanced through various engineering deliver a respectable specific energy that is comparable
approaches are displayed in Fig.  5a. Morphological with inorganic electrodes even at a charge/discharge rate
control of organic particles has been demonstrated of 10C or higher48,109–112.
to be effective. For example, the modification of the
morphology to a sheet-type rather than a sphere-type Electrode-level specific power assessment of redox-active
morphology with organic 32 or the reduction of the organic materials. The power capability has been
particle size to the nanoscale with organic 36 results improved substantially using the approaches listed

Nature Reviews | Materials

0123456789();:
Reviews

in Fig. 5a; however, organic electrodes still require a large poor cycling performance in its molecular form122,
amount of conducting agent, leading to reduced specific was tailored into a polymer form119,123–127 (organic 38).
power and energy at the electrode level. Figure 5b (and Further inclusion of advanced carbon material in the
Supplementary Fig. 2b, where references for each mate- electrode led to a capacity retention of about 57% after
rial are noted) present the rate capability of the afore- 9,000 cycles (organic 15)128. Although polymerization
mentioned redox-active organic materials based on improves the cycle life of organic electrodes, there are
electrode-level capacity. Even though radical polymers clear trade-offs; the polymerization results in poor rate
exhibit power capabilities that are competitive with those capability, stemming from the low electronic conductivi-
of inorganic materials (Fig. 5a), practical rate capabilities, ties of common polymers, and reduced capacity because
considering the weight of the entire electrode, are far of the addition of redox-inactive backbones61. To address
from what is delivered at the material level. Fabricating these trade-offs, a few new polymers are noteworthy. For
electrodes with a large amount of advanced carbon mate- instance, a ladder-type polymer (organic 22), synthe-
rial may not be an optimal strategy for high-performance sized with a configuration minimizing redox-inactive
organic rechargeable batteries because of the associated components, achieves a high capacity along with long
reduction in the volumetric and gravimetric energy and cycle life120. A donor–acceptor polymer (organic 39)
the power density. The organic battery field still awaits a that directly links the electron-donating polymer back-
solution that would enable high power density without bones with electron-accepting pendants129 simultane-
sacrificing the energy density, possibly by combining ously achieves high cycle stability and rate capability.
material- and electrode-level strategies. The use of advanced carbon materials is another way to
secure cycle stability, as they stably anchor redox-active
Comparative analysis of cycle life organic materials against dissolution through π–π inter-
The cycle life of an organic rechargeable battery is actions. These approaches were effective for redox-active
governed mainly by common factors, such as the age- compounds including organic 3•10, 14, 15, 17, 28, 35
ing and degradation of the electrolyte and electrodes and 37 (refs.34,48,95,109–111,128).
because of repetitive electrochemical reactions, which The dissolution issue of organic electrodes can
also contribute to the cycle stability of conventional also be tackled by engineering the electrolyte proper-
lithium-ion batteries113. However, several attributes of ties. Because the free solvents in the electrolyte trigger
organic materials are particularly critical for the cycle the solvation and dissociation of organic compounds,
stability of organic rechargeable batteries, such as attempts to reduce or eliminate free solvents in the elec-
the structural integrity of the organic crystal6. Unlike trolyte have been sought after. One way to reduce the
transition-metal-based electrode materials, organic effect of free solvents is to adopt high-concentration
crystals are held together by weak van der Waals forces electrolytes 130–132. Because the number of free sol-
between molecules. Accordingly, redox-active organic vent molecules greatly decreases with the change of
materials are often vulnerable to dissolution in the liquid the solvation structure in high-concentration elec-
electrolyte and can fall from the electrode through solva- trolytes, the dissolution of the organic electrodes is
tion by the solvent molecules in the electrolyte. The con- subsequently obstructed. Small redox-active organic
tinuous loss of redox-active organic materials from the materials such as organics 13, 20 and 40 are able to
electrode leads to a gradual decrease of capacity upon retain stable cycling performance without any molec-
cycling, which is a key obstacle limiting their practical ular engineering because of the suppressed dissolution
use. Moreover, the organic crystal structure becomes in the high-concentration electrolyte, as depicted in
easily disordered and degraded with repeated guest-ion refs.36,55,94, Fig. 5c. The use of solvent-free electrolytes,
insertion and extraction, causing undesirable changes in such as ionic liquids, is also a promising route to sup-
the electrochemical activity40,114. press the dissolution22,133–136. An electrode made from
organic 18 (calix[6]quinone) could exhibit a capacity
Strategies to increase the cycle life of redox-active retention of 69% after 1,000 cycles with the ionic liquid
organic materials. Many strategies have addressed these N-methyl-N-propylpyrrolidinium bis(trifluorometh-
issues and improved the cycle stability. In Fig. 5c (and in anesulfonyl)amide134. This performance represents
Supplementary Fig. 2c, where references for each mate- an important enhancement over the conventional
rial are noted), we compare and summarize the cycle per- electrolyte system (ethylene carbonate/dimethyl car-
formances of redox-active organic electrodes obtained bonate containing 1 M lithium hexafluorophosphate),
through various strategies with respect to the reported which delivers only 53.7% of the initial capacity after
cycle number and the capacity retention rate. One of the 100 cycles137. Similarly, polymer- or plastic-crystal-type
most intuitive examples includes dimer-/trimer- and/or electrolytes could support the cycle stability of
polymer-based organic species35,38,39,45,110,115–121 (denoted redox-active compounds such as organics 18, 19, 37
with pentagonal data points), such as organic 31 (ref.116) and 41 (refs.109,133,138,139). The low molecular mobility
and organic polymer 16 (ref.35), which have displayed in these systems could hinder the free movement of
cycle stability as high as 93% after 10,000 cycles35. When the molecules, suppressing the undesirable dissolu-
individual redox motifs are anchored to each other by tion. Furthermore, the introduction of an inorganic
covalent bonds, the enlarged size of the molecules hin- solid electrolyte, a new electrolyte type investigated for
ders the solvation process, thus reducing their tendency lithium-ion battery research140,141, is an effective coun-
to be dissolved. As another example, anthraquinone, termeasure, precluding the dissolution of redox-active
which has high solubility and subsequently suffers from organic materials. Though this latter approach remains

www.nature.com/natrevmats

0123456789();:
Reviews

in the early stages of research and has been adopted Earth-abundant metal-ion organic rechargeable
only for organics 30, 42 and 43 (refs.142–144), its obvious batteries
merits, particularly the prevention of organic materials Concerns about the supply of lithium raw materials
dissolution140,145–147, may open up new opportunities for have motivated the exploration of battery chemistries
organic electrode materials. based on Earth-abundant metal ions such as sodium,
Capacity degradation is attributable not only to the potassium, magnesium, calcium and aluminium154–159.
dissolution of the organic electrode itself but also to side Batteries based on sodium ions and potassium ions have
reactions caused by the crossover of the dissolved species been studied the most extensively because the operating
in the electrochemical cell, leading to counter-electrode mechanism for the monovalent-ion system is expected
passivation or to the shuttle effect, in which dissolved to be similar to that of lithium-ion batteries. Moreover,
species move between the anode and the cathode with they share several important electrode material groups
repetitive oxidation and reduction148–150. The passing such as AxMO2 (A = Li, K or Na; M = transition metal),
of the dissolved redox-active organic materials to the allowing the reversible intercalation of not only lith-
counter-electrode can be inhibited through rational ium ions but also sodium and potassium ions160–164.
separator engineering. One simple approach is to coat Nevertheless, the use of larger ions such as Na+ (1.02 Å)
an additional layer such as graphene or Nafion on com- and K+ (1.38 Å), relative to Li+ (0.76 Å), often involves
monly used polypropylene or polyethylene separators a greater alternation of the crystal structure, which
to regulate the transport through the separator, because requires a larger volumetric change of the electrode and
additional layers allow small lithium ions and electro- more rapid degradation of the intercalation host165–171.
lytes to permeate but usually inhibit the permeation of Unlike the rigid structure of conventional inorganic
organic species. Enhanced cycle stability was achievable intercalation hosts, redox-active organic materials offer
for organic electrodes from organics 3, 13, 26 and 44 a much more flexible framework. Because organic
by employing a coated separator in comparison with a crystals are generally bound by weak van der Waals
bare separator under the same electrode and electro- forces, a space can be provided that is large enough to
lyte conditions97,151–153. The electrode with organic 44 accommodate various guest ions, compared with rigid
was capable of retaining 76% of the initial capacity after inorganic compounds, rendering redox-active organic
400 cycles through the employment of a Nafion-coated materials less susceptible to the size of charge-carrying
separator, whereas the capacity retention was only 64% ions6,172. This feature enables the stable insertion of larger
after 50 cycles with the bare polypropylene separator152. guest-ions such as Na+ and K+ (refs.30,173–184) in various
A new type of permselective separator has also been redox-active organic compounds, as presented in Fig. 6a
proposed, where novel membrane engineering could (and in Supplementary Fig. 3a, where references for
enable the precise control of the pore size, selectively each material are noted). For example, organic 13 can
blocking the redox-active organic molecules while reversibly store both Na+ and K+ ions, utilizing the two
transporting the ions in the electrolyte 149. In that carbonyl motifs in the structure181,182,185. Li (ref.186), Na
regard, a monolithic gel-type membrane with perm­ (ref.181) and K (ref.182) half-cells employing electrodes
selective properties was synthesized, taking advantage with organic 13 can deliver specific capacities of 127, 107
of the nanopores of metal–organic framework (MOF) and 105 mAh g−1, respectively, showing that the choice of
materials. The pores in MOF materials can be adjusted metal has an insignificant effect on the specific capacity.
to desired sizes and distributions, thus allowing the Furthermore, a respectable cycle and power performance
customized design of the membrane separator against is possible in Na and K cells, where around 94% and 87%
target organic molecules. A model organic compound, of the initial capacity can be delivered after 200 cycles and
quinone derivative organic 45, displayed a remark­ 1,000 cycles, respectively, while retaining 81 mAh g−1 at
able cycle stability with a retention of about 83% after 895 mA g−1 for the Na cell and 70 mAh g−1 at 4,475 mA g−1
2,000 cycles at a current density of 269 mA g −1 by for the K cell. These results offer a notable contrast to typ-
employing a MOF-gel separator based on zinc and ical inorganic hosts, whose intercalation capabilities vary
2-methylimidazole. substantially depending on the guest-ion species61,155.
Similarly, given its azo redox motif, organic 50 can store
Alternative energy storage platforms two equivalent ions in Li (ref.44), Na (ref.177) and K (ref.184)
The applicability of redox-active organic materials has half-cells. All half-cells presented decent and compa-
usually been examined in lithium-ion-battery configu- rable capacity retention, indicating that redox-active
rations. However, despite the extensive efforts described, organic materials are less susceptible to the type of guest
the application of redox-active organic materials still ion. Moreover, organic 49 delivered a high capacity of
presents unresolved issues, such as the inevitable sol- 232 mAh g−1, based on the full utilization of all four car-
ubility into the electrolyte and the low power density, bonyl motifs, and maintained 97% of the initial capacity
only improvable by substantially sacrificing the energy after 10,000 cycles, which is considered an outstanding
density. Efforts to produce battery platforms beyond stability for sodium-ion battery cathodes180.
lithium-ion batteries (the so-called post-lithium-ion Redox-active organic materials can also be suc-
batteries) have led to new opportunities for redox-active cessfully used in multivalent-ion batteries. Despite the
organic materials. Because of their diverse physical potential merits of multivalent-ion battery systems187,188,
and chemical configurations, these organic materials the lack of a suitable host for multivalent ions has
are more easily applied to these new platforms than remained one of the critical impediments. One of the
conventional transition-metal-based electrode materials. main issues is the severely sluggish diffusion kinetics

Nature Reviews | Materials

0123456789();:
Reviews

a 10–3 b 10–3
Sodium-ion batteries Magnesium-ion batteries
Potassium-ion batteries 49a Calcium-ion batteries 53

Capacity decay rate (% per cycle)

Capacity decay rate (% per cycle)


Aluminium-ion batteries
50a Inorganic 6a 46b
48
10–2 46a 10–2 52 38b
13c 49b 26b
Inorganic 3
47b Inorganic 8
50b
54 30c
Inorganic 5
10–1 13d 51 10–1 13e
1a
Inorganic 2
Inorganic 1
1d Inorganic 6b
Inorganic 4
Inorganic 7
47a
100 100
0 100 200 300 400 0 100 200 300
Initial specific capacity (mAh g–1) Initial specific capacity (mAh g–1)
c 15,000 d Energy density
Nonaqueous
Aqueous 260.4C
Specific power (W kg–1)

50C
10,000 Cost Power
40.7C merit density

5,000

7C
3.4C 1C
1C Cycle life Energy efficiency
0.34C
0
Organic 10 Organic 55 Organic 13 Organic 30 Vanadium Zn-Fe
ORFB 1 ORFB 2

Fig. 6 | Potential of employing redox-active organic materials for post-lithium-ion battery configurations.
a,b | Initial specific capacity and capacity decay rate of some organic materials (blue) and representative inorganic
materials (red) for monovalent-ion batteries (part a) and multivalent-ion batteries (part b). The initial specific capacity
is the reversible specific discharge capacity. We note that the y axis is inversed such that the points at the top-right
side of the plot indicate better performing materials. Inorganic 1, Na0.66Li0.18Mn0.71Ni0.21Co0.08O2; inorganic 2, Na3V(PO4)2;
inorganic 3, Na3MnZr(PO4)3; inorganic 4, K0.45MnO2; inorganic 5, K2Ni2TeO6; inorganic 6, Mo6S8; inorganic 7, Ti2S4;
inorganic 8, Na0.5VPO4.8F0.7. c | Specific power of selected organic batteries with aqueous and nonaqueous electrolyte.
All the specific powers are based on the C-rate and are presented with the highest reported value in each reference.
Organic 10, nonaqueous (ref.208) and aqueous (ref.207); organic 55, nonaqueous (ref.206) and aqueous (ref.205); organic 13,
nonaqueous (ref.182) and aqueous (ref.209); organic 30, nonaqueous (ref.22) and aqueous (ref.210). d | Performance and
cost evaluations for selected inorganic and organic redox-flow batteries222–226. The scores of each evaluation item were
normalized based on the maximum and minimum values of each item. Cost merits were calculated based on the inverse
value of the cost per kilogram of the active material. The cycle life was estimated from the capacity retention of each
system per year. ORFB 1, anolyte (1,8-bis(2-(2-(2-hydroxyethoxy)ethoxy)-ethoxy)anthracene-9,10-dione) and catholyte
(potassium ferrocyanide); ORFB 2, anolyte ((((9,10-dioxo-9,10-dihydroanthracene-2,6-diyl)bis(oxy))bis(propane-3,1-diyl))
bis(phosphonic acid)) and catholyte (potassium ferrocyanide).

typically observed in most intercalation hosts for multi- ion insertions in a magnesium half-cell150. Moreover,
valent ions188–192. As such, only a few inorganic host mate- 82% of the initial capacity is successfully retained even
rials have been reported to date, including Chevrel phases after 200 cycles, which is almost comparable to the
(Mo6S8, Mo6Se8) and vanadium oxides (V2O5, VO2), performance observed in the lithium cell22,150. A multi­
for magnesium- or aluminium-ion batteries6,193. valent aluminium complex ion can be reversibly stored
Although their development remains in its in the triangular macrocycle organic 52, which forms
infancy, several redox-active organic materials have a layered organic crystal structure. The insertion and
been identified as promising cathode materials for extraction of a cationic aluminium complex allows
multivalent-ion batteries, as presented in Fig. 6b (and in a reversible capacity of 75 mAh g–1 along with a cycl­
Supplementary Fig. 3b, where references for each mate- ability of up to 5,000 cycles196. The synthesis of the
rial are noted)150,194–200. Organic 30, containing four tetradiketone macrocycle organic 53 with a reduced
carbonyl redox motifs, can deliver a remarkably high benzene-ring-to-carbonyl-group ratio also leads to
capacity of 266 mAh g−1, which corresponds to two Mg2+ a high-capacity aluminium-ion battery cathode 197.

www.nature.com/natrevmats

0123456789();:
Reviews

Divalent ions (AlCl2+) are preferably stored over their As the field of aqueous organic rechargeable batter-
competing monovalent counterpart (AlCl2+), leading to a ies continues to grow, approaches that have been effec-
higher specific capacity of 237 mAh g−1. In that regard, tive in aqueous inorganic battery applications should
a truly multivalent aluminium-ion battery based on be explored. For example, the confinement of the water
organic active materials can be realized, unlike previously solvent within a polymer network reduces the activity of
reported systems relying on monovalent aluminium water214, and a spatially separated acid–alkaline dual elec-
complex ions for the main redox reaction. trolyte expands the electrochemical window of the full
cell by lowering the hydrogen-evolution reaction poten­
Aqueous organic rechargeable batteries tial and elevating the oxygen-evolution reaction potential
The electrolytes currently employed in lithium-ion based on their pH dependence 215. Applying such
batteries are flammable and toxic 201–203. Aqueous approaches to aqueous organic batteries should help to
rechargeable batteries can circumvent these problems overcome concerns about the low energy density.
while providing a much higher ionic conductivity
(about 1 S cm−1) than typical organic-based electrolytes Organic redox-flow batteries
(10−3–10−2 S cm−1)204. Thus, their use may help to com- Redox-flow batteries, which rely on the solution-based
pensate for the low kinetic performance of redox-active redox chemistry of active materials, have often employed
organic materials. In this regard, several attempts have soluble inorganic redox pairs such as V/V (ref. 216) ,
been made to achieve high power capabilities from Zn/Fe (ref.217) or Zn/Br (ref.218). Given the concerns
organic electrodes in aqueous battery systems, as depicted about their highly volatile price and toxicity, there has
in Fig. 6c (and in Supplementary Fig. 3c, where references been increasing interest in replacing these redox pairs
for each material are noted)22,182,205–210. For example, when with redox-active organic materials. Redox-active
organic 13 was used in an electrode in a sodium half-cell organic materials have advantages over their inor-
with an ethylene carbonate/diethyl carbonate electrolyte, ganic counterparts not only because of their potential
only 65% of the capacity could be delivered when the cost-effectiveness but also because of their chemical
current rate was increased from 0.07C to 7C (ref.181). In tunability, which enables target electrochemical perfor-
contrast, the same electrode could retain nearly 62.6% of mances through the modulation of solubility or molec-
the capacity when the current rate was increased to the ular size. In Fig. 6d, we compare performance and cost
extremely high value of 260.4 C from 1.6C when employ- metrics such as energy density, power density, energy
ing a 2 M zinc chloride (ZnCl2) aqueous electrolyte in efficiency, cycle life and cost of organic redox-flow bat-
a zinc half-cell209. The specific power of 12.7 kW per teries in reference to representative inorganic redox-flow
kilogram of material of this system was even higher than batteries. In the graph, the scores of each evaluation
that of conventional lithium-ion batteries (around 1 kW item were normalized based on the maximum and
per kilogram of material)9. Moreover, organic 30 with a minimum values of each item, which were collected
2 M zinc sulfate (ZnSO4) aqueous electrolyte exhibited from previous studies on inorganic vanadium-based
a specific power of 10.2 kW per kilogram of material at and Zn/Fe-based redox-flow batteries and organic
50C in a zinc half-cell, which far exceeds the specific 1,8-bis(2-(2-(2-hydroxyethoxy)ethoxy)-ethoxy)
power of 0.83 kW per kilogram of material obtained anthracene-9,10-dione/ferrocyanide (ORFB 1)-based
when the same organic was exploited in a nonaqueous and (((9,10-dioxo-9,10-dihydroanthracene-2,6-diyl)
system composed of 1 M lithium hexafluorophosphate bis(oxy))bis(propane-3,1-diyl))bis(phosphonic acid)/
and ethylene carbonate/diethyl carbonate22,210. ferrocyanide (ORFB 2)-based redox-flow batteries219–223.
Although aqueous electrolytes can offer extraordinary It is clear from Fig. 6d that the performance of organic
power capability for redox-active organic electrodes, redox-flow batteries can complement some aspects of
their introduction results in a narrow electro­chemical representative inorganic-based redox-flow batteries.
window (theoretically 1.23 V for water splitting). The In terms of energy density, tailoring the solubilizing
narrow electrochemical window would lead to not only group has been an effective strategy to increase the solu-
a limited deliverable energy density but also to various bility of the active material, thus enhancing the volumet-
side effects such as swelling of the cell by gas genera- ric specific capacity of organic redox-flow batteries224,225.
tion during repeated cycles211. The adoption of a highly Introducing a non-ionic polyethylene-glycol-based
concentrated electrolyte, a so-called water-in-salt elec- functional group to anthraquinone produced a
trolyte, is one of the strategies that may overcome these water-miscible 1,8-bis(2-(2-(2-hydroxyethoxy)ethoxy)-
shortcomings130,131,212. In the water-in-salt system, the ethoxy)anthracene-9,10-dione, which provided a
electrochemical window is widened because of its unique volumetric capacity of 80.4 Ah per litre and an energy
solvation structure, which leads to a reduction of the elec- density of 25.2 Wh per litre for ORFB 1 (ref.222). The
trochemical activity of water. Concentrated water-in-salt molecular structure of a representative ferrocyanide
electrolytes with 30 m potassium bis(fluorosulfonyl)imide catholyte can be redesigned by symmetry breaking of
effectively suppress the hydrogen evolution, extend- their ligands226, where the substitution of hydrophilic
ing the electrochemical window to nearly 4 V (ref.213). bipyridine ligands for cyanide ligands allows a higher
Accordingly, organic 13 can function as a stable anode solubility of 1.09 M in water, compared with 0.62 M for
in aqueous potassium batteries with improved cycling sodium ferrocyanide. Moreover, the redox potential
stability. Moreover, the dissolution of the redox-active for sodium ferrocyanide can be enhanced from 0.26 V
organic material could be substantially suppressed in (versus Ag/AgCl) to 0.65 V (versus Ag/AgCl) because of
water-in-salt electrolytes. the weaker interaction of the bipyridine ligand with the

Nature Reviews | Materials

0123456789();:
Reviews

Fe centre, relative to that of the cyanide ligand with Moreover, optimization procedures should be performed
the Fe centre. with a fundamental understanding of the compatibility
The low cycle life, which is mainly attributable to the of the organic electrodes with the electrolyte systems,
instability of redox-active organic materials at charged which has yet to be explored. The type of energy-storage
states, is the main bottleneck hindering the practical use platform that will most effectively exploit the benefits
of organic redox-flow batteries. As the molecules in solu- of redox-active organic materials should also be taken
tion are relatively free to interact with other elements, into account. For instance, organic rechargeable batter-
various side reactions, such as dimerization or nucleop- ies are generally disadvantageous in terms of volumetric
hilic attack from solvent molecules227,228, are likely to be energy or power density because of their intrinsically
more facile than for solid-state materials. These unde- low density229. To establish practically feasible organic
sirable reactions decrease the amount of active materials batteries, innovative electrode engineering along with
and subsequently shorten the lifetime of the redox-flow system design are warranted. An alternative strategy may
batteries. However, these limitations can be poten- be to establish target applications for which environmen-
tially overcome. For example, more positively charged tal friendliness and cost-effectiveness are crucial, such as
functional groups can inhibit radical dimerization in stationary energy storage systems, which usually store a
the charged state of both a ferrocene derivative and large amount of intermittent renewable energy, rather
organic 56 through electrostatic repulsion224. Redesign than applications for mobile devices.
of redox-active organic materials also mitigates the cross- Redox-active organic materials are regarded as
over known to greatly contribute to cycle degradation, eco-friendly materials, as they are generally produced
resulting in a marked capacity retention of 99.9943% per under mild synthetic conditions with low environ-
cycle and 99.967% per day. Furthermore, the study of the mental impact and can even be obtained from natural
degradation mechanism of carboxylated anthraquinone products32,33,230,231. Nonetheless, because the produc-
inspired a molecular redesign of carboxylated anthra­ tion and recycling processes of redox-active organic
quinone to phosphonated anthraquinone, restricting materials have not been thoroughly examined, it could
the carboxylate-mediated intramolecular nucleophilic be said that the sustainability of redox-active organic
substitution223. As a result, ORFB 2, with a (((9,10-dioxo- materials is simply conjecture. However, initial steps
9,10-dihydroanthracene-2,6-diyl)bis(oxy))bis(propane- towards recycling redox-active organic materials have
3,1-diyl))bis(phosphonic acid) anolyte and a potassium been taken to promote recyclability from the decompo-
ferrocyanide catholyte, demonstrated a remarkably low sition products232. Although this strategy must be further
capacity decay rate of 0.014% per day. These accom- developed in order to recycle such a battery completely,
plishments suggest that redox-active organic materials the possibility of recycling redox-active organic materials
are rapidly approaching the level needed for the practical improves the outlook of organic rechargeable batteries
application of redox-flow battery systems. for next-generation energy storage. Future efforts should
also examine how prospective organic electrode materi-
Perspective als can be produced with a minimal environmental foot-
Growing concerns about global environmental pol- print. Moreover, similar attention should be paid to the
lution have triggered the development of sustainable other components that constitute organic rechargeable
and eco-friendly battery chemistries. In that regard, batteries, considering that conventionally used battery
organic rechargeable batteries are considered promising components, such as the binder and electrolyte233–235, are
next-generation systems that could meet the demands often either not environmentally benign or not based
of this age. on sustainable chemistry. The appropriate selection or
Electrode-level performance metrics should be tailoring of redox-active organic materials may enable
more carefully assessed for the practical realization of the replacement of these components with environ-
organic rechargeable battery systems, eventually being mentally and economically more viable options. With
extended to cell-level performance metric evaluation. continued and concerted efforts to improve the perfor-
That is, the type or amount of electrolyte should also mance and sustainability of organic batteries, a greener
be considered, in order to attain a high gravimetric or rechargeable world is probably not too far off.
volumetric energy density for the cells, although such
performance-metric standards have not yet been set94,96. Published online xx xx xxxx

1. Wang, C., Chen, B., Yu, Y., Wang, Y. & Zhang, W. secondary battery category is more environmentally 11. Lakraychi, A. E., Dolhem, F., Vlad, A. & Becuwe, M.
Carbon footprint analysis of lithium ion secondary friendly and promising based on footprint family Organic negative electrode materials for metal-ion and
battery industry: two case studies from China. indicators? J. Clean. Prod. 276, 124244 (2020). molecular-ion batteries: progress and challenges from
J. Clean. Prod. 163, 241–251 (2017). 6. Poizot, P. et al. Opportunities and challenges for a molecular engineering perspective. Adv. Energy
2. Ambrose, H. & Kendall, A. Effects of battery chemistry organic electrodes in electrochemical energy storage. Mater. 11, 2101562 (2021).
and performance on the life cycle greenhouse gas Chem. Rev. 120, 6490–6557 (2020). 12. Stocker, T. et al. (eds) Climate Change 2013:
intensity of electric mobility. Transp. Res. D 47, 7. Esser, B. et al. A perspective on organic electrode The Physical Science Basis: Working Group I
182–194 (2016). materials and technologies for next generation Contribution To The Fifth Assessment Report
3. Kwak, W.-J. et al. Lithium–oxygen batteries and batteries. J. Power Sources 482, 228814 (2021). Of The Intergovernmental Panel On Climate Change
related systems: potential, status, and future. 8. Armand, M. & Tarascon, J.-M. Building better (Cambridge Univ. Press, 2014).
Chem. Rev. 120, 6626–6683 (2020). batteries. Nature 451, 652–657 (2008). 13. Larcher, D. & Tarascon, J. M. Towards greener and
4. Ma, L., Hendrickson, K. E., Wei, S. & Archer, L. A. 9. Lu, Y. & Chen, J. Prospects of organic electrode more sustainable batteries for electrical energy
Nanomaterials: science and applications in the materials for practical lithium batteries. storage. Nat. Chem. 7, 19–29 (2015).
lithium–sulfur battery. Nano Today 10, 315–338 Nat. Rev. Chem. 4, 127–142 (2020). 14. Sun, X., Luo, X., Zhang, Z., Meng, F. & Yang, J.
(2015). 10. Lee, S., Hong, J. & Kang, K. Redox-active organic Life cycle assessment of lithium nickel cobalt
5. Wang, L., Hu, J., Yu, Y., Huang, K. & Hu, Y. compounds for future sustainable energy storage manganese oxide (NCM) batteries for electric passenger
Lithium-air, lithium-sulfur, and sodium-ion, which system. Adv. Energy Mater. 10, 2001445 (2020). vehicles. J. Clean. Prod. 273, 123006 (2020).

www.nature.com/natrevmats

0123456789();:
Reviews

15. Rosental, M., Fröhlich, T. & Liebich, A. Life cycle ultrafast battery cathode. ACS Appl. Mater. Interf. 10, 67. Suguro, M., Iwasa, S., Kusachi, Y., Morioka, Y. &
assessment of carbon capture and utilization for 6346–6353 (2018). Nakahara, K. Cationic polymerization of poly
the production of large volume organic chemicals. 42. Peng, Z., Yi, X., Liu, Z., Shang, J. & Wang, D. (vinyl ether) bearing a TEMPO radical: a new
Front. Clim. 2, 9 (2020). Triphenylamine-based metal–organic frameworks cathode-active material for organic radical batteries.
16. Wernet, G., Conradt, S., Isenring, H. P., as cathode materials in lithium-ion batteries with Macromol. Rapid Commun. 28, 1929–1933 (2007).
Jiménez-González, C. & Hungerbühler, K. Life cycle coexistence of redox active sites, high working voltage, 68. Wang, S., Li, F., Easley, A. D. & Lutkenhaus, J. L.
assessment of fine chemical production: a case study and high rate stability. ACS Appl. Mater. Interf. 8, Real-time insight into the doping mechanism of
of pharmaceutical synthesis. Int. J. Life Cycle Assess. 14578–14585 (2016). redox-active organic radical polymers. Nat. Mater. 18,
15, 294–303 (2010). 43. Otteny, F. et al. Poly (vinylphenoxazine) as 69–75 (2019).
17. Wernet, G., Mutel, C., Hellweg, S. & Hungerbühler, K. fast-charging cathode material for organic batteries. 69. Oyaizu, K. & Nishide, H. Radical polymers for organic
The environmental importance of energy use in ACS Sustain. Chem. Eng. 8, 238–247 (2019). electronic devices: a radical departure from conjugated
chemical production. J. Ind. Ecol. 15, 96–107 (2011). 44. Luo, C. et al. Azo compounds as a family of polymers? Adv. Mater. 21, 2339–2344 (2009).
18. Wentker, M., Greenwood, M. & Leker, J. A bottom-up organic electrode materials for alkali-ion batteries. 70. Nigrey, P. J., MacInnes, D. Jr, Nairns, D. P.,
approach to lithium-ion battery cost modeling with a Proc. Natl Acad. Sci. USA 115, 2004–2009 (2018). MacDiarmid, A. G. & Heeger, A. J. Lightweight
focus on cathode active materials. Energies 12, 504 45. Singh, V. et al. Thiazole-linked covalent organic rechargeable storage batteries using polyacetylene,
(2019). framework promoting fast two-electron transfer (CH)x as the cathode-active material. J. Electrochem.
19. Battery Shipment And Installation For EV/ESS for lithium-organic batteries. Adv. Energy Mater. Soc. 128, 1651 (1981).
(SNE Research, 2021). 11, 2003735 (2021). 71. Kaneto, K., Yoshino, K. & Inuishi, Y. Characteristics of
20. Yang, Z. et al. Alkaline benzoquinone aqueous flow 46. Wu, C. et al. Azo-linked covalent triazine-based polythiophene battery. Jpn. J. Appl. Phys. 22, L567
battery for large-scale storage of electrical energy. framework as organic cathodes for ultrastable (1983).
Adv. Energy Mater. 8, 1702056 (2018). capacitor-type lithium-ion batteries. Energy Storage 72. MacDiarmid, A. G., Yang, L., Huang, W. & Humphrey, B.
21. Williams, D., Byrne, J. & Driscoll, J. A high energy Mater. 36, 347–354 (2021). Polyaniline: electrochemistry and application to
density lithium/dichloroisocyanuric acid battery 47. Lee, M. et al. Redox cofactor from biological energy rechargeable batteries. Synth. Met. 18, 393–398
system. J. Electrochem. Soc. 116, 2 (1969). transduction as molecularly tunable energy-storage (1987).
22. Liang, Y., Zhang, P. & Chen, J. Function-oriented compound. Angew. Chem. Int. Ed. 52, 8322–8328 73. Liu, Q. et al. Approaching the capacity limit of lithium
design of conjugated carbonyl compound electrodes (2013). cobalt oxide in lithium ion batteries via lanthanum and
for high energy lithium batteries. Chem. Sci. 4, 48. Peng, C. et al. Reversible multi-electron redox aluminium doping. Nat. Energy 3, 936–943 (2018).
1330–1337 (2013). chemistry of π-conjugated N-containing heteroaromatic 74. Gu, R. et al. Improved electrochemical performances
23. Liang, Y. et al. Universal quinone electrodes for long molecule-based organic cathodes. Nat. Energy 2, 1–9 of LiCoO2 at elevated voltage and temperature
cycle life aqueous rechargeable batteries. Nat. Mater. (2017). with an in situ formed spinel coating layer. ACS Appl.
16, 841–848 (2017). 49. Schon, T. B., Tilley, A. J., Bridges, C. R., Mater. Interf. 10, 31271–31279 (2018).
24. Chen, H. et al. From biomass to a renewable LixC6O6 Miltenburg, M. B. & Seferos, D. S. Bio-derived 75. Kim, U.-H. et al. Cation ordered Ni-rich layered
organic electrode for sustainable Li-ion batteries. polymers for sustainable lithium-ion batteries. cathode for ultra-long battery life. Energy Environ. Sci.
ChemSusChem 1, 348 (2008). Adv. Funct. Mater. 26, 6896–6903 (2016). 14, 1573–1583 (2021).
25. Chen, H. et al. Lithium salt of tetrahydroxybenzoquinone: 50. Hong, J. et al. Biologically inspired pteridine redox 76. Neudeck, S. et al. Molecular surface modification of
toward the development of a sustainable Li-ion battery. centres for rechargeable batteries. Nat. Commun. 5, NCM622 cathode material using organophosphates
J. Am. Chem. Soc. 131, 8984–8988 (2009). 5335 (2014). for improved Li-ion battery full-cells. ACS Appl. Mater.
26. Lee, M. et al. High-performance sodium–organic 51. Kim, J. et al. Biological nicotinamide cofactor as Interf. 10, 20487–20498 (2018).
battery by realizing four-sodium storage in disodium a redox-active motif for reversible electrochemical 77. Song, J. et al. Enhancement of the rate capability
rhodizonate. Nat. Energy 2, 861–868 (2017). energy storage. Angew. Chem. Int. Ed. 58, of LiFePO4 by a new highly graphitic carbon-coating
27. Armand, M. et al. Conjugated dicarboxylate anodes 16764–16769 (2019). method. ACS Appl. Mater. Interfaces 8, 15225–15231
for Li-ion batteries. Nat. Mater. 8, 120–125 (2009). 52. Fujihara, Y. et al. Electrical conductivity-relay between (2016).
28. Jouhara, A. et al. Raising the redox potential in organic charge-transfer and radical salts toward 78. Luo, C., Fan, X., Ma, Z., Gao, T. & Wang, C. Self-healing
carboxyphenolate-based positive organic materials conductive additive-free rechargeable battery. chemistry between organic material and binder for
via cation substitution. Nat. Commun. 9, 4401 (2018). ACS Appl. Mater. Interf. 12, 25748–25755 (2020). stable sodium-ion batteries. Chem 3, 1050–1062
29. Geng, J., Bonnet, J.-P., Renault, S., Dolhem, F. & 53. Hanyu, Y. & Honma, I. Rechargeable quasi-solid (2017).
Poizot, P. Evaluation of polyketones with N-cyclic state lithium battery with organic crystalline cathode. 79. Lu, Y. et al. Cyclohexanehexone with ultrahigh
structure as electrode material for electrochemical Sci. Rep. 2, 453 (2012). capacity as cathode materials for lithium-ion batteries.
energy storage: case of tetraketopiperazine unit. 54. Häupler, B. et al. PolyTCAQ in organic batteries: Angew. Chem. Int. Ed. 58, 7020–7024 (2019).
Energy Environ. Sci. 3, 1929–1933 (2010). enhanced capacity at constant cell potential using 80. Janek, J. & Zeier, W. G. A solid future for battery
30. Song, Z., Zhan, H. & Zhou, Y. Polyimides: promising two-electron-redox-reactions. J. Mater. Chem. A 2, development. Nat. Energy 1, 16141 (2016).
energy-storage materials. Angew. Chem. Int. Ed. 49, 8999–9001 (2014). 81. Kasnatscheew, J. et al. The truth about the 1st cycle
8444–8448 (2010). 55. Wang, J. et al. Conjugated sulfonamides as a class Coulombic efficiency of LiNi1/3Co1/3Mn1/3O2 (NCM)
31. Xing, Z. et al. A perylene anhydride crystal of organic lithium-ion positive electrodes. Nat. Mater. cathodes. Phys. Chem. Chem. Phys. 18, 3956–3965
as a reversible electrode for K-ion batteries. 20, 665–673 (2021). (2016).
Energy Storage Mater. 2, 63–68 (2016). 56. Speer, M. E. et al. Thianthrene-functionalized 82. Liang, Y., Zhang, P., Yang, S., Tao, Z. & Chen, J. Fused
32. Miroshnikov, M. et al. Bioderived molecular electrodes polynorbornenes as high-voltage materials for organic heteroaromatic organic compounds for high-power
for next-generation energy-storage materials. cathode-based dual-ion batteries. Chem. Commun. 51, electrodes of rechargeable lithium batteries. Adv.
ChemSusChem 13, 2186–2204 (2020). 15261–15264 (2015). Energy Mater. 3, 600–605 (2013).
33. Lee, B. et al. Exploiting biological systems: toward 57. Wild, A., Strumpf, M., Häupler, B., Hager, M. D. 83. Yao, M., Senoh, H., Araki, M., Sakai, T. &
eco-friendly and high-efficiency rechargeable batteries. & Schubert, U. S. All-organic battery composed Yasuda, K. Organic positive-electrode materials
Joule 2, 61–75 (2018). of thianthrene- and TCAQ-based polymers. based on dialkoxybenzoquinone derivatives for use in
34. Lee, S. et al. Charge-transfer complexes for high-power Adv. Energy Mater. 7, 1601415 (2017). rechargeable lithium batteries. ECS Trans. 28, 3 (2010).
organic rechargeable batteries. Energy Storage Mater. 58. Kim, J. et al. A p–n fusion strategy to design bipolar 84. Nishida, S., Yamamoto, Y., Takui, T. & Morita, Y.
20, 462–469 (2019). organic materials for high-energy-density symmetric Organic rechargeable batteries with tailored voltage
35. Kolek, M. et al. Ultra-high cycling stability of poly batteries. J. Mater. Chem. A 9, 14485–14494 (2021). and cycle performance. ChemSusChem 6, 794–797
(vinylphenothiazine) as a battery cathode material 59. Zhao, H. et al. Organic thiocarboxylate electrodes (2013).
resulting from π–π interactions. Energy Environ. Sci. for a room-temperature sodium-ion battery delivering 85. Lee, S., Kwon, J. E., Hong, J., Park, S. Y. & Kang, K.
10, 2334–2341 (2017). an ultrahigh capacity. Angew. Chem. Int. Ed. 56, The role of substituents in determining the redox
36. Lee, K. et al. Phenoxazine as a high-voltage p-type 15334–15338 (2017). potential of organic electrode materials in Li and
redox center for organic battery cathode materials: 60. Zhang, B. et al. Isometric thionated naphthalene Na rechargeable batteries: electronic effects vs.
small structural reorganization for faster charging diimides as organic cathodes for high capacity lithium substituent-Li/Na ionic interaction. J. Mater. Chem. A
and narrow operating voltage. Energy Environ. Sci. batteries. Chem. Mater. 32, 10575–10583 (2020). 7, 11438–11443 (2019).
13, 4142–4156 (2020). 61. Lee, S. et al. Recent progress in organic electrodes 86. Sieuw, L. et al. Through-space charge modulation
37. Lee, M. et al. Multi-electron redox phenazine for for Li and Na rechargeable batteries. Adv. Mater. 30, overriding substituent effect: rise of the redox
ready-to-charge organic batteries. Green. Chem. 19, 1704682 (2018). potential at 3.35 V in a lithium-phenolate
2980–2985 (2017). 62. Visco, S. J. & DeJonghe, L. C. Ionic conductivity of stereoelectronic isomer. Chem. Mater. 32,
38. Dai, G. et al. Manipulation of conjugation to stabilize organosulfur melts for advanced storage electrodes. 9996–10006 (2020).
N redox-active centers for the design of high-voltage J. Electrochem. Soc. 135, 2905 (1988). 87. Rambabu, D. et al. An electrically conducting Li-ion
organic battery cathode. Energy Storage Mater. 16, 63. Zhou, J. et al. Selenium-doped cathodes for metal–organic framework. J. Am. Chem. Soc. 143,
236–242 (2019). lithium–organosulfur batteries with greatly improved 11641–11650 (2021).
39. Huang, L. et al. π-Extended dihydrophenazine-based volumetric capacity and Coulombic efficiency. 88. Liang, Y. & Yao, Y. Positioning organic electrode
polymeric cathode material for high-performance Adv. Mater. 29, 1701294 (2017). materials in the battery landscape. Joule 2,
organic batteries. ACS Sustain. Chem. Eng. 8, 64. Visco, S., Liu, M. & De Jonghe, L. Ambient 1690–1706 (2018).
17868–17875 (2020). temperature high-rate lithium/organosulfur batteries. 89. Yang, H. et al. Molecular engineering of carbonyl
40. Peterson, B. M., Gannett, C. N., Melecio-Zambrano, L., J. Electrochem. Soc. 137, 1191 (1990). organic electrodes for rechargeable metal-ion batteries:
Fors, B. P. & Abruña, H. Effect of structural ordering 65. Shadike, Z. et al. Review on organosulfur materials fundamentals, recent advances, and challenges.
on the charge storage mechanism of p-type organic for rechargeable lithium batteries. Mater. Horiz. 8, Energy Environ. Sci. 14, 4228–4267 (2021).
electrode materials. ACS Appl. Mater. Interf. 13, 471–500 (2021). 90. Lee, Y. K. The effect of active material, conductive
7135–7141 (2021). 66. Nakahara, K. et al. Rechargeable batteries with organic additives, and binder in a cathode composite
41. Yamamoto, K., Suemasa, D., Masuda, K., Aita, K. & radical cathodes. Chem. Phys. Lett. 359, 351–354 electrode on battery performance. Energies 12,
Endo, T. Hyperbranched triphenylamine polymer for (2002). 658 (2019).

Nature Reviews | Materials

0123456789();:
Reviews

91. Zheng, H., Yang, R., Liu, G., Song, X. & Battaglia, V. S. 116. Kwon, J. E. et al. Triptycene-based quinone molecules 142. Hao, F. et al. High-energy all-solid-state organic–
Cooperation between active material, polymeric binder showing multi-electron redox reactions for large lithium batteries based on ceramic electrolytes.
and conductive carbon additive in lithium ion battery capacity and high energy organic cathode materials ACS Energy Lett. 6, 201–207 (2020).
cathode. J. Phys. Chem. C 116, 4875–4882 (2012). in Li-ion batteries. J. Mater. Chem. A 6, 3134–3140 143. Chi, X. et al. Tailored organic electrode material
92. Fédèle, L. et al. Mesoscale texturation of organic- (2018). compatible with sulfide electrolyte for stable all-solid-
based negative electrode material through in situ 117. Muench, S. et al. Polymer-based organic batteries. state sodium batteries. Angew. Chem. Int. Ed. 57,
proton reduction of conjugated carboxylic acid. Chem. Rev. 116, 9438–9484 (2016). 2630–2634 (2018).
Chem. Mater. 31, 6224–6230 (2019). 118. Otteny, F. et al. Unlocking full discharge capacities of 144. Luo, C. et al. Solid-state electrolyte anchored with a
93. Jouhara, A. et al. Playing with the p-doping mechanism poly (vinylphenothiazine) as battery cathode material carboxylated azo compound for all-solid-state lithium
to lower the carbon loading in n-type insertion organic by decreasing polymer mobility through cross-linking. batteries. Angew. Chem. Int. Ed. 57, 8567–8571
electrodes: first feasibility study with binder-free Adv. Energy Mater. 8, 1802151 (2018). (2018).
composite electrodes. J. Electrochem. Soc. 167, 119. Song, Z. et al. Polyanthraquinone as a reliable 145. Sun, Y.-K. Promising all-solid-state batteries for future
070540 (2020). organic electrode for stable and fast lithium storage. electric vehicles. ACS Energy Lett. 5, 3221–3223
94. Mao, M. et al. Electronic conductive inorganic Angew. Chem. Int. Ed. 54, 13947–13951 (2015). (2020).
cathodes promising high-energy organic batteries. 120. Zhao, Y. et al. Balance cathode-active and 146. Yang, X., Adair, K. R., Gao, X. & Sun, X. Recent
Adv. Mater. 33, 2005781 (2021). anode-active groups in one conjugated polymer advances and perspectives on thin electrolytes
95. Iordache, A. et al. From an enhanced understanding towards high-performance all-organic lithium-ion for high-energy-density solid-state lithium batteries.
to commercially viable electrodes: the case of PTCLi4 batteries. Nano Energy 86, 106055 (2021). Energy Environ. Sci. 14, 643–671 (2021).
as sustainable organic lithium-ion anode material. 121. Song, Z., Qian, Y., Zhang, T., Otani, M. & Zhou, H. 147. Cheng, X.-B., Zhao, C.-Z., Yao, Y.-X., Liu, H. &
Adv. Sustain. Syst. 1, 1600032 (2017). Poly(benzoquinonyl sulfide) as a high-energy organic Zhang, Q. Recent advances in energy chemistry
96. Molina, A. et al. Electrode engineering of redox-active cathode for rechargeable Li and Na batteries. Adv. Sci. between solid-state electrolyte and safe lithium-metal
conjugated microporous polymers for ultra-high 2, 1500124 (2015). anodes. Chem 5, 74–96 (2019).
areal capacity organic batteries. ACS Energy Lett. 5, 122. Song, Z., Zhan, H. & Zhou, Y. Anthraquinone based 148. Lau, V. W.-h, Moudrakovski, I., Yang, J., Zhang, J. &
2945–2953 (2020). polymer as high performance cathode material for Kang, Y.-M. Uncovering the shuttle effect in organic
97. Fang, C. et al. Immobilizing an organic electrode rechargeable lithium batteries. Chem. Commun. batteries and counter-strategies thereof: a case study
material through π–π interaction for high-performance https://doi.org/10.1039/b814515f (2009). of the N,N′-dimethylphenazine cathode. Angew. Chem.
Li-organic batteries. J. Mater. Chem. A 7, 123. Oyaizu, K., Choi, W. & Nishide, H. Functionalization Int. Ed. 59, 4023–4034 (2020).
22398–22404 (2019). of poly (4-chloromethylstyrene) with anthraquinone 149. Bai, S. et al. Permselective metal–organic framework
98. Michelbacher, C. J. et al. Enabling Fast Charging: pendants for organic anode-active materials. gel membrane enables long-life cycling of rechargeable
A Technology Gap Assessment (Idaho National Polym. Adv. Technol. 22, 1242–1247 (2011). organic batteries. Nat. Nanotechnol. 16, 77–84 (2021).
Laboratory, 2017). 124. Xu, W. et al. Factors affecting the battery performance 150. Dong, H. et al. High-power Mg batteries enabled by
99. Jenn, A., Clark-Sutton, K., Gallaher, M. & Petrusa, J. of anthraquinone-based organic cathode materials. heterogeneous enolization redox chemistry and weakly
Environmental impacts of extreme fast charging. J. Mater. Chem. 22, 4032–4039 (2012). coordinating electrolytes. Nat. Energy 5, 1043–1050
Environ. Res. Lett. 15, 094060 (2020). 125. Zarren, G., Nisar, B. & Sher, F. Synthesis of (2020).
100. Nakahara, K. et al. Cell properties for modified PTMA anthraquinone-based electroactive polymers: a critical 151. Sun, Y. et al. Functional separator with a lightweight
cathodes of organic radical batteries. J. Power Sources review. Mater. Today Sustain. 5, 100019 (2019). carbon-coating for stable, high-capacity organic
165, 398–402 (2007). 126. Zhou, Y. et al. Polyanthraquinone-based nanostructured lithium batteries. Chem. Eng. J. 418, 129404 (2021).
101. Nakahara, K. et al. Al-laminated film packaged organic electrode material capable of high-performance 152. Song, Z., Qian, Y., Otani, M. & Zhou, H. Stable Li–
radical battery for high-power applications. J. Power pseudocapacitive energy storage in aprotic electrolyte. organic batteries with nafion-based sandwich-type
Sources 163, 1110–1113 (2007). Nano Energy 15, 654–661 (2015). separators. Adv. Energy Mater. 6, 1501780 (2016).
102. Janoschka, T., Hager, M. D. & Schubert, U. S. 127. Zhao, L. et al. A novel polyquinone cathode material 153. Belanger, R. L. et al. Diffusion control of organic
Powering up the future: radical polymers for battery for rechargeable lithium batteries. J. Power Sources cathode materials in lithium metal battery. Sci. Rep. 9,
applications. Adv. Mater. 24, 6397–6409 (2012). 233, 23–27 (2013). 1213 (2019).
103. Rostro, L., Wong, S. H. & Boudouris, B. W. Solid state 128. Molina, A. et al. New anthraquinone-based conjugated 154. Yabuuchi, N. et al. P2-type Nax[Fe1/2Mn1/2]O2 made
electrical conductivity of radical polymers as a function microporous polymer cathode with ultrahigh specific from earth-abundant elements for rechargeable Na
of pendant group oxidation state. Macromolecules surface area for high-performance lithium-ion batteries. Nat. Mater. 11, 512–517 (2012).
47, 3713–3719 (2014). batteries. Adv. Funct. Mater. 30, 1908074 (2020). 155. Rajagopalan, R., Tang, Y., Ji, X., Jia, C. & Wang, H.
104. Joo, Y., Agarkar, V., Sung, S. H., Savoie, B. M. & 129. Wang, X., Zhou, J. & Tang, W. Poly (dithieno [3, 2-b: Advancements and challenges in potassium ion
Boudouris, B. W. A nonconjugated radical polymer 2′, 3′-d] pyrrole) twisting redox pendants enabling batteries: a comprehensive review. Adv. Funct. Mater.
glass with high electrical conductivity. Science 359, high current durability in all-organic proton battery. 30, 1909486 (2020).
1391–1395 (2018). Energy Storage Mater. 36, 1–9 (2021). 156. Kim, H. et al. Recent progress and perspective in
105. Wang, S. et al. Organic Li4C8H2O6 nanosheets for 130. Suo, L. et al. “Water-in-salt” electrolyte enables electrode materials for K-ion batteries. Adv. Energy
lithium-ion batteries. Nano Lett. 13, 4404–4409 high-voltage aqueous lithium-ion chemistries. Science Mater. 8, 1702384 (2018).
(2013). 350, 938–943 (2015). 157. Li, M. et al. Design strategies for nonaqueous
106. Zhuo, S. et al. Size control of zwitterionic polymer 131. Yamada, Y. et al. Hydrate-melt electrolytes for multivalent-ion and monovalent-ion battery anodes.
micro/nanospheres and its dependence on sodium high-energy-density aqueous batteries. Nat. Energy 1, Nat. Rev. Mater. 5, 276–294 (2020).
storage. Nanoscale Horiz. 4, 1092–1098 (2019). 16129 (2016). 158. Liang, Y., Dong, H., Aurbach, D. & Yao, Y. Current
107. Wang, Y. & Weng, G. J. in Micromechanics And 132. Suo, L., Hu, Y.-S., Li, H., Armand, M. & Chen, L. status and future directions of multivalent metal-ion
Nanomechanics Of Composite Solids 123–156 A new class of solvent-in-salt electrolyte for high-energy batteries. Nat. Energy 5, 646–656 (2020).
(Springer, 2018). rechargeable metallic lithium batteries. Nat. Commun. 159. Elia, G. A. et al. An overview and future perspectives
108. Zhao, S. et al. Nanoengineering of advanced 4, 1481 (2013). of aluminum batteries. Adv. Mater. 28, 7564–7579
carbon materials for sodium-ion batteries. Small 17, 133. Ma, T., Liu, L., Wang, J., Lu, Y. & Chen, J. Charge (2016).
e2007431 (2021). storage mechanism and structural evolution of 160. Delmas, C., Braconnier, J.-J., Fouassier, C. &
109. Kim, H. W. et al. Binder-free organic cathode viologen crystals as the cathode of lithium batteries. Hagenmuller, P. Electrochemical intercalation of
based on nitroxide radical polymer-functionalized Angew. Chem. Int. Ed. 59, 11533–11539 (2020). sodium in NaxCoO2 bronzes. Solid. State Ion. 3-4,
carbon nanotubes and gel polymer electrolyte for 134. Zhang, X., Zhou, W., Zhang, M., Yang, Z. & Huang, W. 165–169 (1981).
high-performance sodium organic polymer batteries. Superior performance for lithium-ion battery with 161. Kim, H. et al. K-ion batteries based on a P2-type
J. Mater. Chem. A 8, 17980–17986 (2020). organic cathode and ionic liquid electrolyte. K0.6CoO2 cathode. Adv. Energy Mater. 7, 1700098
110. Xie, J., Wang, Z., Xu, Z. J. & Zhang, Q. Toward J. Energy Chem. 52, 28–32 (2021). (2017).
a high-performance all-plastic full battery with 135. Qin, J. et al. A metal-free battery with pure ionic liquid 162. Reed, J., Ceder, G. & Van Der Ven, A. Layered-to-spinel
a’single organic polymer as both cathode and anode. electrolyte. iScience 15, 16–27 (2019). phase transition in LixMnO2. Electrochem. Solid.
Adv. Energy Mater. 8, 1703509 (2018). 136. Fang, Y.-B., Zheng, W., Li, L. & Yuan, W.-H. An State Lett. 4, A78 (2001).
111. Lin, C.-H., Lee, J.-T., Yang, D.-R., Chen, H.-W. & ultrahigh rate ionic liquid dual-ion battery based on a 163. Ma, X., Chen, H. & Ceder, G. Electrochemical
Wu, S.-T. Nitroxide radical polymer/carbon-nanotube- poly (anthraquinonyl sulfide) anode. ACS Appl. Energy properties of monoclinic NaMnO2. J. Electrochem.
array electrodes with improved C-rate performance in Mater. 3, 12276–12283 (2020). Soc. 158, A1307 (2011).
organic radical batteries. RSC Adv. 5, 33044–33048 137. Huang, W. et al. Calix[6]quinone as high-performance 164. Kim, H. et al. Investigation of potassium storage in
(2015). cathode for lithium-ion battery. Sci. China Mater. 63, layered P3-type K0.5MnO2 cathode. Adv. Mater. 29,
112. Zhao, Q., Wang, J., Chen, C., Ma, T. & Chen, J. 339–346 (2020). 1702480 (2017).
Nanostructured organic electrode materials grown on 138. Huang, W. et al. Synthesis and application of calix [6] 165. Yabuuchi, N., Kubota, K., Dahbi, M. & Komaba, S.
graphene with covalent-bond interaction for high-rate quinone as a high-capacity organic cathode for Research development on sodium-ion batteries.
and ultra-long-life lithium-ion batteries. Nano Res. 10, plastic crystal electrolyte-based lithium-ion batteries. Chem. Rev. 114, 11636–11682 (2014).
4245–4255 (2017). Energy Storage Mater. 26, 465–471 (2020). 166. Pramudita, J. C., Sehrawat, D., Goonetilleke, D. &
113. Pender, J. P. et al. Electrode degradation in lithium-ion 139. Sun, H. et al. High-performance organic lithium-ion Sharma, N. An initial review of the status of electrode
batteries. ACS Nano 14, 1243–1295 (2020). battery with plastic crystal electrolyte. Org. Electron. materials for potassium-ion batteries. Adv. Energy
114. Rodríguez-Pérez, I. A. et al. Mg-ion battery electrode: 87, 105966 (2020). Mater. 7, 1602911 (2017).
an organic solid’s herringbone structure squeezed 140. Zhao, Q., Stalin, S., Zhao, C.-Z. & Archer, L. A. 167. Guo, S. et al. A layered P2- and O3-type composite
upon Mg-ion insertion. J. Am. Chem. Soc. 139, Designing solid-state electrolytes for safe, energy-dense as a high-energy cathode for rechargeable sodium-ion
13031–13037 (2017). batteries. Nat. Rev. Mater. 5, 229–252 (2020). batteries. Angew. Chem. Int. Ed. 54, 5894–5899 (2015).
115. Zhang, K., Xie, Y., Monteiro, M. J. & Jia, Z. 141. Cao, S. et al. Modeling, preparation, and elemental 168. Kim, J., Yoon, G., Kim, H., Park, Y.-U. & Kang, K.
Triazole-enabled small TEMPO cathodes for doping of Li7La3Zr2O12 garnet-type solid electrolytes: Na3V(PO4)2: a new layered-type cathode material with
lithium-organic batteries. Energy Storage Mater. 35, a review. J. Korean Ceram. Soc. 56, 111–129 high water stability and power capability for Na-ion
122–129 (2021). (2019). batteries. Chem. Mater. 30, 3683–3689 (2018).

www.nature.com/natrevmats

0123456789();:
Reviews

169. Gao, H. et al. Na3MnZr(PO4)3: a high-voltage 197. Yoo, D.-J., Heeney, M., Glöcklhofer, F. & Choi, J. W. batteries. Energy Storage Mater. 42, 185–192
cathode for sodium batteries. J. Am. Chem. Soc. 140, Tetradiketone macrocycle for divalent aluminium ion (2021).
18192–18199 (2018). batteries. Nat. Commun. 12, 2386 (2021). 226. Li, X. et al. Symmetry-breaking design of an organic iron
170. Liu, C.-L., Luo, S.-H., Huang, H.-B., Zhai, Y.-C. & 198. Walter, M., Kravchyk, K. V., Böfer, C., Widmer, R. & complex catholyte for a long cyclability aqueous organic
Wang, Z.-W. Layered potassium-deficient P2- and Kovalenko, M. V. Polypyrenes as high-performance redox flow battery. Nat. Energy 6, 873–881 (2021).
P3-type cathode materials KxMnO2 for K-ion batteries. cathode materials for aluminum batteries. Adv. Mater. 227. Kwabi, D. G., Ji, Y. & Aziz, M. J. Electrolyte lifetime in
Chem. Eng. J. 356, 53–59 (2019). 30, 1705644 (2018). aqueous organic redox flow batteries: a critical review.
171. Masese, T. et al. Rechargeable potassium-ion batteries 199. Han, C., Li, H., Li, Y., Zhu, J. & Zhi, C. Proton-assisted Chem. Rev. 120, 6467–6489 (2020).
with honeycomb-layered tellurates as high voltage calcium-ion storage in aromatic organic molecular 228. Feng, R. et al. Reversible ketone hydrogenation and
cathodes and fast potassium-ion conductors. crystal with coplanar stacked structure. Nat. Commun. dehydrogenation for aqueous organic redox flow
Nat. Commun. 9, 3823 (2018). 12, 2400 (2021). batteries. Science 372, 836–840 (2021).
172. Xie, J. & Zhang, Q. Recent Progress in multivalent 200. Gheytani, S. et al. An aqueous Ca-ion battery. Adv. Sci. 229. Xie, J. & Lu, Y.-C. Towards practical organic batteries.
metal (Mg, Zn, Ca, and Al) and metal-ion rechargeable 4, 1700465 (2017). Nat. Mater. 20, 581–583 (2021).
batteries with organic materials as promising 201. Liu, Y. et al. Cathode design for aqueous rechargeable 230. Hu, P. et al. Renewable-biomolecule-based full lithium-ion
electrodes. Small 15, 1805061 (2019). multivalent ion batteries: challenges and opportunities. batteries. Adv. Mater. 28, 3486–3492 (2016).
173. Jian, Z., Liang, Y., Rodríguez-Pérez, I. A., Yao, Y. Adv. Funct. Mater. 31, 2010445 (2021). 231. Sun, T. et al. A biodegradable polydopamine-derived
& Ji, X. Poly (anthraquinonyl sulfide) cathode for 202. Kim, H. et al. Aqueous rechargeable Li and Na ion electrode material for high-capacity and long-life
potassium-ion batteries. Electrochem. Commun. 71, batteries. Chem. Rev. 114, 11788–11827 (2014). lithium-ion and sodium-ion batteries. Angew. Chem.
5–8 (2016). 203. Huang, J. et al. Recent progress of rechargeable Int. Ed. 55, 10662–10666 (2016).
174. Zhang, Y. et al. Dispersion–assembly approach batteries using mild aqueous electrolytes. 232. Nguyen, T. P. et al. Polypeptide organic radical
to synthesize three-dimensional graphene/polymer Small Methods 3, 1800272 (2019). batteries. Nature 593, 61–66 (2021).
composite aerogel as a powerful organic cathode 204. Demir-Cakan, R., Palacin, M. R. & Croguennec, L. 233. Wang, C. et al. A selectively permeable membrane for
for rechargeable Li and Na batteries. ACS Appl. Rechargeable aqueous electrolyte batteries: enhancing cyclability of organic sodium-ion batteries.
Mater. Interf. 9, 15549–15556 (2017). from univalent to multivalent cation chemistry. Adv. Mater. 28, 9182–9187 (2016).
175. Zhao, Q. et al. Oxocarbon salts for fast rechargeable J. Mater. Chem. A 7, 20519–20539 (2019). 234. Patil, N. et al. An ultrahigh performance zinc-organic
batteries. Angew. Chem. Int. Ed. 55, 12528–12532 205. Wang, Y. et al. Binding zinc ions by carboxyl groups from battery using poly (catechol) cathode in Zn(TFSI)2-
(2016). adjacent molecules toward long-life aqueous zinc–organic based concentrated aqueous electrolytes. Adv. Energy
176. Slesarenko, A. et al. New tetraazapentacene-based batteries. Adv. Mater. 32, 2000338 (2020). Mater. 11, 2100939 (2021).
redox-active material as a promising high-capacity 206. Ma, T., Zhao, Q., Wang, J., Pan, Z. & Chen, J. A sulfur 235. Tie, Z. & Niu, Z. Design strategies for high-performance
organic cathode for lithium and potassium batteries. heterocyclic quinone cathode and a multifunctional aqueous zn/organic batteries. Angew. Chem. Int. Ed.
J. Power Sources 435, 226724 (2019). binder for a high-performance rechargeable lithium-ion 59, 21293–21303 (2020).
177. Luo, C. et al. Reversible redox chemistry of azo battery. Angew. Chem. Int. Ed. 55, 6428–6432 (2016). 236. Forster, P. V. et al. in Climate Change 2007: The
compounds for sodium-ion batteries. Angew. Chem. 207. Wang, Q., Liu, Y. & Chen, P. Phenazine-based Physical Science Basis. Contribution Of Working
Int. Ed. 57, 2879–2883 (2018). organic cathode for aqueous zinc secondary batteries. Group I To The Fourth Assessment Report Of
178. Zhang, Y. et al. Enhanced cycle performance of J. Power Sources 468, 228401 (2020). The Intergovernmental Panel On Climate Change
polyimide cathode using a quasi-solid-state electrolyte. 208. Tian, B. et al. Amino group enhanced phenazine (eds Solomon, S. et al.) (Cambridge Univ. Press, 2007).
J. Phys. Chem. C 122, 22294–22300 (2018). derivatives as electrode materials for lithium storage. 237. Myhre, G. D. et al. in Climate Change 2013: The
179. Tang, W. et al. Highly stable and high Chem. Commun. 53, 2914–2917 (2017). Physical Science Basis. Contribution Of Working
rate-performance Na-ion batteries using polyanionic 209. Zhang, H., Fang, Y., Yang, F., Liu, X. & Lu, X. Group I To The Fifth Assessment Report Of The
anthraquinone as the organic cathode. ChemSusChem Aromatic organic molecular crystal with enhanced Intergovernmental Panel On Climate Change
12, 2181–2185 (2019). π–π stacking interaction for ultrafast Zn-ion storage. (eds Stocker, T. F. et al.) (Cambridge Univ. Press, 2013).
180. Tang, M. et al. Tailoring π-conjugated systems: Energy Environ. Sci. 13, 2515–2523 (2020).
from π-π stacking to high-rate-performance organic 210. Guo, Z. et al. An environmentally friendly and flexible Acknowledgements
cathodes. Chem 4, 2600–2614 (2018). aqueous zinc battery using an organic cathode. This study was supported by the Creative Materials Discovery
181. Luo, W., Allen, M., Raju, V. & Ji, X. An organic pigment Angew. Chem. Int. Ed. 57, 11737–11741 (2018). Program through the National Research Foundation of Korea
as a high-performance cathode for sodium-ion 211. Han, C., Zhu, J., Zhi, C. & Li, H. The rise of aqueous (NRF), funded by the Ministry of Science, ICT and Future
batteries. Adv. Energy Mater. 4, 1400554 (2014). rechargeable batteries with organic electrode materials. Planning (NRF-2017M3D1A1039553). This work was also
182. Fan, L., Ma, R., Wang, J., Yang, H. & Lu, B. An ultrafast J. Mater. Chem. A 8, 15479–15512 (2020). supported by the Center for Nanoparticle Research at
and highly stable potassium–organic battery. Adv. 212. Suo, L. et al. Advanced high-voltage aqueous lithium-ion Institute for Basic Science (IBS) (IBS-R006-A2).
Mater. 30, 1805486 (2018). battery enabled by “water-in-bisalt” electrolyte.
183. Tang, M. et al. An organic cathode with high capacities Angew. Chem. Int. Ed. 55, 7136–7141 (2016). Author contributions
for fast-charge potassium-ion batteries. J. Mater. 213. Chen, H. et al. Use of a water-in-salt electrolyte to J.K. led the process, and prepared the content in the
Chem. A 7, 486–492 (2019). avoid organic material dissolution and enhance ‘Applications in rechargeable batteries’ section and
184. Liang, Y. et al. An organic anode for high temperature the kinetics of aqueous potassium ion batteries. the ‘Perspective’ section. Y. Kim prepared the content of the
potassium-ion batteries. Adv. Energy Mater. 9, Sustain. Energy Fuels 4, 128–131 (2020). ‘Alternative energy storage platforms’ section. J.Y. prepared
1802986 (2019). 214. Xie, J., Liang, Z. & Lu, Y.-C. Molecular crowding the content of the ‘Introduction’, ‘Assessment of redox-active
185. Chen, Y. et al. Organic electrode for non-aqueous electrolytes for high-voltage aqueous batteries. organics’ and ‘Redox-active organic materials’ sections. J.K.,
potassium-ion batteries. Nano Energy 18, 205–211 Nat. Mater. 19, 1006–1011 (2020). Y. Kim and J.Y. revised and edited the manuscript before pub-
(2015). 215. Liu, C., Chi, X., Han, Q. & Liu, Y. A high energy density lication. All authors participated in researching data and
186. Sharma, P., Damien, D., Nagarajan, K., Shaijumon, M. M. aqueous battery achieved by dual dissolution/deposition substantial discussion about organization of manuscript.
& Hariharan, M. Perylene-polyimide-based organic reactions separated in acid-alkaline electrolyte.
electrode materials for rechargeable lithium batteries. Adv. Energy Mater. 10, 1903589 (2020).
Competing interests
The authors declare no competing interests.
J. Phys. Chem. Lett. 4, 3192–3197 (2013). 216. Ding, C., Zhang, H., Li, X., Liu, T. & Xing, F. Vanadium
187. Lu, Y., Tu, Z. & Archer, L. A. Stable lithium flow battery for energy storage: prospects and
Peer review information
electrodeposition in liquid and nanoporous solid challenges. J. Phys. Chem. Lett. 4, 1281–1294 (2013).
Nature Reviews Materials thanks Stefan Freunberger,
electrolytes. Nat. Mater. 13, 961–969 (2014). 217. Gong, K. et al. A zinc–iron redox-flow battery
Alexandru Vlad and the other, anonymous, reviewer(s) for
188. Qin, K., Huang, J., Holguin, K. & Luo, C. Recent under $100 per kW h of system capital cost.
their contribution to the peer review of this work.
advances in developing organic electrode materials for Energy Environ. Sci. 8, 2941–2945 (2015).
multivalent rechargeable batteries. Energy Environ. Sci. 218. Lai, Q., Zhang, H., Li, X., Zhang, L. & Cheng, Y. A novel Publisher’s note
13, 3950–3992 (2020). single flow zinc–bromine battery with improved Springer Nature remains neutral with regard to jurisdictional
189. Aurbach, D. et al. Prototype systems for rechargeable energy density. J. Power Sources 235, 1–4 (2013). claims in published maps and institutional affiliations.
magnesium batteries. Nature 407, 724–727 (2000). 219. Sánchez-Díez, E. et al. Redox flow batteries: status
190. Sun, X. et al. A high capacity thiospinel cathode for Mg and perspective towards sustainable stationary energy Springer Nature or its licensor holds exclusive rights to this
batteries. Energy Environ. Sci. 9, 2273–2277 (2016). storage. J. Power Sources 481, 228804 (2021). article under a publishing agreement with the author(s) or
191. Geng, L., Lv, G., Xing, X. & Guo, J. Reversible 220. Rodby, K. E. et al. Assessing the levelized cost of other rightsholder(s); author self-archiving of the accepted
electrochemical intercalation of aluminum in Mo6S8. vanadium redox flow batteries with capacity fade and manuscript version of this article is solely governed by the
Chem. Mater. 27, 4926–4929 (2015). rebalancing. J. Power Sources 460, 227958 (2020). terms of such publishing agreement and applicable law.
192. Xu, Z.-L. et al. A new high-voltage calcium intercalation 221. Yuan, Z., Duan, Y., Liu, T., Zhang, H. & Li, X.
host for ultra-stable and high-power calcium Toward a low-cost alkaline zinc-iron flow battery with Supplementary information
rechargeable batteries. Nat. Commun. 12, 3369 (2021). a polybenzimidazole custom membrane for stationary The online version contains supplementary material available
193. Wu, F., Yang, H., Bai, Y. & Wu, C. Paving the path energy storage. iScience 3, 40–49 (2018). at https://doi.org/10.1038/s41578-022-00478-1.
toward reliable cathode materials for aluminum-ion 222. Jin, S. et al. A water-miscible quinone flow battery
batteries. Adv. Mater. 31, 1806510 (2019). with high volumetric capacity and energy density.
194. Pan, B. et al. Polyanthraquinone-based organic ACS Energy Lett. 4, 1342–1348 (2019).
Related links
London metal exchange, 2021: https://www.lme.com/en/
cathode for high-performance rechargeable 223. Ji, Y. et al. A phosphonate-functionalized quinone redox
Metals/EV/LME-Cobalt#Trading+day+summary
magnesium-ion batteries. Adv. Energy Mater. 6, flow battery at near-neutral pH with record capacity
shanghai metal market, 2021: https://www.metal.com/price/
1600140 (2016). retention rate. Adv. Energy Mater. 9, 1900039 (2019).
New%20Energy/Ternary-precursor-material
195. Cui, L. et al. Salt-controlled dissolution in pigment 224. Beh, E. S. et al. A neutral pH aqueous organic–
United states Geological survey (UsGs): https://minerals.
cathode for high-capacity and long-life magnesium organometallic redox flow battery with extremely high
usgs.gov/minerals/pubs/mcs/2020/%20mcs2021.pdf
organic batteries. Nano Energy 65, 103902 (2019). capacity retention. ACS Energy Lett. 2, 639–644 (2017).
196. Kim, D. J. et al. Rechargeable aluminium organic 225. Kwon, G. et al. Highly persistent triphenylamine-
batteries. Nat. Energy 4, 51–59 (2019). based catholyte for durable organic redox flow © Springer Nature Limited 2022

Nature Reviews | Materials

0123456789();:

You might also like